0%

Lebesgue Integral

1. Lebesgue Integral of Non-negative Simple Function

Definition 1.1    Suppose \(E\) is measurable, \(E_i\) are disjoint measurable sets, and \[f(x)=\sum_{i=1}^s c_i\chi_{E_i}(x), c_i \geq 0.\] The Lebesgue integral of \(f\) on \(E\) is \[\int_E f(x)\text{d}x=\sum_{i=1}^s c_i \cdot m(E_i).\] Suppose \(A \subset E\), and \(f|_A(x)=c_i\), \(x \in A \cap E_i\), then \[\int_A f(x)\text{d}x=\sum_{i=1}^s c_i \cdot m(A \cap E_i).\]

Theorem 1.1    Suppose \(A\) and \(B\) are disjoint measurable subsets of \(E\), then \[\int_{A \cup B} f(x)\text{d}x=\int_A f(x)\text{d}x+\int_B f(x)\text{d}x.\]

Proof. Since \(A \cap B=\varnothing\), then \[\begin{aligned} \int_{A \cup B} f(x)\text{d}x&=\sum_{i=1}^s c_i \cdot m((A \cup B) \cap E_i) \\&=\sum_{i=1}^s c_i \cdot m((A \cap E_i) \cup (B \cap E_i)) \\&=\sum_{i=1}^s c_i \cdot (m(A \cap E_i)+m(B \cap E_i)) \\&=\sum_{i=1}^s c_i \cdot m(A \cap E_i)+\sum_{i=1}^s c_i \cdot m(B \cap E_i) \\&=\int_A f(x)\text{d}x+\int_B f(x)\text{d}x. \end{aligned}\]

\(\square\)

Theorem 1.2    Suppose \(\{A_n\}_{n=1}^\infty\) is an increasing collection of measurable subsets of \(E\), and \[E=\bigcup_{n=1}^\infty A_n=\lim_{n \to \infty} A_n.\] Then \[\int_E f(x)\text{d}x=\lim_{n \to \infty}\int_{A_n} f(x)\text{d}x.\]

Proof. We have \[\begin{aligned} \lim_{n \to \infty} \int_{A_n} f(x)\text{d}x&=\lim_{n \to \infty} \sum_{i=1}^s c_i \cdot m(A_n \cap E_i) \\&=\sum_{i=1}^s c_i \cdot \lim_{n \to \infty} m(A_n \cap E_i) \\&=\sum_{i=1}^s c_i \cdot m(E \cap E_i) \\&=\int_E f(x)\text{d}x. \end{aligned}\]

\(\square\)

Theorem 1.3    Suppose \(\alpha\) and \(\beta\) are non-negative real numbers. Then \[\int_E (\alpha f(x)+\beta g(x))\text{d}x=\alpha\int_E f(x)\text{d}x+\beta\int_E g(x)\text{d}x.\]

Proof. Suppose \(c \in \mathbb{R} \geq 0\), we have \(\displaystyle cf(x)=\sum_{i=1}^s cc_i \cdot \chi_{E_i}(x)\), where \(cc_i \geq 0\). Then \(cf\) is a non-negative simple function. Therefore, \[\begin{aligned} \int_E cf(x)\text{d}x&=\sum_{i=1}^s cc_i \cdot m(E_i) \\&=c\sum_{i=1}^s c_i \cdot m(E_i) \\&=c\int_E f(x)\text{d}x. \end{aligned}\]

Suppose \(\displaystyle f(x)=\sum_{i=1}^s c_i \cdot \chi_{E_i}(x)\) and \(\displaystyle g(x)=\sum_{i=1}^p d_i \cdot \chi_{\widetilde{E}_i}(x)\).

When \(x \in E_i \cap \widetilde{E}_j\), \(f(x)+g(x)=c_i+d_j \geq 0\). Hence, \[f(x)+g(x)=\sum_{\substack{1 \leq i \leq s \\ 1 \leq j \leq p}} (c_i+d_j) \cdot \chi_{E_i \cap \widetilde{E}_j}(x),\] and \(f+g\) is a non-negative simple function. Therefore, \[\begin{aligned} \int_E (f(x)+g(x))\text{d}x&=\sum_{\substack{1 \leq i \leq s \\ 1 \leq j \leq p}} (c_i+d_j) \cdot m(E_i \cap \widetilde{E}_j) \\&=\sum_{\substack{1 \leq i \leq s \\ 1 \leq j \leq p}} c_i \cdot m(E_i \cap \widetilde{E}_j)+\sum_{\substack{1 \leq i \leq s \\ 1 \leq j \leq p}} d_j \cdot m(E_i \cap \widetilde{E}_j) \\&=\sum_{1 \leq i \leq s} c_i\sum_{1 \leq j \leq p} m(E_i \cap \widetilde{E}_j)+\sum_{1 \leq j \leq p} d_j\sum_{1 \leq i \leq s} m(E_i \cap \widetilde{E}_j) \\&=\sum_{1 \leq i \leq s} c_i \cdot m(E_i)+\sum_{1 \leq j \leq p} d_j \cdot m(\widetilde{E}_j) \\&=\int_E f(x)\text{d}x+\int_E g(x)\text{d}x. \end{aligned}\]

Hence, it is easy to show that \[\int_E (\alpha f(x)+\beta g(x))\text{d}x=\alpha\int_E f(x)\text{d}x+\beta\int_E g(x)\text{d}x.\]

\(\square\)

2. Lebesgue Integral of Non-negative Measurable Function

2.1. Definition

Definition 2.1    Suppose \(E\) is measurable, and \(f\) is a non-negative measurable function on \(E\). The Lebesgue integral of \(f\) on \(E\) is \[\int_E f(x)\text{d}x=\sup\left\{\int_E \varphi(x)\text{d}x\right\},\] where \(\varphi\) is a non-negative simple function on \(E\), and \(\varphi(x) \leq f(x)\).

Theorem 2.1    Suppose \(\varphi\) and \(\psi\) are non-negative simple functions on \(E\). If \(\varphi(x) \leq \psi(x)\), then \[\int_E \varphi(x)\text{d}x \leq \int_E \psi(x)\text{d}x.\]

Corollary 2.2    Suppose \(f\) and \(g\) are non-negative measurable functions on \(E\). If \(f(x) \leq g(x)\), then \[\int_E f(x)\text{d}x \leq \int_E g(x)\text{d}x.\]

Proof. For any \(0 \leq \varphi(x) \leq f(x)\), where \(\varphi\) is a non-negative simple function, we have \(0 \leq \varphi(x) \leq g(x)\), \(x \in E\). Then \[\int_E g(x)\text{d}x \geq \int_E \varphi(x)\text{d}x.\] Therefore, \[\int_E g(x)\text{d}x \geq \sup_{0 \leq \varphi(x) \leq f(x)}\left\{\int_E \varphi(x)\text{d}x\right\}=\int_E f(x)\text{d}x.\]

\(\square\)

Definition 2.2    If \(\displaystyle \int_E f(x)\text{d}x<\infty\), then \(f\) is Lebesgue integrable on \(E\).

