Chapter 2 Integration Theory

2.1 The lebesgue integral: basic properties and convergence theorems

2.1.1 Stage one: simple functions

Proposition 2.1 The integral of simpe functions defined bve satifies the following properties:

  1. Independence of the representation. If \(\varphi=\sum_{k=1}^Na_k\chi_{E_k}\) is any representation of \(\varphi\), then \[ \int \varphi=\sum_{k=1}^Na_km(E_k). \]
  2. Linearity.
  3. Additivity.
  4. Monotonicity.
  5. Triangle inequality. If \(\varphi\) is a simple function, then so is \(|\varphi|\), and \[ \left|\int\varphi\right|\le \int|\varphi|. \]

2.1.2 Stage two: bounded functions supported on a set of finite measure

Lemma 2.1 Let \(f\) be a bounded function supported on a set \(E\) of finite measure. If \(\{\varphi_n\}_{n=1}^\infty\) is any sequence of simple functions bounded by \(M\), supported on \(E\), and with \(\varphi_n(x)\to f(x)\) for a.e. \(x\), then:

  1. The limit \(\lim_{n\to\infty}\int\varphi_n\) exists.
  2. If \(f=0\) a.e., then the limit \(\lim_{n\to\infty}\int\varphi_n\) equals 0.
Proof. Setting \(I_n=\int\varphi_n\) and applying Egorov’s theorem which is proven in Chapter 1 we have that for and large \(n\) and \(m\) \[ \begin{aligned} |I_n-I_m| &\le \int_E|\varphi_n-\varphi_m|\\ &=\int_{A_\epsilon}|\varphi_n-\varphi_m|+\int_{E-A_\epsilon}|\varphi_n-\varphi_m|\\ &\le \int_{A_\epsilon}\epsilon\ dx+\int_{E-A_\epsilon}2M\ dx\\ &\le m(E)\epsilon+2M\epsilon. \end{aligned} \] given any \(\epsilon>0\). This proves that \(\{I_n\}\) os a Cauchy sequence nd hence converges. If \(f=0\), letting \(m\) tend to infinity we have \(|I_n-f|=|I_n|\le m(E)\epsilon+2M\epsilon\), which yields \(\lim_{n\to\infty}I_n=0\).

For a bounded function \(f\) that is supported on sets of finite measure, we define its Lebesgue integral by \[ \int f=\lim_{n\to\infty}\int\varphi_n. \] where \(\{\varphi_n\}\) is any sequence of simple functions satisfying: \(|\varphi_n|\le M\), each \(\varphi_n\) is supported on the support of \(f\), and \(\varphi_n(x)\to f(x)\) for a.e. \(x\) as \(n\) tends to infinity.

Next, we must show that \(\int f\) is independent of the limiting sequence \(\{\varphi_n\}\) used, in order for the integral to be well-defined. Suppose that \(\{\psi_n\}\) is another sequence of simple functions that satisfies the properties above. Then, if \(\eta_n = \varphi_n- \psi_n\), the sequence \(\{\eta_n\}\) consists of simple functions bounded by \(2M\), supported on a set of finite measure, and such that \(\eta_n\to 0\) a.e. as \(n\) tends to infinity. Applying the lemma we find \[ \lim_{n\to\infty}\int \varphi_n =\lim_{n\to\infty}\int \psi_n+\lim_{n\to\infty}\int \eta_n=\lim_{n\to\infty}\int \psi_n \] as desired.

Proposition 2.2 Suppose \(f\) and \(g\) are bounded functions supported on sets of finite measure. Then the following properties hold.

  1. Linearity.
  2. Additivity.
  3. Monotonicity.
  4. Triangle inequality.

Theorem 2.1 (Bounded convergence theorem)

Suppose that \(\{f_n\}\) is a sequence of measurable functions that are all bounded by \(M\), are supported on a set \(E\) of finite measure, and \(f_n(x)\to f(x)\) a.e. \(x\) as \(n\to\infty\). Then \(f\) is measurable, bounded ,supported on \(E\) for a.e. \(x\), and \[ \int|f_n-f|\to 0 \text{ as } n\to\infty. \] Consequently, \[ \int f_n\to \int f \text{ as } n\to\infty. \]
Proof. The proof is a reprise of the argument in Lemma 2.1. Given \(\epsilon>0\), we may find, by Egorov’ theorem, \[ \begin{aligned} \int|f_n-f| &\le \int_{A_\epsilon}|f_n-f|+\int_{E-A_\epsilon}|f_n-f|\\ &\le m(E)\epsilon+2M\epsilon. \end{aligned} \] for all large \(n\).

