MathIsimple
π
θ
Δ
ε
RA-1
6-10 hours
Advanced

Measure Foundations & Outer Measure

Master the construction of outer measure via rectangle coverings, understand the Carathéodory measurability criterion, and explore the fundamental properties of measurable sets that form the foundation of modern measure theory.

Learning Objectives
By the end of this course, you will be able to:
1
Define outer measure using countable rectangle coverings and understand its fundamental properties
2
Apply the Carathéodory criterion to determine measurability of sets
3
Prove that measurable sets form a σ-algebra and that outer measure is countably additive on measurable sets
4
Understand regularity properties: approximation by open and closed sets
5
Identify null sets and understand their role in measure theory
6
Connect outer measure to geometric intuition via rectangle coverings

Prerequisites

Before starting this course, you should be familiar with:

  • Basic set theory: unions, intersections, complements, countable collections
  • Topology of ℝⁿ: open sets, closed sets, compact sets
  • Sequences and limits in ℝⁿ
  • Basic properties of rectangles and their volumes

Core Concepts

Fundamental definitions and properties

Definition 1.1: Outer Measure

Let ERnE \subseteq \mathbb{R}^n. The outer measure of EE, denoted m(E)m^*(E), is defined as:

m(E)=inf{j=1Qj:Ej=1Qj, where Qj are open rectangles}m^*(E) = \inf\left\{\sum_{j=1}^{\infty} |Q_j| : E \subseteq \bigcup_{j=1}^{\infty} Q_j, \text{ where } Q_j \text{ are open rectangles}\right\}

where Qj|Q_j| denotes the volume (Lebesgue measure) of the rectangle QjQ_j.

Note: We take the infimum over all possible countable coverings of EE by open rectangles. If no such covering exists (which cannot happen), we define m(E)=m^*(E) = \infty.

Definition 1.2: Measurable Set (Carathéodory Criterion)

A set ERnE \subseteq \mathbb{R}^n is said to be measurable (in the sense of Carathéodory) if for every set ARnA \subseteq \mathbb{R}^n, we have:

m(A)=m(AE)+m(AEc)m^*(A) = m^*(A \cap E) + m^*(A \cap E^c)

where Ec=RnEE^c = \mathbb{R}^n \setminus E denotes the complement of EE.

Definition 1.3: Null Set

A set ERnE \subseteq \mathbb{R}^n is called a null set (or a set of measure zero) if m(E)=0m^*(E) = 0.

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

Theorem 1.1: Properties of Outer Measure

Outer measure has the following fundamental properties:

  1. Non-negativity: m(E)0m^*(E) \geq 0 for all EE, and m()=0m^*(\emptyset) = 0.
  2. Monotonicity: If EFE \subseteq F, then m(E)m(F)m^*(E) \leq m^*(F).
  3. Countable Subadditivity: For any countable collection {Ej}j=1\{E_j\}_{j=1}^{\infty} of sets,
m(j=1Ej)j=1m(Ej)m^*\left(\bigcup_{j=1}^{\infty} E_j\right) \leq \sum_{j=1}^{\infty} m^*(E_j)
  • Translation Invariance: For any xRnx \in \mathbb{R}^n, m(E+x)=m(E)m^*(E + x) = m^*(E), where E+x={y+x:yE}E + x = \{y + x : y \in E\}.
  • Proof of Theorem 1.1:

    Proof:

    (1) Non-negativity follows immediately from the definition, as volumes are non-negative and the infimum of non-negative numbers is non-negative. The empty set can be covered by a single rectangle of arbitrarily small volume, so m()=0m^*(\emptyset) = 0.

    (2) Monotonicity: If EFE \subseteq F, then any covering of FF is also a covering of EE. Therefore, the infimum over coverings of EE is at most the infimum over coverings of FF.

    (3) Countable Subadditivity: Given ϵ>0\epsilon > 0, for each jj, choose a covering {Qj,k}k=1\{Q_{j,k}\}_{k=1}^{\infty} of EjE_j such that

    k=1Qj,k<m(Ej)+ϵ2j\sum_{k=1}^{\infty} |Q_{j,k}| < m^*(E_j) + \frac{\epsilon}{2^j}

    Then {Qj,k:j,kN}\{Q_{j,k} : j, k \in \mathbb{N}\} is a countable covering of j=1Ej\bigcup_{j=1}^{\infty} E_j, and

    m(j=1Ej)j,kQj,k<j=1m(Ej)+ϵm^*\left(\bigcup_{j=1}^{\infty} E_j\right) \leq \sum_{j,k} |Q_{j,k}| < \sum_{j=1}^{\infty} m^*(E_j) + \epsilon

    Since ϵ>0\epsilon > 0 was arbitrary, the result follows.

    (4) Translation Invariance: If {Qj}\{Q_j\} covers EE, then {Qj+x}\{Q_j + x\} covers E+xE + x, and Qj+x=Qj|Q_j + x| = |Q_j|. The result follows from the definition. ∎

    Theorem 1.2: Measurable Sets Form a σ-Algebra

    The collection M\mathcal{M} of all measurable subsets of Rn\mathbb{R}^n forms a σ-algebra. That is:

    1. RnM\mathbb{R}^n \in \mathcal{M}
    2. If EME \in \mathcal{M}, then EcME^c \in \mathcal{M}
    3. If E1,E2,ME_1, E_2, \ldots \in \mathcal{M}, then j=1EjM\bigcup_{j=1}^{\infty} E_j \in \mathcal{M}
    Proof of Theorem 1.2:

    Proof:

    (1) For any AA, we have m(A)=m(ARn)+m(A)=m(A)+0m^*(A) = m^*(A \cap \mathbb{R}^n) + m^*(A \cap \emptyset) = m^*(A) + 0, so Rn\mathbb{R}^n is measurable.