Theorem 2.3    If \(f\) is Lebesgue integrable on \(E\), then \(f\) is finite almost everywhere on \(E\), i.e., \[m(E[f=\infty])=0.\]

Proof. Since \(f(x) \geq f(x) \cdot \chi_{E[f \geq n]}(x)\), then \[\infty>\int_E f(x)\text{d}x \geq \int_E f(x) \cdot \chi_{E[f \geq n]}(x)\text{d}x=\int_{E[f \geq n]} f(x)\text{d}x \geq \int_{E[f \geq n]} n\text{d}x=n \cdot m(E[f \geq n]).\] Therefore, \[m(E[f \geq n]) \leq \frac{1}{n}\int_E f(x)\text{d}x.\] Since \(E[f \geq n] \supset E[f \geq n+1]\) and \(\displaystyle E[f=\infty]=\bigcap_{n=1}^\infty E[f \geq n]\), then \[m(E[f=\infty])=\lim_{n \to \infty} m(E[f \geq n])=0,\] i.e., \(f\) is finite almost everywhere on \(E\).

\(\square\)

Theorem 2.4    Suppose \(A\) is a measurable subset of \(E\). Then \[\int_A f(x)\text{d}x=\int_E f(x) \cdot \chi_A(x)\text{d}x.\]

Proof. We have \[\begin{aligned} \int_A f(x)\text{d}x&=\sup_{\substack{\varphi(x) \leq f(x) \\ x \in A}}\left\{\int_A \varphi(x)\text{d}x\right\} \\&=\sup_{\substack{\varphi(x)\chi_A(x) \leq f(x)\chi_A(x) \\ x \in E}}\left\{\int_A \varphi(x)\text{D}x\right\} \\&=\int_E f(x)\varphi_A(x)\text{d}x. \end{aligned}\]

\(\square\)

Theorem 2.5    Suppose \(A\) and \(B\) are disjoint measurable subsets of \(E\), then \[\int_{A \cup B} f(x)\text{d}x=\int_A f(x)\text{d}x+\int_B f(x)\text{d}x.\]

Proof. Take an arbitrary simple function \(\varphi\) on \(A \cup B\), where \(0 \leq \varphi(x) \leq f(x)\) for all \(x \in E\). We have \[\int_{A \cup B} \varphi(x)\text{d}x=\int_A \varphi(x)\text{d}x+\int_B \varphi(x)\text{d}x \leq \int_A f(x)\text{d}x+\int_B f(x)\text{d}x.\] Therefore, \[\int_{A \cup B} f(x)\text{d}x \leq \int_{A \cup B} \varphi(x)\text{d}x \leq \int_A f(x)\text{d}x+\int_B f(x)\text{d}x.\]

Take an arbitrary simple function \(\varphi_1\) on \(A\), where \(0 \leq \varphi_1(x) \leq f(x)\) for all \(x \in A\), and an arbitrary simple function \(\varphi_2\) on \(B\), where \(0 \leq \varphi_2(x) \leq f(x)\) for all \(x \in B\). Then \[\varphi(x)=\begin{cases} \varphi_1(x), &x \in A \\ \varphi_2(x), &x \in B \end{cases}\] is a simple function on \(A \cup B\), where \(0 \leq \varphi(x) \leq f(x)\). Hence, \[\int_{A \cup B} f(x)\text{d}x \geq \int_{A \cup B} \varphi(x)\text{d}x=\int_A \varphi(x)\text{d}x+\int_B \varphi(x)\text{d}x=\int_A \varphi_1(x)\text{d}x+\int_B \varphi_2(x)\text{d}x.\] Therefore, \[\int_{A \cup B} f(x)\text{d}x \geq \int_A f(x)\text{d}x+\int_B f(x)\text{d}x.\] As a consequence, \[\int_{A \cup B} f(x)\text{d}x=\int_A f(x)\text{d}x+\int_B f(x)\text{d}x.\]

\(\square\)

2.2. Limit and Integral

Theorem 2.6    Suppose \(\alpha\) and \(\beta\) are non-negative real numbers, then \[\int_E (\alpha f(x)+\beta g(x))\text{d}x=\alpha\int_E f(x)\text{d}x+\beta\int_E g(x)\text{d}x.\]

Lemma 2.6.1 (Levi's Monotonicity Theorem)    Suppose \(\{f_n\}_{n=1}^\infty\) is a sequence of non-negative measurable functions on \(E\). For \(x \in E\), and for all \(n\), we have \(f_n(x) \leq f_{n+1}(x)\). Then \[\lim_{n \to \infty} \int_E f_n(x)\text{d}x=\int_E \lim_{n \to \infty} f_n(x)\text{d}x.\]

Proof. Let \(\displaystyle f(x)=\lim_{n \to \infty} f_n(x)\). We know \(f\) is a non-negative measurable function on \(E\), and \(f_n(x) \leq f(x)\) for all \(n \in \mathbb{N}\). Hence, \[\int_E f_n(x)\text{d}x \leq \int_E f(x)\text{d}x.\] Since \[\int_E f_n(x)\text{d}x \leq \int_E f_{n+1}(x)\text{d}x,\] we have \[\lim_{n \to \infty} \int_E f_n(x)\text{d}x \leq \int_E f(x)\text{d}x.\]

Take an arbitrary non-negative simple function \(\varphi\) on \(E\), where \(0 \leq \varphi(x) \leq f(x)\). Take an arbitrary \(0<c<1\). Let \(E_n=E[f_n \geq c\varphi]\), which is a measurable subset of \(E\).

Since \(f_n(x) \leq f_{n+1}(x)\), then \[f_n(x) \geq c\varphi(x) \Rightarrow f_{n+1}(x) \geq c\varphi(x),\] i.e., \(E_n \subset E_{n+1}\).

