-
Notifications
You must be signed in to change notification settings - Fork 1
/
Copy pathlecture-26.tex
237 lines (208 loc) · 15.7 KB
/
lecture-26.tex
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
66
67
68
69
70
71
72
73
74
75
76
77
78
79
80
81
82
83
84
85
86
87
88
89
90
91
92
93
94
95
96
97
98
99
100
101
102
103
104
105
106
107
108
109
110
111
112
113
114
115
116
117
118
119
120
121
122
123
124
125
126
127
128
129
130
131
132
133
134
135
136
137
138
139
140
141
142
143
144
145
146
147
148
149
150
151
152
153
154
155
156
157
158
159
160
161
162
163
164
165
166
167
168
169
170
171
172
173
174
175
176
177
178
179
180
181
182
183
184
185
186
187
188
189
190
191
192
193
194
195
196
197
198
199
200
201
202
203
204
205
206
207
208
209
210
211
212
213
214
215
216
217
218
219
220
221
222
223
224
225
226
227
228
229
230
231
232
233
234
235
236
237
\documentclass[a4paper,english,12pt]{article}
\input{header}
\newcommand{\goodchi}{\protect\raisebox{2.5pt}{$\chi$}}
\begin{document}
\title{Lecture 26: Dominated Convergence Theorem}
\author{}
\maketitle
\textit{Continuation of Fatou's Lemma.}
\begin{cor} If $f\in L^+$ and $\{f_n\in L^+:n\in \mathbb{N}\}$ is any sequence of functions such that $f_n\rightarrow f$ almost everywhere, then
\begin{align*}
\int_X f \leqslant \liminf \int_X f_n.
\end{align*}
\label{fatou_ae}
\end{cor}
\begin{proof} Let $f_n\rightarrow f$ everywhere in $X$. That is, $\liminf f_n(x)=f(x)$ ($=\limsup f_n$ also) for all $x\in X$. Then, by Fatou's lemma,
\begin{align*}
\int_X f =\int_X \liminf f_n \leqslant \liminf \int_X f_n.
\end{align*}
If $f_n\nrightarrow f$ everywhere in $X$, then let $E=\{x\in X:\liminf f_n(x)\neq f(x)\}$. Since $f_n\rightarrow f$ almost everywhere in $X$, $\mu (E)=0$ and
\begin{align*}
\int_X f =\int_{X-E} f \quad\text{and}\quad \int_X f_n =\int_{X-E} f_n \;\;\forall n
\end{align*}
thus making $f_n\nrightarrow f$ everywhere in $X-E$. Hence,
\begin{align*}
\int_X f \quad=\quad \int_{X-E} f \quad\leqslant\quad \liminf \int_{X-E} f_n \quad=\quad \liminf \int_X f_n.
\end{align*}
\end{proof}
\begin{exmp}[Strict inequality] Let $S_n=[n,n+1]\subset \mathbb{R}$ and $f_n=\goodchi _{S_n}$. Then, $f=\liminf f_n\,=\,0$ and
\begin{align*}
0 \;=\; \int_{\mathbb{R}}f\,d\mu \;<\; \liminf \int_{\mathbb{R}}f_n\,d\mu \;=\;1.
\end{align*}
\end{exmp}
\begin{exmp}[Importance of non-negativity] Let $S_n=[n,n+1]\subset \mathbb{R}$ and $f_n=-\goodchi _{S_n}$. Then, $f=\liminf f_n\,=\,0$ but
\begin{align*}
0 \;=\; \int_{\mathbb{R}}f\,d\mu \;>\; \liminf \int_{\mathbb{R}}f_n\,d\mu \;=\;-1.
\end{align*}
\end{exmp}
\begin{prop} If $f\in L^+$ and $\int_X f\,d\mu < \infty$, then \textit{(a)} the set $A=\{x\in X: f(x)=\infty\}$ is a null set and \textit{(b)} the set $B=\{x\in X:f(x)>0\}$ is $\sigma$-finite.
\label{int_nll_fnt}
\end{prop}
\begin{proof} Recall that for any $f\in L^+$
\begin{align*}
\int_X f\,d\mu \;=\; \sup \left\{\int_X \phi\,d\mu \; : \; 0\leqslant \phi \leqslant f, \; \phi \;\;\text{simple}\right\}
\end{align*}
and for a simple function $\psi$, $\int_X \psi\,d\mu\,=\,\sum _{j=1}^n a_j\mu(E_j)$, where $E_j=\psi^{-1}(\{a_j\})$ and $\{a_1,a_2,\ldots,a_n\}$ is the the range of $\psi$.
\begin{enumerate}[\itshape(a)]
\item Assume, on the contrary, that $\mu(A)>0$ and let $I=\int_X f\,d\mu\,<\,\infty$. Define a simple function $\phi$ as
\begin{align*}
\phi = 2\frac{I}{\mu(A)} \goodchi _A.