    (2) The measurability condition is symmetric in EE and EcE^c, so if EE is measurable, so is EcE^c.

    (3) For countable unions, we use induction and the fact that finite unions of measurable sets are measurable (which follows from De Morgan's laws and the complement property). The countable case then follows from continuity properties. ∎

    Theorem 1.3: Countable Additivity on Measurable Sets

    If E1,E2,E_1, E_2, \ldots are disjoint measurable sets, then:

    m(j=1Ej)=j=1m(Ej)m^*\left(\bigcup_{j=1}^{\infty} E_j\right) = \sum_{j=1}^{\infty} m^*(E_j)

    This property is called countable additivity.

    Proof of Theorem 1.3:

    Proof:

    By subadditivity, we already have m(Ej)m(Ej)m^*(\bigcup E_j) \leq \sum m^*(E_j). For the reverse inequality, we use the measurability of the sets and induction. For finite unions, measurability gives additivity. The countable case follows by taking limits. ∎

    Corollary 1.1: Lebesgue Measure

    The restriction of outer measure to the σ-algebra M\mathcal{M} of measurable sets is called Lebesgue measure, denoted mm. That is, for measurable EE, we define:

    m(E)=m(E)m(E) = m^*(E)

    Lebesgue measure is countably additive on measurable sets.

    Example 1.1: Outer Measure of a Point

    Problem: Compute the outer measure of a single point {a}\{a\} in Rn\mathbb{R}^n.

    Solution:

    For any ϵ>0\epsilon > 0, we can cover the point {a}\{a\} by a single open rectangle QQ centered at aa with volume Q<ϵ|Q| < \epsilon.

    Therefore, m({a})ϵm^*(\{a\}) \leq \epsilon for all ϵ>0\epsilon > 0, which implies m({a})=0m^*(\{a\}) = 0.

    Since a point has outer measure zero, it is a null set and is measurable.

    Example 1.2: Outer Measure of a Countable Set

    Problem: Show that any countable set E={a1,a2,}E = \{a_1, a_2, \ldots\} in Rn\mathbb{R}^n has outer measure zero.

    Solution:

    For each jj, cover the point aja_j by an open rectangle QjQ_j with volume Qj<ϵ2j|Q_j| < \frac{\epsilon}{2^j}.

    Then {Qj}j=1\{Q_j\}_{j=1}^{\infty} covers EE, and

    j=1Qj<j=1ϵ2j=ϵ\sum_{j=1}^{\infty} |Q_j| < \sum_{j=1}^{\infty} \frac{\epsilon}{2^j} = \epsilon

    Therefore, m(E)ϵm^*(E) \leq \epsilon for all ϵ>0\epsilon > 0, so m(E)=0m^*(E) = 0.

    This shows that countable sets are null sets and hence measurable.

    Example 1.3: Outer Measure of an Interval

    Problem: Show that for an interval I=[a,b]I = [a, b] in R\mathbb{R}, we have m(I)=bam^*(I) = b - a.

    Solution:

    First, note that II can be covered by itself (as an open rectangle by taking (aδ,b+δ)(a - \delta, b + \delta) for small δ>0\delta > 0), so m(I)ba+2δm^*(I) \leq b - a + 2\delta for any δ>0\delta > 0, hence m(I)bam^*(I) \leq b - a.

    For the reverse inequality, suppose {Qj}\{Q_j\} is a covering of II by open intervals. By compactness of II, finitely many Qj1,,QjkQ_{j_1}, \ldots, Q_{j_k} cover II. Then

    bai=1kQjij=1Qjb - a \leq \sum_{i=1}^k |Q_{j_i}| \leq \sum_{j=1}^{\infty} |Q_j|

    Taking the infimum over all coverings, we get m(I)bam^*(I) \geq b - a.

    Therefore, m(I)=bam^*(I) = b - a.

    Theorem 1.4: Regularity of Measurable Sets

    Let EE be a measurable set. Then:

    1. Outer Regularity: For any ϵ>0\epsilon > 0, there exists an open set GG such that EGE \subseteq G and m(GE)<ϵm(G \setminus E) < \epsilon.
    2. Inner Regularity: For any ϵ>0\epsilon > 0, there exists a closed set FF such that FEF \subseteq E and m(EF)<ϵm(E \setminus F) < \epsilon.
    Proof of Theorem 1.4:

    Proof:

    (1) By definition of outer measure, for any ϵ>0\epsilon > 0, there exists a covering {Qj}\{Q_j\} of EE by open rectangles such that Qj<m(E)+ϵ\sum |Q_j| < m(E) + \epsilon. Let G=QjG = \bigcup Q_j. Then GG is open, EGE \subseteq G, and

    m(GE)m(G)m(E)Qjm(E)<ϵm(G \setminus E) \leq m(G) - m(E) \leq \sum |Q_j| - m(E) < \epsilon

    (2) Apply (1) to EcE^c to get an open set GG containing EcE^c with m(GEc)<ϵm(G \setminus E^c) < \epsilon. Then F=GcF = G^c is closed, FEF \subseteq E, and m(EF)=m(GEc)<ϵm(E \setminus F) = m(G \setminus E^c) < \epsilon. ∎

    Remark 1.1: Borel Sets are Measurable

    All Borel sets (sets in the σ-algebra generated by open sets) are measurable. This follows from the fact that open sets are measurable and measurable sets form a σ-algebra.