Since \(\displaystyle \lim_{n \to \infty} f_n(x)=f(x)\), and \(c\varphi(x)<f(x)\), then when \(n \gg 0\), \(c\varphi(x) \leq f_n(x)\). Hence, \[\bigcup_{n=1}^\infty E_n=E.\]

Therefore, \[\int_E \varphi(x)\text{d}x=\lim_{n \to \infty} \int_{E_n} \varphi(x)\text{d}x.\] Since \[\lim_{n \to \infty} \int_E f_n(x)\text{d}x \geq \lim_{n \to \infty} \int_{E_n} f_n(x)\text{d}x \geq \lim_{n \to \infty} \int_{E_n} c\varphi(x)\text{d}x=c\lim_{n \to \infty} \int_{E_n} \varphi(x)\text{d}x,\] then \[\lim_{n \to \infty} \int_E f_n(x)\text{d}x \geq c\int_E \varphi(x)\text{d}x.\] Take \(c \to 1\), we have \[\lim_{n \to \infty} \int_E f_n(x)\text{d}x \geq \int_E \varphi(x)\text{d}x.\] Hence, \[\lim_{n \to \infty} \int_E f_n(x)\text{d}x \geq \int_E f(x)\text{d}x.\]

Therefore, \[\lim_{n \to \infty} \int_E f_n(x)\text{d}x=\int_E \lim_{n \to \infty} f_n(x)\text{d}x.\]

\(\square\)

Proof. Suppose \(c \in \mathbb{R} \geq 0\). When \(c>0\), we have \[\int_E cf(x)\text{d}x=\sup\left\{\int_E \varphi(x)\text{d}x\right\},\] where \(0 \leq \varphi(x) \leq cf(x)\), and \(\varphi\) is a simple function on \(E\). Let \(\displaystyle \psi(x)=\frac{1}{c}\varphi(x)\), which is a simple function on \(E\), then we have \[\int_E cf(x)\text{d}x=\sup\left\{\int_E c\psi(x)\text{d}x\right\}=\sup\left\{c\int_E \psi(x)\text{d}x\right\}=c\sup\left\{\int_E \psi(x)\text{d}x\right\},\] where \(0 \leq \psi(x) \leq f(x)\). Hence, \[\int_E cf(x)\text{d}x=c\int_E f(x)\text{d}x.\]

When \(c=0\), it is obvious that \[\int_E cf(x)\text{d}x=c\int_E f(x)\text{d}x.\]

There exist two sequences of non-negative simple function \(\{\varphi_n\}\) and \(\{\psi_n\}\) on \(E\) such that \[\lim_{n \to \infty} \varphi_n(x)=f(x), \lim_{n \to \infty} \psi_n(x)=g(x),\] where for all \(x \in E\), \(\varphi_n(x) \leq \varphi_{n+1}(x) \leq f(x)\), and \(\psi_n(x) \leq \psi_{n+1}(x) \leq g(x)\). We know \(\varphi_n+\psi_n\) is a non-negative simple function on \(E\), \(\varphi_n(x)+\psi_n(x) \leq \varphi_{n+1}(x)+\psi_{n+1}(x) \leq f(x)+g(x)\), and \[\lim_{n \to \infty} (\varphi_n(x)+\psi_n(x))=f(x)+g(x).\] By Levi's monotonicity theorem, we have \[\begin{aligned} \int_E \lim_{n \to \infty} (\varphi_n(x)+\psi_n(x))\text{d}x&=\int_E (f(x)+g(x))\text{d}x \\&=\lim_{n \to \infty} \int_E (\varphi_n(x)+\psi_n(x))\text{d}x \\&=\lim_{n \to \infty} \left(\int_E \varphi_n(x)\text{d}x+\int_E \psi_n(x)\text{d}x\right) \\&=\lim_{n \to \infty} \int_E \varphi_n(x)\text{d}x+\lim_{n \to \infty} \int_E \psi_n(x)\text{d}x \\&=\int_E \lim_{n \to \infty} \varphi_n(x)\text{d}x+\int_E \lim_{n \to \infty} \psi_n(x)\text{d}x \\&=\int_E f(x)\text{d}x+\int_E g(x)\text{d}x. \end{aligned}\]

Hence, it is easy to show that \[\int_E (\alpha f(x)+\beta g(x))\text{d}x=\alpha\int_E f(x)\text{d}x+\beta\int_E g(x)\text{d}x.\]

\(\square\)

Theorem 2.7    Suppose \(\{f_n\}_{n=1}^\infty\) is a sequence of non-negative measurable functions on \(E\). Then \[\int_E \left(\sum_{n=1}^\infty f_n(x)\right)\text{d}x=\sum_{n=1}^\infty \int_E f_n(x)\text{d}x.\]

Proof. Let \(\displaystyle g_n(x)=\sum_{k=1}^n f_k(x)\), which is a non-negative measurable function. Besides, \[g_{n+1}(x)=g_n(x)+f_{n+1}(x) \geq g_n(x).\] By Levi's monotonicity theorem, we have \[\lim_{n \to \infty} \int_E g_n(x)\text{d}x=\int_E \lim_{n \to \infty} g_n(x)\text{d}x=\int_E \left(\sum_{n=1}^\infty f_n(x)\right)\text{d}x.\] Since \[\lim_{n \to \infty} \int_E g_n(x)\text{d}x=\lim_{n \to \infty} \sum_{k=1}^n \int_E f_k(x)\text{d}x=\sum_{n=1}^\infty \int_E f_n(x)\text{d}x,\] then \[\int_E \left(\sum_{n=1}^\infty f_n(x)\right)\text{d}x=\sum_{n=1}^\infty \int_E f_n(x)\text{d}x.\]

\(\square\)

Theorem 2.8 (Fatou's Lemma)    Suppose \(\{f_n\}_{n=1}^\infty\) is a sequence of non-negative measurable functions on \(E\). Then \[\int_E \varliminf_{n \to \infty} f_n(x)\text{d}x \leq \varliminf_{n \to \infty} \int_E f_n(x)\text{d}x.\]

Proof. Since \(\displaystyle \left\{\inf_{k \geq n} f_k\right\}\) is an increasing sequence of non-negative measurable functions on \(E\). By Levi's monotonicity theorem, we have \[\int_E \varliminf_{n \to \infty} f_n(x)\text{d}x=\int_E \left(\lim_{n \to \infty} \inf_{k \geq n} f_k(x)\right)\text{d}x=\lim_{n \to \infty} \int_E \inf_{k \geq n} f_k(x)\text{d}x.\] Since \[\int_E \inf_{k \geq n} f_k(x)\text{d}x \leq \inf_{k \geq n} \int_E f_k(x)\text{d}x,\] then \[\lim_{n \to \infty} \int_E \inf_{k \geq n} f_k(x)\text{d}x \leq \lim_{n \to \infty} \inf_{k \geq n} \int_E f_k(x)\text{d}x=\varliminf_{n \to \infty} \int_E f_n(x)\text{d}x.\] Hence, \[\int_E \varliminf_{n \to \infty} f_n(x)\text{d}x \leq \varliminf_{n \to \infty} \int_E f_n(x)\text{d}x.\]