2.1.3 Return to Riemann integrable functions

Theorem 2.2

Suppose \(f\) os Riemann integrable on the closed interval \([a,b]\). Then \(f\) is measurable, and \[ \cal \int_{[a,b]}^Rf(x)\ dx=\int_{[a,b]}^Lf(x)\ dx. \]
Proof. By definition of Riemann integrability, \(f\) is bounded, say \(|f(x)|\le M\), and we may construct two consequences of step functions \(\{\varphi_k\}\) and \(\{\psi_k\}\) that satisfy the following properties: \(|\varphi_k(x)\le M|\) and \(|\psi_k(x)\le M|\) for all \(x\in[a,b]\) and \(k\ge 1\), \[ \varphi_1(x)\le \varphi_2(x) \le \cdots\le f\le\cdots\le \psi_2(x)\le \psi_1, \] and \[ \lim_{k\to\infty}\cal \int_{[a,b]}^R\varphi_k(x)\ dx=\lim_{k\to\infty}\cal \int_{[a,b]}^R\psi_k(x)\ dx=\cal \int_{[a,b]}^Rf(x)\ dx. \] Notice that \[ \cal \int_{[a,b]}^R\varphi_k(x)\ dx=\int_{[a,b]}^L\varphi_k(x)\ dx, \] and \[ \cal \int_{[a,b]}^R\psi_k(x)\ dx=\cal \int_{[a,b]}^L\psi_k(x)\ dx. \] Consider Theorem 2.1 (Bounded convergence theorem) and you will complete the proof.

2.1.4 Stage three: non-negative functions

In the case of such a function \(f\) we define its Lebesgue integral by \[ \int f =\sup_g\int g. \]

Proposition 2.3 The integral of non-negative measurable functions enjoys the following properties:

  1. Linearity.
  2. Additivity.
  3. Monotonicity.
  4. If \(g\) is integrable and \(0\le f\le g\), then \(f\) is integrable.
  5. If \(f\) is integrable, then \(f(x)<\infty\) for a.e. \(x\).
  6. If \(\int f=0\), then \(f(x)=0\) for a.e. \(x\).
    Proof. We just prove the first assertion. Let \(\varphi\), \(\psi\) and \(\eta\) be non-negative functions bounded and supported on sets of finite measure, where \(\varphi\le f\), \(\psi\le g\) and \(\eta\le f+g\), then \(\varphi+\psi\le f+g\). Consequently, \[ \int f+ \int g\le \int (f+g). \] On the other hand, if we define \(\eta_1=\min(f(x),\eta(x))\) and \(\eta_2=\eta-\eta_1\), then \[ \int \eta=\int(\eta_1+\eta_2)=\int \eta_1+\int \eta_2\le\int f+ \int g, \] which means that \[ \int (f+g)\le \int f+ \int g. \]

Lemma 2.2 (Fatou)

Suppose \(\{f_n\}\) is a sequence of measurable functions with \(f_n\ge 0\). If \(\lim_{n\to\infty}f_n(x)=f(x)\) for a.e. \(x\), then \[ \int f\leq \liminf_{n\to\infty}\int f_n. \]
Proof. Suppose \(0\le g\le f\), where \(g\) is bounded and supported on a set \(E\) of finite measure. If we set \(g_n(x)=\min(g(x),f_n(x))\), then \(g_n\to g\) a.e. as \(n\to \infty\). By Theorem 2.1 (Bounded convergence theorem) we have \[ \int g_n\to \int g. \] Since \(g_n\le f_n\), we have \(\int f_n\le \int g_n\), so that \[ \int g\le \liminf_{n\to\infty}\int f_n. \] Taking the supremum over all \(g\) yields the desired inequality.