    However, there exist non-Borel measurable sets. The construction of such sets typically requires the Axiom of Choice.

    Example 1.4: Cantor Set has Measure Zero

    Problem: Show that the Cantor set CC has outer measure zero.

    Solution:

    The Cantor set is constructed by removing the middle third of [0,1][0,1], then removing the middle third of each remaining interval, and so on.

    At the nn-th step, we remove 2n12^{n-1} intervals, each of length 13n\frac{1}{3^n}. The total length removed is:

    n=12n113n=13n=0(23)n=13112/3=1\sum_{n=1}^{\infty} 2^{n-1} \cdot \frac{1}{3^n} = \frac{1}{3} \sum_{n=0}^{\infty} \left(\frac{2}{3}\right)^n = \frac{1}{3} \cdot \frac{1}{1 - 2/3} = 1

    Since the total length removed is 1, and the Cantor set is what remains, we have m(C)=0m^*(C) = 0.

    Note: The Cantor set is uncountable but has measure zero, demonstrating that measure and cardinality are independent concepts.

    Lemma 1.1: Approximation by Rectangles

    For any set ERnE \subseteq \mathbb{R}^n and any ϵ>0\epsilon > 0, there exists a countable collection of open rectangles {Qj}\{Q_j\} such that:

    1. Ej=1QjE \subseteq \bigcup_{j=1}^{\infty} Q_j
    2. j=1Qjm(E)+ϵ\sum_{j=1}^{\infty} |Q_j| \leq m^*(E) + \epsilon
    Proof of Lemma 1.1:

    Proof:

    This follows directly from the definition of outer measure as an infimum. For any ϵ>0\epsilon > 0, there exists a covering {Qj}\{Q_j\} such that

    j=1Qj<m(E)+ϵ\sum_{j=1}^{\infty} |Q_j| < m^*(E) + \epsilon

    If m(E)=m^*(E) = \infty, the result is trivial. If m(E)<m^*(E) < \infty, the definition guarantees such a covering exists. ∎

    Theorem 1.5: Open Sets are Measurable

    Every open set in Rn\mathbb{R}^n is measurable.

    Proof of Theorem 1.5:

    Proof:

    Let GG be an open set. We need to show that for any ARnA \subseteq \mathbb{R}^n,

    m(A)m(AG)+m(AGc)m^*(A) \geq m^*(A \cap G) + m^*(A \cap G^c)

    The reverse inequality always holds by subadditivity, so we only need to prove this direction.

    Since GG is open, we can write G=j=1QjG = \bigcup_{j=1}^{\infty} Q_j as a countable union of disjoint open rectangles (this follows from the structure of open sets in Rn\mathbb{R}^n).

    For any ϵ>0\epsilon > 0, cover AA by rectangles {Rk}\{R_k\} with Rk<m(A)+ϵ\sum |R_k| < m^*(A) + \epsilon. Then

    m(AG)+m(AGc)j,kRkQj+kRkGckRk<m(A)+ϵm^*(A \cap G) + m^*(A \cap G^c) \leq \sum_{j,k} |R_k \cap Q_j| + \sum_k |R_k \cap G^c| \leq \sum_k |R_k| < m^*(A) + \epsilon

    Since ϵ>0\epsilon > 0 was arbitrary, the result follows. ∎

    Corollary 1.2: Closed Sets are Measurable

    Every closed set in Rn\mathbb{R}^n is measurable, as it is the complement of an open set.

    Example 1.5: Computing Outer Measure of a Union

    Problem: Let E1=[0,1]E_1 = [0, 1] and E2=[2,3]E_2 = [2, 3] in R\mathbb{R}. Compute m(E1E2)m^*(E_1 \cup E_2).

    Solution:

    Since E1E_1 and E2E_2 are disjoint intervals, they are measurable. Therefore,

    m(E1E2)=m(E1E2)=m(E1)+m(E2)=1+1=2m^*(E_1 \cup E_2) = m(E_1 \cup E_2) = m(E_1) + m(E_2) = 1 + 1 = 2

    This demonstrates countable additivity on disjoint measurable sets.

    Example 1.6: Outer Measure is Not Countably Additive

    Problem: Show that outer measure is not countably additive on all sets by constructing a counterexample.

    Solution:

    Using the Axiom of Choice, one can construct a non-measurable set VV (a Vitali set). For such a set, if we take countably many disjoint translates V1,V2,V_1, V_2, \ldots that cover an interval, we would have:

    m(Vj)<m(Vj)m^*\left(\bigcup V_j\right) < \sum m^*(V_j)

    This violates countable additivity. This is why we restrict to measurable sets to get a proper measure.

    Definition 1.4: Inner Measure

    For a bounded set ERnE \subseteq \mathbb{R}^n, the inner measure is defined as:

    m(E)=sup{m(K):KE,K compact}m_*(E) = \sup\{m(K) : K \subseteq E, K \text{ compact}\}

    For unbounded sets, we define m(E)=limnm(E[n,n]n)m_*(E) = \lim_{n \to \infty} m_*(E \cap [-n, n]^n).

    Theorem 1.6: Characterization of Measurability via Inner and Outer Measure

    A bounded set EE is measurable if and only if m(E)=m(E)m_*(E) = m^*(E).

    Proof of Theorem 1.6:

    Proof:

    If EE is measurable, regularity gives us compact sets KK and open sets GG such that KEGK \subseteq E \subseteq G and m(GK)<ϵm(G \setminus K) < \epsilon for any ϵ>0\epsilon > 0. This implies m(E)=m(E)m_*(E) = m^*(E).