\(\square\)

The equal sign in the Fatou's lemma does not always hold. For example, let \[f_n(x)=\begin{cases}n, &\displaystyle 0<x<\frac{1}{n} \\0, &\displaystyle x \geq \frac{1}{n}\end{cases}.\] For any \(x \in (0, \infty)\), \(\displaystyle \lim_{n \to \infty} f_n(x)=0\), and thus \[\int_{(0, \infty)} \lim_{n \to \infty} f_n(x)\text{d}x=0.\] However, \[\int_{(0, \infty)} f_n(x)\text{d}x=\int_{\left(0, \frac{1}{n}\right)} f_n(x)\text{d}x+\int_{\left[\frac{1}{n}, \infty\right)} f_n(x)\text{d}x=1,\] and thus \[\lim_{n \to \infty} \int_{(0, \infty)} f_n(x)\text{d}x=1.\]

2.3. Zero Measure Set and Integral

Theorem 2.9    Suppose \(E\) is a zero measure set. Then \[\int_E f(x)\text{d}x=0.\]

Proof. By definition, \[\int_E f(x)\text{d}x=\sup\left\{\int_E \varphi(x)\text{d}x\right\},\] where \(\varphi\) is a non-negative simple function on \(E\), and \(\varphi(x) \leq f(x)\). Since \[\int_E \varphi(x)\text{d}x=0,\] then \[\int_E f(x)\text{d}x=0.\]

\(\square\)

Theorem 2.10    If \(f(x) \leq g(x)\ \text{a.e.}\ x \in E\), then \[\int_E f(x)\text{d}x \leq \int_E g(x)\text{d}x.\]

Proof. Let \(E_1=E[f \leq g]\) and \(E_2=E[f>g]\), where \(m(E_2)=0\), then \(E=E_1 \cup E_2\), and \(E_1 \cap E_2=\varnothing\). Hence, \[\begin{aligned} \int_E f(x)\text{d}x&=\int_{E_1} f(x)\text{d}x+\int_{E_2} f(x)\text{d}x \\& \leq \int_{E_1} g(x)\text{d}x+0 \\&=\int_{E_1} g(x)\text{d}x+\int_{E_2} g(x)\text{d}x \\&=\int_E g(x)\text{d}x. \end{aligned}\]

\(\square\)

Corollary 2.11    If \(f(x)=g(x)\ \text{a.e.}\ x \in E\), then \[\int_E f(x)\text{d}x=\int_E g(x)\text{d}x.\] Moreover, if \(f(x)=0\ \text{a.e.}\ x \in E\), then \[\int_E f(x)\text{d}x=0.\]

Theorem 2.12    If \[\int_E f(x)\text{d}x=0,\] then \(f(x)=0\ \text{a.e.}\ x \in E\).

Proof. Since \[0=\int_E f(x)\text{d}x \geq \int_{E\left[f \geq \frac{1}{n}\right]} f(x)\text{d}x \geq \int_{E\left[f \geq \frac{1}{n}\right]} \frac{1}{n}\text{d}x=\frac{1}{n} \cdot m\left(E\left[f \geq \frac{1}{n}\right]\right),\] then \(\displaystyle m\left(E\left[f \geq \frac{1}{n}\right]\right)=0\). Hence, \[m(E[f \neq 0])=m\left(\bigcup_{n=1}^\infty E\left[f \geq \frac{1}{n}\right]\right) \leq \sum_{n=1}^\infty m\left(E\left[f \geq \frac{1}{n}\right]\right)=0.\] Therefore, \(m(E[f \geq 0])=0\), i.e., \(f(x)=0\ \text{a.e.}\ x \in E\).

\(\square\)

3. Lebesgue Integral of Measurable Function

Definition 3.1    Suppose \(E\) is measurable, and \(f\) is a measurable function on \(E\). When at least one of \(\displaystyle \int_E f^+(x)\text{d}x\) and \(\displaystyle \int_E f^-(x)\text{d}x\) is finite, the Lebesgue integral of \(f\) on \(E\) is \[\int_E f(x)\text{d}x=\int_E f^+(x)\text{d}x-\int_E f^-(x)\text{d}x.\] If \(\displaystyle \int_E f^+(x)\text{d}x\) and \(\displaystyle \int_E f^-(x)\text{d}x\) are finite, then \(f\) is Lebesgue integrable on \(E\). The collection of all Lebesgue integrable functions on \(E\) is denoted \(L(E)\).

Theorem 3.1    Suppose \(E \neq \varnothing\) and \(m(E)=0\). Then any real function \(f\) on \(E\) is Lebesgue integrable, and \[\int_E f(x)\text{d}x=0.\]

Proof. Since \(m(E)=0\), and \(f^+\) and \(f^-\) are non-negative measurable on \(E\), then \[\int_E f^+(x)\text{d}x=\int_E f^-(x)\text{d}x=0.\] Hence, \(f \in L(E)\), and \[\int_E f(x)\text{d}x=\int_E f^+(x)\text{d}x-\int_E f^-(x)\text{d}x=0.\]

\(\square\)

Theorem 3.2    Suppose \(\displaystyle E=\bigcup_{n=1}^\infty E_n\), and \(\{E_n\}\) is a collection of disjoint measurable subsets. If the Lebesgue integral of \(f\) on \(E\) exists, then \[\int_E f(x)\text{d}x=\sum_{n=1}^\infty \int_{E_n} f(x)\text{d}x.\]

Proof. Suppose \(f\) is a non-negative measurable function. We have \[\int_{E_n} f(x)\text{d}x=\int_E f(x) \cdot \chi_{E_n}(x)\text{d}x\] and thus \[\sum_{n=1}^\infty \int_{E_n} f(x)\text{d}x=\sum_{n=1}^\infty \int_E f(x) \cdot \chi_{E_n}(x)\text{d}x=\int_E \sum_{n=1}^\infty f(x) \cdot \chi_{E_n}(x)\text{d}x=\int_E f(x)\text{d}x.\]

Suppose \(f\) is a measurable function. We know at least one of \(\displaystyle \sum_{n=1}^\infty \int_{E_n} f^+(x)\text{d}x\) and \(\displaystyle \sum_{n=1}^\infty \int_{E_n} f^-(x)\text{d}x\) is convergent. Hence, \[\begin{aligned} \int_E f(x)\text{d}x&=\int_E f^+(x)\text{d}x-\int_E f^-(x)\text{d}x \\&=\left(\sum_{n=1}^\infty \int_{E_n} f^+(x)\text{d}x\right)-\left(\sum_{n=1}^\infty \int_{E_n} f^-(x)\text{d}x\right) \\&=\sum_{n=1}^\infty \left(\int_{E_n} f^+(x)\text{d}x-\int_{E_n} f^-(x)\text{d}x\right) \\&=\sum_{n=1}^\infty \int_{E_n} f(x)\text{d}x. \end{aligned}\]

\(\square\)

Theorem 3.3    \(f \in L(E) \Leftrightarrow |f| \in L(E)\).