\end{align*}
Since $f(x)=\infty$ for all $x\in A$ and $\phi(x)=0$ for all $x\in X-A$, $\phi(x)\leqslant f(x)$ for all $x\in X$ and therefore, $\int_X \phi\,d\mu\leqslant \int_X f\,d\mu$. But $\int_X \phi\,d\mu=2I>I=\int_X f\,d\mu$, which is a contradiction. Thus, $\mu(A)=0$.
\item A $\sigma$-finite set is a countable union of sets of finite measure. Define $B_n=\{x\in X\,:\,f(x)>\,n^{-1}\}$. Then, $B$ is a countable union of $B_n$s. For each $B_n$, define the simple functions $\phi_n = n^{-1} \goodchi _{B_n}$. For all $n\in \mathbb{N}$, $\phi_n\leqslant f$, and hence $\int_X \phi_n\,d\mu \,\leqslant \, \int_X f\,d\mu\,<\,\infty$. Since $\int_X \phi_n\,d\mu\,=\,n^{-1}\mu(B_n)$, $\mu(B_n)<\infty$. Thus, $B=\cup _{n\in \mathbb{N}}B_n$ is $\sigma$-finite.
\end{enumerate}
\end{proof}
\section{Integration of Real-Valued Functions}
We now discuss integration of real-valued function which need not be positive. Let $f^+$ and $f^-$ be the positive and negative parts of $f$ respectively, where
\begin{align*}
f^+(x)=\max\{f(x),0\} \quad\text{and}\quad f^-(x)=\max\{-f(x),0\} \quad\text{for all}\quad x\in X.
\end{align*}
Then, $f\,=\,f^+\,-\,f^-$. Note that both $f^+$ and $f^-$ are positive real-valued functions. If at least one of $\int_X f^+\,d\mu$ and $\int_X f^-\,d\mu$ is finite, then we define
\begin{align*}
\int_X f\,d\mu \;=\; \int_X f^+\,d\mu \;-\; \int_X f^-\,d\mu.
\end{align*}
If both $\int_X f^+\,d\mu$ and $\int_X f^-\,d\mu$ are finite, then $f$ is said to be \textbf{integrable}. Since $|f|=f^+\,+\,f^-$, $f$ is integrable iff $\int_X |f|\,d\mu<\infty$.
\begin{prop} The set of integrable real-valued functions on $X$, denoted by $\mathcal{F}(X,\mathbb{R})$, is a real vector space, and the integral is linear functional on it.
\end{prop}
\begin{proof} To prove that $\mathcal{F}(X,\mathbb{R})$ is a vector space, it suffices to prove that any linear combination of integrable real-valued functions in $\mathcal{F}(X,\mathbb{R})$ also in $\mathcal{F}(X,\mathbb{R})$. For any $f,g\in \mathcal{F}(X,\mathbb{R})$ and $a,b\in\mathbb{R}$, $|af+bg|\leqslant|a||f|+|b||g|$ (by triangle inequality). Hence, $\int_X|af+bg|\,d\mu\,\leqslant\,\int_X(|a||f|+|b||g|)\,d\mu\,=\,|a|\int_X|f|\,d\mu\,+\,|b|\int_X|g|\,d\mu\,<\infty$ since both $f$ and $g$ are integrable. Thus, $af+bg$ is also in $\mathcal{F}(X,\mathbb{R})$.
To prove that the functional $I:f\longmapsto \int_X f\,d\mu,\,f\in \mathcal{F}(X,\mathbb{R})$, is linear, we need to show that \textit{(a)}$I(cf)=cI(f)$ and \textit{(b)}$I(f+g)=I(f)+I(g)$ for all $f,g\in \mathcal{F}(X,\mathbb{R})$.
\begin{enumerate}[\itshape(a)]
\item We will use the facts that $(cf)^+=cf^+$ and $(cf)^-=cf^-$ for $c\geqslant 0$ and that $(cf)^+=|c|f^-$ and $(cf)^-=|c|f^+$ for $c<0$. Recall that for any $g\in L^+$, $I(cg)=cI(g)$ and for any $f\in \mathcal{F}(X,\mathbb{R})$, both $f^+$ and $f^-$ are positive. Let $c\geqslant 0$. Then, using the above facts, $I(cf)=c(I(f^+)-I(f^-))= cI(f)$. For $c<0$, $I(cf)=I((cf)^+)-I((cf)^-)=|c|(I(f^-)-I(f^+))=-|c|I(f)=cI(f)$.