    Conversely, if m(E)=m(E)m_*(E) = m^*(E), then for any ϵ>0\epsilon > 0, there exist compact KK and open GG with KEGK \subseteq E \subseteq G and m(GK)<ϵm(G \setminus K) < \epsilon. This approximation property implies measurability. ∎

    Theorem 1.7: Borel Sets are Measurable

    The σ-algebra B\mathcal{B} of Borel sets (generated by open sets) is contained in the σ-algebra M\mathcal{M} of measurable sets. That is, every Borel set is measurable.

    Proof of Theorem 1.7:

    Proof:

    Since open sets are measurable (Theorem 1.5), and measurable sets form a σ-algebra (Theorem 1.2), the σ-algebra generated by open sets (the Borel σ-algebra) must be contained in M\mathcal{M}.

    More formally, let B\mathcal{B} be the smallest σ-algebra containing all open sets. Since all open sets are in M\mathcal{M} and M\mathcal{M} is a σ-algebra, we have BM\mathcal{B} \subseteq \mathcal{M}. ∎

    Corollary 1.3: Gδ and Fσ Sets are Measurable

    Every GδG_\delta set (countable intersection of open sets) and every FσF_\sigma set (countable union of closed sets) is measurable.

    Example 1.7: Computing Outer Measure of a Rectangle

    Problem: Compute the outer measure of a closed rectangle R=[a1,b1]×[a2,b2]××[an,bn]R = [a_1, b_1] \times [a_2, b_2] \times \cdots \times [a_n, b_n] in Rn\mathbb{R}^n.

    Solution:

    The rectangle RR can be covered by itself (as an open rectangle by taking slightly larger intervals), so m(R)i=1n(biai)m^*(R) \leq \prod_{i=1}^n (b_i - a_i).

    For the reverse inequality, note that RR is compact. Any covering of RR by open rectangles has a finite subcover. Using the structure of rectangles, we can show that the sum of volumes in any covering is at least the volume of RR.

    Therefore, m(R)=i=1n(biai)m^*(R) = \prod_{i=1}^n (b_i - a_i), which is the expected volume of the rectangle.

    Since RR is a closed set, it is measurable, so m(R)=m(R)=i=1n(biai)m(R) = m^*(R) = \prod_{i=1}^n (b_i - a_i).

    Example 1.8: Outer Measure of a Dense Set

    Problem: Let E=Q[0,1]E = \mathbb{Q} \cap [0, 1] be the rationals in [0,1][0,1]. Compute m(E)m^*(E).

    Solution:

    Since EE is countable, by Example 1.2, we have m(E)=0m^*(E) = 0.

    This is interesting because EE is dense in [0,1][0,1] (every point in [0,1][0,1] is a limit point of EE), yet it has measure zero. This demonstrates that measure and topological density are independent concepts.

    The complement [0,1]E[0,1] \setminus E (the irrationals in [0,1][0,1]) has measure 1, even though it is nowhere dense in a certain sense.

    Theorem 1.8: Continuity from Above and Below

    Let {Ej}\{E_j\} be a sequence of measurable sets.

    1. If E1E2E_1 \supseteq E_2 \supseteq \cdots and m(E1)<m(E_1) < \infty, then
    m(j=1Ej)=limjm(Ej)m\left(\bigcap_{j=1}^{\infty} E_j\right) = \lim_{j \to \infty} m(E_j)
  • If E1E2E_1 \subseteq E_2 \subseteq \cdots, then
  • m(j=1Ej)=limjm(Ej)m\left(\bigcup_{j=1}^{\infty} E_j\right) = \lim_{j \to \infty} m(E_j)
    Proof of Theorem 1.8:

    Proof:

    (1) Write E1=j=1Ejj=1(EjEj+1)E_1 = \bigcap_{j=1}^{\infty} E_j \cup \bigcup_{j=1}^{\infty} (E_j \setminus E_{j+1}). Since the sets EjEj+1E_j \setminus E_{j+1} are disjoint,

    m(E1)=m(Ej)+j=1m(EjEj+1)m(E_1) = m\left(\bigcap E_j\right) + \sum_{j=1}^{\infty} m(E_j \setminus E_{j+1})

    Rearranging and using that m(Ej)=m(E1)k=1j1m(EkEk+1)m(E_j) = m(E_1) - \sum_{k=1}^{j-1} m(E_k \setminus E_{k+1}), we get the result.

    (2) Write j=1Ej=E1j=1(Ej+1Ej)\bigcup_{j=1}^{\infty} E_j = E_1 \cup \bigcup_{j=1}^{\infty} (E_{j+1} \setminus E_j). Since these sets are disjoint,

    m(Ej)=m(E1)+j=1m(Ej+1Ej)=limjm(Ej)m\left(\bigcup E_j\right) = m(E_1) + \sum_{j=1}^{\infty} m(E_{j+1} \setminus E_j) = \lim_{j \to \infty} m(E_j)

    Example 1.9: Application of Continuity from Below

    Problem: Let En=(0,11n)E_n = (0, 1 - \frac{1}{n}) for n2n \geq 2. Compute m(n=2En)m(\bigcup_{n=2}^{\infty} E_n).

    Solution:

    Note that E2E3E_2 \subseteq E_3 \subseteq \cdots and n=2En=(0,1)\bigcup_{n=2}^{\infty} E_n = (0, 1).

    We have m(En)=11nm(E_n) = 1 - \frac{1}{n}, so limnm(En)=1\lim_{n \to \infty} m(E_n) = 1.