Proof. (\(\Leftarrow\)) Suppose \(f \in L(E)\). Then \(\displaystyle \int_E f^+(x)\text{d}x<\infty\) and \(\displaystyle \int_E f^-(x)\text{d}x\), and thus \[\int_E |f(x)|\text{d}x=\int_E (f^+(x)+f^-(x))\text{d}x=\int_E f^+(x)\text{d}x+\int_E f^-(x)\text{d}x<\infty,\] i.e., \(|f| \in L(E)\).

(\(\Rightarrow\)) Suppose \(|f| \in L(E)\). Then \(f^+(x) \leq |f(x)|\) and \(f^-(x) \leq |f(x)|\). Hence, \[\int_E f^+(x)\text{d}x \leq \int_E |f(x)|\text{d}x<\infty\] and \[\int_E f^-(x)\text{d}x \leq \int_E |f(x)|\text{d}x<\infty.\] Therefore, \(f \in L(E)\).

\(\square\)

Theorem 3.4    Suppose \(|f(x)| \leq g(x)\ \text{a.e.}\ x \in E\), and \(g\) is a non-negative Lebesgue integrable function on \(E\). Then \(f \in L(E)\), and \[\left|\int_E f(x)\text{d}x\right| \leq \int_E |f(x)|\text{d}x \leq \int_E g(x)\text{d}x.\]

Proof. Since \(|f(x)| \leq g(x)\ \text{a.e.}\ x \in E\), \(|f|\) and \(g\) are non-negative measurable functions on \(E\), and \(g\) is Lebesgue integrable, then \[\int_E |f(x)|\text{d}x \leq \int_E g(x)\text{d}x<\infty,\] i.e., \(|f| \in L(E)\). Hence, \(f \in L(E)\).

In addition, \[\begin{aligned} \left|\int_E f(x)\text{d}x\right|&=\left|\int_E f^+(x)\text{d}x-\int_E f^-(x)\text{d}x\right| \\&\leq \left|\int_E f^+(x)\text{d}x\right|+\left|\int_E f^-(x)\text{d}x\right| \\&=\int_E f^+(x)\text{d}x+\int_E f^-(x)\text{d}x \\&=\int_E (f^+(x)+f^-(x))\text{d}x \\&=\int_E |f(x)|\text{d}x. \end{aligned}\]

\(\square\)

Theorem 3.5    Suppose \(f \in L(E)\), and \(A\) is a measurable subset of \(E\), then \[\forall \varepsilon>0, \exists \delta>0\ \text{s.t.}\ m(A)<\delta \Rightarrow \left|\int_A f(x)\text{d}x\right| \leq \int_A |f(x)|\text{d}x<\varepsilon.\]

Proof. Take an arbitrary \(\varepsilon>0\). Suppose \(f\) is a non-negative simple function. We know \(f(x) \leq M\), for all \(x \in A\). Take \(\displaystyle \delta=\frac{\varepsilon}{M+1}\), when \(m(A)<\delta\), we have \[\int_A f(x)\text{d}x \leq M \cdot m(A)<\varepsilon.\]

Suppose \(f\) is a measurable function, then \(|f|\) is a non-negative measurable function. There exists a non-negative simple function such that \(\varphi(x) \leq |f(x)|\) and \[\int_E \varphi(x)\text{d}x \leq \int_E |f(x)|\text{d}x \leq \int_E \varphi(x)+\frac{\varepsilon}{2}.\]

We take a \(\delta\) such that when \(m(A)<\delta\), \(\displaystyle \int_A \varphi(x)\text{d}x<\frac{\varepsilon}{2}\). Hence, \[\begin{aligned} \int_A |f(x)|\text{d}x&=\int_A \varphi(x)\text{d}x+\int_A (|f(x)|-\varphi(x))\text{d}x \\&\leq \int_A \varphi(x)\text{d}x+\int_E(|f(x)|-\varphi(x))\text{d}x \\&=\int_A \varphi(x)\text{d}x+\int_E |f(x)|\text{d}x-\int_E \varphi(x)\text{d}x \\&<\frac{\varepsilon}{2}+\frac{\varepsilon}{2}=\varepsilon. \end{aligned}\]

\(\square\)

Theorem 3.6    Suppose \(f\) and \(g\) are Lebesgue integrable functions, and \(\alpha\) and \(\beta\) are real numbers. Then \(\alpha f+\beta g \in L(E)\), and \[\int_E (\alpha f(x)+\beta g(x))\text{d}x=\alpha\int_E f(x)\text{d}x+\beta\int_E g(x)\text{d}x.\]

Proof. Suppose \(c \in \mathbb{R}\). When \(c=0\), it is obvious that \(cf \in L(E)\), and \[\int_E cf(x)\text{d}x=c\int_E f(x)\text{d}x.\]

When \(c>0\), \((cf)^+=cf^+\) and \((cf)^-=cf^-\).

Since \(f \in L(E)\), then \(\displaystyle \int_E f^+(x)\text{d}x<\infty\) and \(\displaystyle \int_E f^-(x)\text{d}x<\infty\), and thus \[\int_E (cf)^+(x)\text{d}x=\int_E cf^+(x)\text{d}x=c\int_E f^+(x)\text{d}x<\infty\] and \[\int_E (cf)^-(x)\text{d}x=\int_E cf^-(x)\text{d}x=c\int_E f^-(x)\text{d}x<\infty.\] Therefore, \(cf(x) \in L(E)\), and \[\begin{aligned} \int_E cf(x)\text{d}x&=\int_E (cf)^+(x)\text{d}x-\int_E (cf)^-(x)\text{d}x \\&=c\int_E f^+(x)\text{d}x-c\int_E f^-(x)\text{d}x \\&=c\left(\int_E f^+(x)\text{d}x-\int_E f^-(x)\text{d}x\right) \\&=c\int_E f(x)\text{d}x. \end{aligned}\]

When \(c<0\), \((cf)^+=|c|f^-\) and \((cf)^-=|c|f^+\). Similarly, we can show that \(cf \in L(E)\), and \[\int_E cf(x)\text{d}x=c\int_E f(x)\text{d}x.\]

In addition, we know \(|f|\) and \(|g|\) are Lebesgue integrable, and \[\int_E(|f|+|g|)(x)\text{d}x=\int_E |f|(x)\text{d}x+\int_E |g|(x)\text{d}x<\infty,\] i.e., \(|f|+|g| \in L(E)\). Since \(|f+g| \leq |f|+|g|\), then \(|f+g| \in L(E)\).