Corollary 2.1 Suppose \(f\) is a non-negative measurable function, and \(\{f_n\}\) a sequence of non-negative measurable functions with \(f_n(x) \le f(x)\) and \(f_n(x)\to f(x)\) for a.e. \(x\). Then \[ \lim_{n\to\infty}\int f_n=\int f. \]

Corollary 2.2 (Monotone convergence theorem) Suppose \(\{f_n\}\) a sequence of non-negative measurable functions with \(f_n(x) \nearrow f(x)\). Then \[ \lim_{n\to\infty}\int f_n=\int f. \]

Corollary 2.3 Consider a series \(\sum_{k=1}^{\infty}a_k(x)\), where \(a_k(x)\ge 0\) is measurable for every \(k\ge 1\). Then \[ \int \sum_{k=1}^{\infty}a_k(x)\ dx =\sum_{k=1}^{\infty}\int a_k(x)\ dx. \]

2.1.5 General form

In this case, we define the Lebesgue integral of \(f\) by \[ \int f =\int f^+-\int f^-. \]

Proposition 2.4 The integral of Lebesgue integrable functions is linear, additive, monotonic, and satisfies the triangle inequality.

Proposition 2.5 Suppose \(f\) is integral on \(\mathbb{R}^d\). Then for every \(\epsilon>0\):

  1. There exists a set of finite measure \(B\) (a ball, for example) such that \[ \int_{B^c}|f|<\epsilon. \]
  2. There is a \(\delta>0\) such that \[ \int_E|f|<\epsilon\quad \text{ whenever }m(E)<\delta. \]
    Proof. Assume that \(f\ge 0\): 1. \(B_N=(-N,N)\); 2. \(E_N=\{x:f(x)\le N\}\).

Theorem 2.3 (Dominated convergence theorem)

Suppose \(\{f_n\}\) is a sequence of measurable functions such that \(f_n\to f\) a.e. \(x\) as \(n\) tends to infinity. If \(f_n(x)\le g(x)\), where \(g\) is integrable, then \[ \int |f_n-f|\to 0 \quad \text{ as }n\to \infty, \] and consequently \[ \int f_n \to \int f. \]
Proof. For each \(N\ge 0\) let \(E_N=\{x:|x|\le N, g(x)\le N\}\). \[ \begin{aligned} \int|f_n-f|&=\int_{E_N}|f_n-f|+\int_{E_N^c}|f_n-f|\\ &\le \int_{E_N}|f_n-f|+2\int_{E_N^c}g\\ &\le \epsilon +2\epsilon \end{aligned} \] for all large \(n\). This prove the theorem.

2.1.6 Complex-valued functions

The collection of all complex-valued integrable functions on a measurable subset \(E\subset \mathbb{R}^d\) forms a vector space over \(\mathbb{C}\).

2.2 The space \(L^1\) of integrable functions

For any integrable function \(f\) on \(\mathbb{R}^d\) we define the \(L^1\)-norm of \(f\), \[ \|f\|=\|f\|_{L^1}=\|f\|_{L^1(\mathbb{R}^d)}=\int_{\mathbb{R}^d}|f|. \]

Proposition 2.6 Suppose \(f\) and \(g\) are two functions in \(L^1(\mathbb{R}^d)\).

  1. \(\|af\|=|a|\|f\|\) for all \(a\in\mathbb{C}\).
  2. \(\|f+g\|\le \|f\|+\|g\|\).
  3. \(\|f\|=0\) iff \(f=0\) a.e.
  4. \(d(f,g)=\|f-g\|\) defines a metric on \(L^1(\mathbb{R}^d)\).

Theorem 2.4 (Riesz-Fischer) The vector space \(L^1\) is complete in its metric.

Corollary 2.4 If \(\{f_n\}_{n=1}^\infty\) converges to \(f\) in \(L^1\), the nthere exists a subsequence \(\{f_{n_k}\}_{k=1}^\infty\) such that \[ f_{n_k}(x)\to f(x)\quad a.e. x. \]

We say that a family \(\cal G\) of integrable functions is dense in \(L^1\) if for ant \(f\in L^1\) and \(\epsilon>0\), there exists \(g\in\cal G\) so that \(\|f-g\|<\epsilon\).

Theorem 2.5 The following families of functions are dense in \(L^1(\mathbb{R}^d)\):

  1. The simple functions.
  2. The step functions.
  3. The continuous functions of compact support.

2.2.1 Invariance properties

2.2.2 Translations and continuity

Proposition 2.7

Suppose \(f\in L^1(\mathbb{R}^d)\). Then \[ \|f_h-f\|\to 0\quad \text{ as }h\to 0. \]
Proof. By Theorem 2.5, find a continuous function with compact support to approximate \(f\).

2.3 Fubini’s theorem

2.4 A Fourier inversion formula

2.5 Exercise

2.6 Problem

:::