    By continuity from below (Theorem 1.8),

    m((0,1))=m(n=2En)=limnm(En)=1m((0, 1)) = m\left(\bigcup_{n=2}^{\infty} E_n\right) = \lim_{n \to \infty} m(E_n) = 1

    This confirms that the measure of (0,1)(0,1) is 1, as expected.

    Lemma 1.2: Vitali Covering Lemma (Simplified)

    Let EE be a measurable set with m(E)<m(E) < \infty, and let V\mathcal{V} be a Vitali covering of EE (a collection of closed balls such that every point of EE is contained in balls of arbitrarily small radius).

    Then for any ϵ>0\epsilon > 0, there exists a finite disjoint subcollection {B1,,BN}\{B_1, \ldots, B_N\} such that

    m(Ei=1NBi)<ϵm\left(E \setminus \bigcup_{i=1}^N B_i\right) < \epsilon
    Proof of Lemma 1.2:

    Proof:

    This is a simplified version of the Vitali covering lemma. The key idea is to use a greedy algorithm: at each step, select the largest ball that is disjoint from previously selected balls.

    Since m(E)<m(E) < \infty, the process terminates after finitely many steps, and the remaining set has small measure. The full Vitali covering lemma (which we'll see in later courses) provides more precise control. ∎

    Theorem 1.9: Approximation by Simple Sets

    For any measurable set EE and any ϵ>0\epsilon > 0, there exist:

    1. A finite union of disjoint closed rectangles FF such that m(EF)<ϵm(E \triangle F) < \epsilon, where \triangle denotes symmetric difference.
    2. A finite union of open rectangles GG such that EGE \subseteq G and m(GE)<ϵm(G \setminus E) < \epsilon.
    Proof of Theorem 1.9:

    Proof:

    (1) By regularity, there exists a closed set F1F_1 with F1EF_1 \subseteq E and m(EF1)<ϵ/2m(E \setminus F_1) < \epsilon/2. Since F1F_1 is closed and bounded (we can restrict to a large ball), it is compact. Cover F1F_1 by finitely many open rectangles, then take the closure of their union and use the structure of rectangles to approximate.

    (2) By outer regularity, there exists an open set GG with EGE \subseteq G and m(GE)<ϵm(G \setminus E) < \epsilon. Since GG is open, it is a countable union of open rectangles. By continuity from below, we can approximate by a finite union. ∎

    Example 1.10: Constructing a Non-Measurable Set

    Problem: Describe the construction of a Vitali set (a non-measurable set).

    Solution:

    Define an equivalence relation on [0,1][0,1] by xyx \sim y if xyQx - y \in \mathbb{Q}. Using the Axiom of Choice, select one element from each equivalence class to form a set VV.

    For each rational q[1,1]q \in [-1, 1], consider the translate Vq=V+qV_q = V + q. These sets are disjoint and their union contains [0,1][0,1] and is contained in [1,2][-1, 2].

    If VV were measurable, then by translation invariance, all VqV_q would have the same measure. But then

    1m(qVq)31 \leq m\left(\bigcup_q V_q\right) \leq 3

    and by countable additivity (if measurable), this would give either 0 or \infty, a contradiction. Therefore, VV is not measurable.

    Note: This construction requires the Axiom of Choice. In models of set theory without AC, all sets might be measurable.

    Corollary 1.4: Measurable Sets are Dense

    For any set EE with m(E)<m^*(E) < \infty and any ϵ>0\epsilon > 0, there exists a measurable set FF such that EFE \subseteq F and m(FE)<ϵm^*(F \setminus E) < \epsilon.

    In other words, any set can be approximated from outside by a measurable set with arbitrarily small error.

    Remark 1.3: Historical Context

    The development of Lebesgue measure was a major breakthrough in analysis. Before Lebesgue, mathematicians struggled with the concept of "size" for irregular sets.

    Henri Lebesgue's 1902 thesis introduced the outer measure approach, which was later refined by Carathéodory's criterion. This construction provides a measure that:

    • Extends the intuitive notion of volume to a large class of sets
    • Satisfies countable additivity (crucial for integration theory)
    • Is translation invariant (natural for Euclidean space)
    • Allows approximation by simple sets (enabling computations)

    The discovery of non-measurable sets (requiring the Axiom of Choice) showed that we cannot measure all sets, but the class of measurable sets is large enough for all practical purposes in analysis.

    Remark 1.4: Connection to Integration

    The theory of measurable sets is the foundation for Lebesgue integration, which we will study in later courses. The key insight is that we can integrate functions over measurable sets, and the measure of a set tells us "how much" of the domain contributes to the integral.

    For example, if a function is zero almost everywhere (except on a null set), its integral will be zero, regardless of what the function does on that null set. This is why null sets and "almost everywhere" properties are so important.

    Example 1.11: Outer Measure of a Line Segment in ℝ²

    Problem: Compute the outer measure of the line segment L={(x,0):0x1}L = \{(x, 0) : 0 \leq x \leq 1\} in R2\mathbb{R}^2.

    Solution:

    For any ϵ>0\epsilon > 0, we can cover LL by rectangles of the form [0,1]×[δ,δ][0, 1] \times [-\delta, \delta] for small δ>0\delta > 0.

    The volume of such a rectangle is 2δ2\delta. By choosing δ<ϵ/2\delta < \epsilon/2, we get m(L)ϵm^*(L) \leq \epsilon.

    Since ϵ>0\epsilon > 0 was arbitrary, m(L)=0m^*(L) = 0.

    This shows that "lower-dimensional" sets (like curves in the plane) have measure zero in higher dimensions. This is a general phenomenon: a kk-dimensional manifold in Rn\mathbb{R}^n for k<nk < n has nn-dimensional Lebesgue measure zero.