Since \(f+g=(f+g)^+-(f+g)^-=f^+-f^-+g^+-g^-\), then \[(f+g)^++f^-+g^-=(f+g)^-+f^++g^+.\] Hence, \[\int_E ((f+g)^+(x)+f^-(x)+g^-(x))\text{d}x=\int_E ((f+g)^-(x)+f^+(x)+g^+(x))\text{d}x,\] and thus \[\int_E (f+g)^+(x)\text{d}x+\int_E f^-(x)\text{d}x+\int_E g^-(x)\text{d}x=\int_E (f+g)^-(x)\text{d}x+\int_E f^+(x)\text{d}x+\int_E g^+(x)\text{d}x.\] Therefore, \[\int_E (f+g)^+(x)\text{d}x-\int_E (f+g)^-(x)\text{d}x=\int_E f^+(x)\text{d}x-\int_E f^-(x)\text{d}x+\int_E g^+(x)\text{d}x-\int_E g^-(x)\text{d}x,\] i.e., \[\int_E (f(x)+g(x))\text{d}x=\int_E f(x)\text{d}x+\int_E g(x)\text{d}x.\]

Hence, it is easy to show that \(\alpha f+\beta g \in L(E)\), and \[\int_E (\alpha f(x)+\beta g(x))\text{d}x=\alpha\int_E f(x)\text{d}x+\beta\int_E g(x)\text{d}x.\]

\(\square\)

3.1. Lebesgue's Dominated Convergence Theorem

Theorem 3.7 (Lebesgue's Dominated Convergence Theorem, DCT)    Suppose \(\{f_n\}\) is a sequence of measurable functions on \(E\), and \(\displaystyle \lim_{n \to \infty} f_n(x)=f(x)\ \text{a.e.}\ x \in E\). If there exists a non-negative Lebesgue integrable function \(F\) on \(E\) such that for all \(n \in \mathbb{N}\), \(|f_n(x)| \leq F(x)\ \text{a.e.}\ x \in E\), then \[\lim_{n \to \infty} \int_E f_n(x)\text{d}x=\int_E f(x)\text{d}x.\]

Proof. Since \(|f_n(x)| \leq F(x)\ \text{a.e.}\ x \in E\), then \(|f(x)| \leq F(x)\ \text{a.e.}\ x \in E\). Since \(F\) is Lebesgue integrable, then \(f\) and \(f_n\) are Lebesgue integrable, and thus \(F\), \(f\) and \(f_n\) are finite almost everywhere.

Assume without loss of generality that \(F\), \(f\) and \(f_n\) are finite. Let \(g_n(x)=|f_n(x)-f(x)| \geq 0\). We know \(g_n(x) \leq |f_n(x)|+|f(x)| \leq 2F(x)\ \text{a.e.}\ x \in E\), then \(2F-g_n\) is a non-negative measurable function.

By Fatou's lemma, \[\int_E \varliminf_{n \to \infty} (2F(x)-g_n(x))\text{d}x \leq \varliminf_{n \to \infty} \int_E (2F(x)-g_n(x))\text{d}x.\] Hence, \[\int_E 2F(x)\text{d}x-\int_E \varlimsup_{n \to \infty} g_n(x)\text{d}x \leq \int_E 2F(x)\text{d}x-\varlimsup_{n \to \infty} \int_E g_n(x)\text{d}x,\] and thus \[\varlimsup_{n \to \infty} \int_E g_n(x)\text{d}x \leq \int_E \varlimsup_{n \to \infty} g_n(x)\text{d}x.\] Since \(\displaystyle \varlimsup_{n \to \infty} g_n(x)\text{d}x=0\ \text{a.e.}\ x \in E\), then \[\varlimsup_{n \to \infty} \int_E g_n(x)\text{d}x=0,\] and thus \[\lim_{n \to \infty} \int_E g_n(x)\text{d}x=\lim_{n \to \infty} \int_E |f_n(x)-f(x)|\text{d}x=0.\] Since \[\left|\int_E (f_n(x)-f(x))\text{d}x\right| \leq \int_E |f_n(x)-f(x)|\text{d}x,\] then \[\lim_{n \to \infty} \int_E (f_n(x)-f(x))\text{d}x=0,\] i.e., \[\lim_{n \to \infty} \int_E f_n(x)\text{d}x=\int_E f(x)\text{d}x.\]

\(\square\)

Theorem 3.8    Suppose \(\{f_n\}\) is a sequence of measurable functions on \(E\), and \(f_n \Rightarrow f\). If there exists a non-negative Lebesgue integrable function \(F\) on \(E\) such that for all \(n \in \mathbb{N}\), \(|f_n(x)| \leq F(x)\ \text{a.e.}\ x \in E\), then \[\lim_{n \to \infty} \int_E f_n(x)\text{d}x=\int_E f(x)\text{d}x.\]

Proof. Since \(|f_n(x)| \leq F(x)\ \text{a.e.}\ x \in E\), then \(|f(x)| \leq F(x)\ \text{a.e.}\ x \in E\). Since \(F\) is Lebesgue integrable, then \(f\) and \(f_n\) are Lebesgue integrable, and thus \(F\), \(f\) and \(f_n\) are finite almost everywhere.

Assume without loss of generality that \(F\), \(f\) and \(f_n\) are finite. Suppose \[\int_E |f_n(x)-f(x)|\text{d}x \not\to0,\] then we can find a subsequence \(\{f_{n_k}\}_{k=1}^\infty\) such that \(\displaystyle \lim_{k \to \infty} \int_E |f_{n_k}(x)-f(x)|\text{d}x=\alpha>0\).

By Riesz's theorem, there exists a subsequence \(\{f_{n_{k_i}}\}_{i=1}^\infty\) such that \(f_{n_{k_i}}(x) \to f(x)\ \text{a.e.}\ x \in E\), and thus \[\lim_{i \to \infty} \int_E |f_{n_{k_i}}(x)-f(x)|\text{d}x=0,\] which is a contradiction.