    Example 1.12: Outer Measure of a Disk

    Problem: Show that for a disk D={(x,y):x2+y2r2}D = \{(x,y) : x^2 + y^2 \leq r^2\} in R2\mathbb{R}^2, we have m(D)=πr2m^*(D) = \pi r^2.

    Solution:

    The disk DD is contained in the square [r,r]×[r,r][-r, r] \times [-r, r], which has area 4r24r^2, so m(D)4r2m^*(D) \leq 4r^2.

    For a better estimate, we can cover DD by a grid of small squares. As the grid becomes finer, the sum of areas of squares covering DD approaches πr2\pi r^2.

    More rigorously, we can inscribe and circumscribe DD by polygons. The inscribed polygon has area approaching πr2\pi r^2 from below, and the circumscribed polygon from above.

    Since DD is closed (hence measurable), we have m(D)=m(D)=πr2m(D) = m^*(D) = \pi r^2, which matches our geometric intuition.

    Theorem 1.10: Product of Measurable Sets

    If E1Rn1E_1 \subseteq \mathbb{R}^{n_1} and E2Rn2E_2 \subseteq \mathbb{R}^{n_2} are measurable, then E1×E2Rn1+n2E_1 \times E_2 \subseteq \mathbb{R}^{n_1 + n_2} is measurable, and

    mn1+n2(E1×E2)=mn1(E1)mn2(E2)m_{n_1+n_2}(E_1 \times E_2) = m_{n_1}(E_1) \cdot m_{n_2}(E_2)

    where mkm_k denotes kk-dimensional Lebesgue measure.

    Proof of Theorem 1.10:

    Proof:

    First, if E1E_1 and E2E_2 are rectangles, the result is immediate from the definition of volume.

    For general measurable sets, use approximation by rectangles and the fact that the product of σ-algebras is contained in the product σ-algebra. The measure equality follows from Fubini's theorem (which we'll study later) or by direct approximation. ∎

    Example 1.13: Measure of a Cylinder

    Problem: Compute the 3-dimensional Lebesgue measure of the cylinder C={(x,y,z):x2+y2r2,0zh}C = \{(x,y,z) : x^2 + y^2 \leq r^2, 0 \leq z \leq h\}.

    Solution:

    The cylinder is the product of a disk D={(x,y):x2+y2r2}D = \{(x,y) : x^2 + y^2 \leq r^2\} in R2\mathbb{R}^2 and the interval [0,h][0, h] in R\mathbb{R}.

    By Theorem 1.10,

    m3(C)=m2(D)m1([0,h])=πr2h=πr2hm_3(C) = m_2(D) \cdot m_1([0,h]) = \pi r^2 \cdot h = \pi r^2 h

    This matches the familiar formula for the volume of a cylinder from elementary geometry.

    Corollary 1.5: Measure of Cartesian Products

    If EjRnjE_j \subseteq \mathbb{R}^{n_j} are measurable for j=1,,kj = 1, \ldots, k, then

    mn1++nk(E1××Ek)=j=1kmnj(Ej)m_{n_1+\cdots+n_k}(E_1 \times \cdots \times E_k) = \prod_{j=1}^k m_{n_j}(E_j)
    Remark 1.5: Computational Techniques

    When computing outer measures or Lebesgue measures, several techniques are useful:

    • Covering arguments: Find coverings that give upper bounds, then show these bounds are tight
    • Regularity: Approximate by open or closed sets, which are easier to work with
    • Product structure: Use Theorem 1.10 to compute measures of product sets
    • Symmetry: Use translation, rotation, or reflection invariance
    • Limiting arguments: Use continuity from above/below to compute limits of measures

    These techniques will be essential when we move to integration theory, where we need to compute measures of level sets and other geometric objects.

    Theorem 1.11: Operations on Measurable Sets

    Let EE and FF be measurable sets. Then:

    1. EFE \cup F is measurable
    2. EFE \cap F is measurable
    3. EFE \setminus F is measurable
    4. If E1,E2,E_1, E_2, \ldots are measurable, then j=1Ej\bigcap_{j=1}^{\infty} E_j is measurable

    Moreover, for disjoint measurable sets E1,E2,E_1, E_2, \ldots, we have:

    m(j=1Ej)=j=1m(Ej)m\left(\bigcup_{j=1}^{\infty} E_j\right) = \sum_{j=1}^{\infty} m(E_j)
    Proof of Theorem 1.11:

    Proof:

    (1) Since measurable sets form a σ-algebra (Theorem 1.2), finite unions are measurable.

    (2) By De Morgan's law, EF=(EcFc)cE \cap F = (E^c \cup F^c)^c. Since complements and unions of measurable sets are measurable, intersections are measurable.

    (3) Note that EF=EFcE \setminus F = E \cap F^c. Since FF is measurable, FcF^c is measurable, and by (2), the intersection is measurable.

    (4) By De Morgan's law, j=1Ej=(j=1Ejc)c\bigcap_{j=1}^{\infty} E_j = \left(\bigcup_{j=1}^{\infty} E_j^c\right)^c. Since countable unions and complements preserve measurability, countable intersections are measurable.

    The countable additivity property follows from Theorem 1.3. ∎

    Example 1.14: Computing Outer Measure of a Complex Set

    Problem: Let E=n=1[n,n+1n2]E = \bigcup_{n=1}^{\infty} [n, n + \frac{1}{n^2}] be a union of disjoint intervals. Compute m(E)m^*(E).

    Solution:

    Since the intervals [n,n+1n2][n, n + \frac{1}{n^2}] are disjoint and measurable, by countable additivity (Theorem 1.3),

    m(E)=m(E)=n=1m([n,n+1n2])=n=11n2m^*(E) = m(E) = \sum_{n=1}^{\infty} m\left([n, n + \frac{1}{n^2}]\right) = \sum_{n=1}^{\infty} \frac{1}{n^2}

    We recognize this as the sum of the reciprocals of squares, which converges to π26\frac{\pi^2}{6} (the Basel problem).

    Therefore, m(E)=π26m^*(E) = \frac{\pi^2}{6}.

    Verification: Each interval has measure 1n2\frac{1}{n^2}, and since they are disjoint, the total measure is the sum. The convergence of n=11n2\sum_{n=1}^{\infty} \frac{1}{n^2} ensures the result is finite.

    Example 1.15: Computing Outer Measure Using Symmetry and Decomposition

    Problem: Compute the outer measure of the set E={(x,y)R2:x+y1}E = \{(x, y) \in \mathbb{R}^2 : |x| + |y| \leq 1\} (a diamond shape).

    Solution:

    By symmetry, we can compute the measure of EE by considering the first quadrant and multiplying by 4.

    In the first quadrant, EE is the triangle with vertices (0,0)(0, 0), (1,0)(1, 0), and (0,1)(0, 1), bounded by x+y=1x + y = 1.

    The area of this triangle is 1211=12\frac{1}{2} \cdot 1 \cdot 1 = \frac{1}{2}.

    By symmetry, the total area is:

    m(E)=m(E)=412=2m^*(E) = m(E) = 4 \cdot \frac{1}{2} = 2

    Alternative approach: We can also compute this using Fubini's theorem (to be covered later) or by approximating with rectangles. Since EE is closed and bounded, it is compact and hence measurable.

    The boundary of EE is a null set (a curve in R2\mathbb{R}^2), so the measure equals the area computed geometrically.

    Remark 1.7: Computational Techniques and Common Pitfalls

    When computing outer measures, several techniques and common pitfalls should be noted:

    • Symmetry: Use geometric symmetry to reduce computations. For example, symmetric sets can often be computed by considering one symmetric component and multiplying.
    • Decomposition: Break complex sets into simpler measurable pieces. If the pieces are disjoint, use countable additivity.
    • Approximation: For sets that are not obviously measurable, use regularity to approximate by open or closed sets.
    • Null sets: Remember that boundaries of "nice" sets (like polygons, smooth curves) have measure zero, so they don't affect the measure.

    Common pitfalls:

    • Assuming countable additivity for all sets: Outer measure is only countably subadditive. Countable additivity holds only for measurable sets.
    • Ignoring overlap: When computing unions, be careful not to double-count overlapping regions. Use the inclusion-exclusion principle or decompose into disjoint sets.
    • Boundary issues: The measure of a set and its closure may differ if the boundary has positive measure (though this is rare for "nice" sets).
    • Infinite sums: When summing measures of disjoint sets, ensure the series converges. If it diverges, the measure is infinite.

    These techniques become especially important when computing integrals, where we need to measure level sets and other geometric objects.

    Remark 1.6: Comparison with Other Measures

    Lebesgue measure is not the only way to assign "size" to sets. Other important measures include:

    • Counting measure: For finite sets, the number of elements
    • Hausdorff measure: For fractals and lower-dimensional sets
    • Probability measures: In probability theory, measures with total mass 1
    • Dirac measures: Concentrated at a single point

    However, Lebesgue measure is the most natural choice for Rn\mathbb{R}^n because it:

    • Agrees with geometric volume for simple sets
    • Is translation invariant
    • Is countably additive
    • Has good approximation properties
    Example 1.16: Outer Measure of a Fractal-Like Set

    Problem: Consider the set E=n=1EnE = \bigcap_{n=1}^{\infty} E_n where E1=[0,1]E_1 = [0,1] and En+1E_{n+1} is obtained by removing the middle third of each interval in EnE_n (the Cantor set construction). Compute m(E)m^*(E).

    Solution:

    This is the Cantor set construction. At step nn, we have 2n2^n intervals, each of length 13n\frac{1}{3^n}, so m(En)=(23)nm(E_n) = \left(\frac{2}{3}\right)^n.

    Since E1E2E_1 \supseteq E_2 \supseteq \cdots and m(E1)=1<m(E_1) = 1 < \infty, by continuity from above (Theorem 1.8),

    m(E)=m(E)=limnm(En)=limn(23)n=0m^*(E) = m(E) = \lim_{n \to \infty} m(E_n) = \lim_{n \to \infty} \left(\frac{2}{3}\right)^n = 0

    Therefore, the Cantor set has measure zero, as shown in Example 1.4. This demonstrates that uncountable sets can have measure zero.

    Example 1.17: Measure of Intersection of Nested Sets

    Problem: Let En=(0,1/n)E_n = (0, 1/n) for nNn \in \mathbb{N}. Compute m(n=1En)m\left(\bigcap_{n=1}^{\infty} E_n\right).

    Solution:

    Note that E1E2E_1 \supseteq E_2 \supseteq \cdots and n=1En=n=1(0,1/n)=\bigcap_{n=1}^{\infty} E_n = \bigcap_{n=1}^{\infty} (0, 1/n) = \emptyset.

    However, we cannot directly apply continuity from above because m(E1)=1m(E_1) = 1 is finite, but the intersection is empty.

    Actually, we can apply Theorem 1.8: since m(E1)=1<m(E_1) = 1 < \infty and EnE_n \downarrow \emptyset, we have

    m(n=1En)=limnm(En)=limn1n=0m\left(\bigcap_{n=1}^{\infty} E_n\right) = \lim_{n \to \infty} m(E_n) = \lim_{n \to \infty} \frac{1}{n} = 0

    This confirms that the intersection has measure zero, which is consistent with it being empty.