\(\square\)

Theorem 3.9    Suppose \(\{f_n\}\) is a sequence of Lebesgue integrable functions on \(E\). If \[\sum_{n=1}^\infty \int_E |f_n(x)|\text{d}x<\infty,\] then \(\displaystyle \sum_{n=1}^\infty f_n(x)<\infty\ \text{a.e.}\ x \in E\), \(\displaystyle \sum_{n=1}^\infty f_n\) is Lebesgue integrable on \(E\), and \[\int_E \left(\sum_{n=1}^\infty f_n(x)\right)\text{d}x=\sum_{n=1}^\infty \int_E f_n(x)\text{d}x.\]

Proof. Let \[F(x)=\sum_{n=1}^\infty |f_n(x)|.\] We have \[\int_E F(x)\text{d}x=\sum_{n=1}^\infty \int_E |f_n(x)|\text{d}x<\infty.\] Hence, \(F\) is a non-negative Lebesgue integrable function on \(E\), and thus \(F\) is finite almost everywhere on \(E\), i.e., \(\displaystyle \sum_{n=1}^\infty |f_n(x)|<\infty\ \text{a.e.}\ x \in E\). Therefore, \(\displaystyle \sum_{n=1}^\infty f_n(x)<\infty\ \text{a.e.}\ x \in E\).

Let \(\displaystyle g_n(x)=\sum_{k=1}^n f_k(x)\) and \(\displaystyle g(x)=\sum_{n=1}^\infty f_n(x)=\lim_{n \to \infty} g_n(x)\). We have \[|g_n(x)|=\left|\sum_{k=1}^n f_k(x)\right| \leq \sum_{k=1}^n |f_k(x)| \leq F(x).\] By Lebesgue's dominated convergence theorem, \(g \in L(E)\), and \[\lim_{n \to \infty} \int_E g_n(x)\text{d}x=\int_E g(x)\text{d}x,\] i.e., \(\displaystyle \sum_{n=1}^\infty f_n\) is Lebesgue integrable on \(E\), and \[\sum_{n=1}^\infty \int_E f_n(x)\text{d}x=\lim_{n \to \infty} \sum_{k=1}^n \int_E f_k(x)\text{d}x=\lim_{n \to \infty} \int_E \sum_{k=1}^n f_k(x)\text{d}x=\int_E \left(\sum_{n=1}^\infty f_n(x)\right)\text{d}x.\]

\(\square\)

Theorem 3.10    Suppose \(f\) is a real function on \(E \times (a, b)\). If for all \(t \in (a, b)\), \(f(x, t)\) (as a function of \(x\)) is Lebesgue integrable on \(E\), for \(\text{a.e.}\ x \in E\), \(f(x, t)\) (as a function of \(t\)) is differentiable on \((a, b)\), and there exists a non-negative Lebesgue integrable function \(F\) on \(E\) such that \[\left|\frac{\partial}{\partial t}f(x, t)\right| \leq F(x),\] then \(\displaystyle \frac{\partial}{\partial t}f(x, t)\) is Lebesgue integrable on \(E\), \(\displaystyle \int_E f(x, t)\text{d}x\) is differentiable on \((a, b)\), and \[\frac{\text{d}}{\text{d}t}\int_E f(x, t)\text{d}x=\int_E \frac{\partial}{\partial t}f(x, t)\text{d}x.\]

Proof. Fix a \(t \in (a, b)\) and take an arbitrary sequence \(\{h_n\}\), where \(h_n \to 0\) and \(h_n \neq 0\). Let \[g_n(x)=\frac{f(x, t+h_n)-f(x, t)}{h_n},\] then \[\lim_{n \to \infty} g_n(x)=\frac{\partial}{\partial t}f(x, t),\] and \[|g_n(x)|=\left|\frac{\partial}{\partial t}f(x, t+\theta_nh_n)\right| \leq F(x),\] where \(0 \leq \theta_n \leq 1\).

Since \(\displaystyle \frac{\partial}{\partial t}f(x, t)\) is Lebesgue integrable on \(E\), and by Lebesgue's dominated convergence theorem, \[\lim_{n \to \infty} \frac{\displaystyle \int_E f(x, t+h_n)\text{d}x-\int_E f(x, t)\text{d}x}{h_n}=\lim_{n \to \infty} \int_E g_n(x)\text{d}x=\int_E \lim_{n \to \infty} g_n(x)\text{d}x=\int_E \frac{\partial}{\partial t}f(x, t)\text{d}x<\infty.\] Therefore, \(\displaystyle \int_E f(x, t)\text{d}t\) is differentiable on \((a, b)\), and \[\frac{\text{d}}{\text{d}t}\int_E f(x, t)\text{d}x=\int_E \frac{\partial}{\partial t}f(x, t)\text{d}x.\]

\(\square\)

4. Riemann Integral and Lebesgue Integral

Recall 4.1    Suppose \(f\) is bounded on \([a, b]\). A partition sequence of \([a, b]\), \(\{P^{(n)}\}\) is a finite sequence of values \(x_i\) such that \[a=x_0^{(n)}<x_1^{(n)}<\cdots<x_{k_n}^{(n)}=b.\] Let \(M_i^{(n)}=\sup\{f(x): x_{i-1}^{(n)} \leq x \leq x_i^{(n)}\}\) and \(m_i^{(n)}=\inf\{f(x): x_{i-1}^{(n)} \leq x \leq x_i^{(n)}\}\). The upper Darboux sum of \(f\) with respect to \(P^{(n)}\) is \[S(f, P^{(n)})=\sum_{i=1}^{k_n} M_i^{(n)}(x_i^{(n)}-x_{i-1}^{(n)}).\] The lower Darboux sum of \(f\) with respect to \(P^{(n)}\) is \[s(f, P^{(n)})=\sum_{i=1}^{k_n} m_i^{(n)}(x_i^{(n)}-x_{i-1}^{(n)}).\] The upper Darboux integral of \(f\) is \[\overline{\int_a^b} f(x)\text{d}x=\inf_{P^{(n)}} S(f, P^{(n)})=\lim_{n \to \infty} S(f, P^{(n)}).\] The lower Darboux integral of \(f\) is \[\underline{\int_a^b} f(x)\text{d}x=\sup_{P^{(n)}} s(f, P^{(n)})=\lim_{n \to \infty} s(f, P^{(n)}).\] \(f(x)\) is Riemann integrable on \([a, b]\) if and only if \[\overline{\int_a^b} f(x)\text{d}x=\underline{\int_a^b} f(x)\text{d}x.\]

Recall 4.2    The amplitude at \(x\) is \[\omega(x)=\lim_{\delta \to 0^+}\sup\{|f(y)-f(z)|: y, z \in (x-\delta, x+\delta) \cap [a, b]\}.\]

Theorem 4.1    A bounded function \(f\) is Riemann integrable on \([a, b]\) if and only if \(f\) is continuous almost everywhere on \([a, b]\).