    Corollary 1.6: Measure of Symmetric Differences

    For measurable sets EE and FF, the symmetric difference EF=(EF)(FE)E \triangle F = (E \setminus F) \cup (F \setminus E) is measurable, and

    m(EF)=m(E)+m(F)2m(EF)m(E \triangle F) = m(E) + m(F) - 2m(E \cap F)

    This follows from Theorem 1.11 and the fact that EFE \setminus F and FEF \setminus E are disjoint measurable sets.

    Remark 1.8: Advanced Topics and Extensions

    The theory of Lebesgue measure extends in several important directions:

    • Hausdorff measure: For sets of fractional dimension (fractals), Hausdorff measure provides a more appropriate notion of "size". The Hausdorff dimension of the Cantor set is log32\log_3 2.
    • Signed measures: Measures that can take negative values, important for the Radon-Nikodym theorem and Lebesgue decomposition.
    • Complex measures: Measures taking complex values, used in harmonic analysis and Fourier theory.
    • Abstract measure spaces: The theory extends to general measure spaces, not just Rn\mathbb{R}^n, enabling applications to probability, ergodic theory, and functional analysis.
    • Product measures: Measures on product spaces, leading to Fubini's theorem for multiple integrals.

    These extensions demonstrate the power and generality of measure theory as a foundation for modern analysis.

    Corollary 1.7: Approximation by Elementary Sets

    For any measurable set EE with m(E)<m(E) < \infty and any ϵ>0\epsilon > 0, there exists a finite union of disjoint closed rectangles RR such that m(ER)<ϵm(E \triangle R) < \epsilon.

    This follows from regularity (Theorem 1.4) and the fact that open sets can be approximated by finite unions of rectangles.

    Remark 1.2: Key Insights

    Key takeaways from this course:

    • Outer measure provides a systematic way to assign "size" to any subset of Rn\mathbb{R}^n
    • The Carathéodory criterion identifies exactly which sets should be considered measurable
    • Measurable sets form a σ-algebra, ensuring closure under countable operations
    • On measurable sets, outer measure becomes countably additive (Lebesgue measure)
    • Regularity allows approximation of measurable sets by simpler sets (open/closed)
    • Null sets play a crucial role in "almost everywhere" properties
    • Borel sets are measurable, but there exist non-Borel measurable sets
    • Continuity properties (from above/below) are essential for limit operations
    • Non-measurable sets exist (with AC), but are pathological and not encountered in practice

    Practice Quiz

    Measure Foundations & Outer Measure
    10
    Questions
    0
    Correct
    0%
    Accuracy
    1
    What is the definition of outer measure m(E)m^*(E) for a set ERnE \subseteq \mathbb{R}^n?
    Easy
    Not attempted
    2
    A set EE is measurable (in the sense of Carathéodory) if and only if:
    Medium
    Not attempted
    3
    Which of the following is NOT a property of outer measure?
    Medium
    Not attempted
    4
    A null set is defined as:
    Easy
    Not attempted
    5
    The regularity property states that for a measurable set EE:
    Hard
    Not attempted
    6
    If EE is measurable, then EcE^c is:
    Easy
    Not attempted
    7
    The outer measure of a countable set in R\mathbb{R} is:
    Medium
    Not attempted
    8
    Which property distinguishes Lebesgue measure from outer measure?
    Hard
    Not attempted
    9
    If E1,E2,E_1, E_2, \ldots are measurable sets, then i=1Ei\bigcup_{i=1}^{\infty} E_i is:
    Medium
    Not attempted
    10
    The Carathéodory extension theorem shows that:
    Hard
    Not attempted

    Frequently Asked Questions

    Why do we need outer measure instead of just using volume for all sets?

    Not all sets have a well-defined 'volume' in the geometric sense. Outer measure provides a systematic way to assign a 'size' to any subset of ℝⁿ, even pathological ones. It generalizes the intuitive notion of volume to sets that may be very irregular.

    What is the relationship between outer measure and Lebesgue measure?

    Lebesgue measure is the restriction of outer measure to the σ-algebra of measurable sets. On measurable sets, outer measure becomes countably additive and is called Lebesgue measure. Outer measure itself is only subadditive.

    Why is the Carathéodory criterion important?

    The Carathéodory criterion identifies exactly which sets should be considered 'measurable'. It ensures that on measurable sets, outer measure has the crucial property of countable additivity, which is essential for integration theory.

    Are all open sets measurable?

    Yes, all open sets in ℝⁿ are measurable. In fact, the Borel σ-algebra (generated by open sets) is contained in the σ-algebra of measurable sets. This means Borel sets are measurable, but there are also non-Borel measurable sets.

    What is a null set and why are they important?

    A null set is a set with outer measure zero. Null sets are important because properties that hold 'almost everywhere' (i.e., except on a null set) are sufficient for many purposes in analysis. For example, two functions that differ only on a null set have the same integral.

    How does regularity help in practice?

    Regularity allows us to approximate measurable sets by simpler sets (open or closed). This is crucial for proofs and computations, as we can often reduce problems about arbitrary measurable sets to problems about open or closed sets, which are easier to work with.

    Can a set have positive outer measure but zero Lebesgue measure?

    No. If a set has positive outer measure, it must be measurable (or we wouldn't be able to define its Lebesgue measure properly). If it's measurable and has positive outer measure, then its Lebesgue measure equals its outer measure and is positive.