Proof. Let \[h_n(x)=\begin{cases} M_i^{(n)}-m_i^{(n)}, &x_{i-1}^{(n)}<x<x_i^{(n)} \\ 0, &x=x_{i-1}^{(n)} \vee x=x_i^{(n)} \end{cases}.\] We know \(h_n\) is a non-negative measurable function. Suppose \(|f(x)| \leq M\), \(x \in [a, b]\), then \[|h_n(x)| \leq M_i^{(n)}+m_i^{(n)} \leq 2M.\] Since \(g(x)=2M\) is Lebesgue integrable on \([a, b]\), then \[\begin{aligned} \overline{\int_a^b} f(x)\text{d}x-\underline{\int_a^b} f(x)\text{d}x&=\lim_{n \to \infty} (S(f, P^{(n)})-s(f, P^{(n)})) \\&=\lim_{n \to \infty} \sum_{i=1}^{k_n} (M_i^{(n)}-m_i^{(n)})(x_i^{(n)}-x_{i-1}^{(n)}) \\&=\lim_{n \to \infty} \int_{[a, b]} h_n(x)\text{d}x \\&=\int_{[a, b]} \lim_{n \to \infty} h_n(x)\text{d}x & \text{(DCT)} \\&=\int_{[a, b]} \omega(x)\text{d}x=0. \end{aligned}\]

Therefore, \(f\) is Riemann integrable on \([a, b]\) if and only if \[\begin{aligned} &\quad \quad \int_{[a, b]} \omega(x)\text{d}x=0 \\&\Leftrightarrow \omega(x)=0\ \text{a.e.}\ x \in [a, b] \\&\Leftrightarrow \text{$f$ is continuous almost everywhere on $[a, b]$.} \end{aligned}\]

\(\square\)

Theorem 4.2    If a bounded function \(f\) is Riemann integrable on \([a, b]\), then \(f\) is Lebesgue integrable on \([a, b]\), and \[\int_{[a, b]} f(x)\text{d}x=\int_a^b f(x)\text{d}x.\]

Proof. Since \(f\) is Riemann integrable on \([a, b]\), then \(f\) is continuous almost everywhere on \([a, b]\). Hence, \(f\) is a bounded measurable function on \([a, b]\), and thus \(f \in L([a, b])\).

Let \[g_n(x)=\begin{cases} M_i^{(n)}, &x_{i-1}^{(n)}<x<x_i^{(n)} \\ 0, &x=x_{i-1}^{(n)} \vee x=x_i^{(n)} \end{cases},\] where \(\displaystyle \lim_{n \to \infty} g_n(x)=f(x)\ \text{a.e.}\ x \in [a, b]\), and \(g_n\) is bounded. We have \[\begin{aligned} \int_a^b f(x)\text{d}x&=\lim_{n \to \infty} \sum_{i=1}^{k_n} M_i^{(n)}(x_i^{(n)}-x_{i-1}^{(n)}) \\&=\lim_{n \to \infty} \int_{[a, b]} g_n(x)\text{d}x \\&=\int_{[a, b]} \lim_{n \to \infty} g_n(x)\text{d}x \\&=\int_{[a, b]} f(x)\text{d}x. \end{aligned}\]

\(\square\)

Theorem 4.4    Suppose \(f\) is a non-negative real function on \([a, b]\). If for all \(A>a\), \(f\) is Riemann integrable on \([a, A]\), and \(\displaystyle \int_a^{\infty} f(x)\text{d}x<\infty\), then \(f\) is Lebesgue integrable on \([a, \infty)\), and \[\int_{[a, \infty)} f(x)\text{d}x=\int_a^\infty f(x)\text{d}x.\]

Proof. Take an arbitrary sequence \(\{A_n\}\) such that \(A_n>a\) and \(\displaystyle \lim_{n \to \infty} A_n=\infty\). Hence, \[\int_{[a, \infty)} f(x)\text{d}x=\lim_{n \to \infty} \int_{[a, A_n]} f(x)\text{d}x=\lim_{n \to \infty}\int_a^{A_n} f(x)\text{d}x=\int_a^\infty f(x)\text{d}x.\]

In addition, if \(\displaystyle \int_a^\infty f(x)\text{d}x\) diverges, then \(\displaystyle \int_{[a, \infty)} f(x)\text{d}x=\infty\).

\(\square\)

If \(f\) is not non-negative function, then Lebesgue integral is not a generalization of Riemann improper integral.

5. Geometric Interpretation of Lebesgue Integral

Definition 5.1    Suppose \(f\) is a non-negative function on \(E \subset \mathbb{R}^n\). We define \[G(E, f)=\{(x, z): x \in E, 0 \leq z<f(x)\} \subset \mathbb{R}^{n+1}.\]

Theorem 5.1    Suppose \(f\) is a non-negative function on \(E \subset \mathbb{R}^n\), then

    (1) \(f\) is measurable on \(E\) if and only if \(G(E, f)\) is measurable;

    (2) If \(f\) is measurable on \(E\), then \[\int_E f(x)\text{d}x=m(G(E, f)).\]

6. Fubini's Theorem

Theorem 6.1 (Fubini's Theorem)    Suppose \(A \subset \mathbb{R}^p\) and \(B \subset \mathbb{R}^q\) are measurable. If \(f\) is a non-negative measurable function on \(A \times B\), then for \(\text{a.e.} x \in A\), \(f(x, y)\) (as a function of \(y\)) is measurable on \(B\), and \[\int_{A \times B} f(x, y)\text{d}x\text{d}y=\int_A \left(\int_B f(x, y)\text{d}y\right)\text{d}x.\] If \(f\) is Lebesgue integrable on \(A \times B\), then for \(\text{a.e.} x \in A\), \(f(x, y)\) (as a function of \(y\)) is Lebesgue integrable on \(B\), \(\displaystyle \int_B f(x, y)\text{d}y\) (as a function of \(x\)) is Lebesgue integrable on \(A\), and \[\int_{A \times B} f(x, y)\text{d}x\text{d}y=\int_A \left(\int_B f(x, y)\text{d}y\right)\text{d}x.\]

Example 6.1    Let \(\displaystyle f(x, y)=\frac{x^2-y^2}{(x^2+y^2)^2}\) and \(E=(0, 1) \times (0, 1)\). We have \[\int_{(0, 1)} \left(\int_{(0, 1)} f(x, y)\text{d}y\right)\text{d}x=\frac{\pi}{4}\] and \[\int_{(0, 1)} \left(\int_{(0, 1)} f(x, y)\text{d}x\right)\text{d}y=-\frac{\pi}{4}.\] By Fubini's theorem, \(f\) is not Lebesgue integrable.