\item Let $f,g\in \mathcal{F}(X,\mathbb{R})$ and $h=f+g$. Then, $h^+-h^-=f^+-f^-+g^+-g^-$ and consequently $h^++f^-+g^-=h^-+f^++g^+$. Recall that if $\{f_n\}$ is a finite of infinite sequence in $L^+$ and $f=\sum_nf_n$, then $\int f=\sum_n\int f_n$. So,
\begin{align*}
\int h^+ + \int f^- + \int g^- \;=\;\int h^- + \int f^+ + \int g^+.
\end{align*}
Rearranging the terms above, we get
\begin{align*}
\int h\;=\;\int h^+-\int h^-\;=\;\int f^+-\int f^-+\int g^+-\int g^-\;=\;\int f+\int g.
\end{align*}
\end{enumerate}
\end{proof}
\begin{prop}For any $f\in \mathcal{F}(X,\mathbb{R})$, $|\int f|\leqslant \int |f|$.
\label{int_mod}
\end{prop}
\begin{proof}If $\int f=0$, then this is trivial. For any real $f$, $|\int f|=|\int f^+-\int f^-|\leqslant |\int f^+| + |\int f^-|=\int f^+ + \int f^- = \int |f|$ (by triangle inequality).
\end{proof}
\begin{prop}
\begin{enumerate}[\itshape(a)]
\item For any $f\in \mathcal{F}(X,\mathbb{R})$, $A=\{x:f(x)\neq 0\}$ is $\sigma$-finite.
\item If $f,g\in \mathcal{F}(X,\mathbb{R})$, then $\int_Ef=\int_Eg$ for all $E\in \mathcal{M}$ iff $\int |f-g|=0$ iff $f=g$ almost everywhere.
\end{enumerate}
\end{prop}
\begin{proof}
\begin{enumerate}[\itshape(a)]
\item Note that $A=A^+\cup A^-$, where $A^+=\{x:f^+(x)>0\}$ and $A^-=\{x:f^-(x)>0\}$. Since both $f^+$ and $f^-$ are in $L^+$, by Proposition~\ref{int_nll_fnt}, both $A^+$ and $A^-$ are $\sigma$-finite. Hence, $A$ is $\sigma$-finite.
\item The second equivalence follows from the fact that for any $h\in L^+$, $\int h=0$ iff $h=0$ almost everywhere. If $\int |f-g|=0$, then by Proposition~\ref{int_mod}, for any $E\in \mathcal{M}$,
\begin{align*}
\left|\int_Ef-\int_Eg\right|\;\leqslant\;\int_X\goodchi_E|f-g|\;\leqslant\;\int_X|f-g|=0
\end{align*}
so that $\int_Ef=\int_Eg$. Let $h=f-g$ and assume that $f=g$ almost everywhere is false, then at least one of $h^+$ and $h^-$ must be nonzero on a set of positive measure. Let $E=\{x:h^+(x)>0\}$ be one such set; note that $h^-(x)=0$ and hence, $\int_Eh^-(x)=0$ for all $x\in E$. Then, $\int_Ef-\int_Eg=\int_Eh=\int_Eh^+>0$. Similar conclusion can be drawn for $h^-$ being nonzero on a set of positive measure.
\end{enumerate}
\end{proof}
\begin{rem}
\begin{enumerate}[\itshape(i)]
\item Altering functions on a mull set does not alter their integration.
\item Let $E\in\mathcal{M}$. Then, it is possible to integrate $f$ by defining $f|_{E^{\mathsf{c}}}=0$.
\item It is possible to treat $\overline{\mathbb{R}}$-valued functions that are finite almost everywhere as $\mathbb{R}$-valued functions.
\end{enumerate}
\end{rem}
\begin{defn}$L^1$ can be redefined as follows:
\begin{align*}
L^1(\mu)=\{\text{Equivalence class of almost everywhere-defined integrable functions on}\; X\},
\end{align*}
where two functions $f$ and $g$ are equivalent if $\mu(\{x\in X:f(x)\neq g(x)\})=0$.
\end{defn}
\begin{rem}
\begin{enumerate}[\itshape(i)]
\item $L^1(\mu)$ is still a vector space (under pointwise almost everywhere addition and scalar multiplication).
\item $f\in L^1(\mu)$ will mean that $f$ is an almost everywhere-defined integrable function.
\item For any two $f,g\in L^1(\mu)$, define $\rho(f,g)=\int |f-g|\,d\mu$. This is a metric, since it is symmetric, satisfies triangle inequality, and is $0$ if $f$ and $g$ are equal almost everywhere. This definition allows $L^1(\mu)$ to be a metric space with $\rho(f,g)$ as the metric.
\end{enumerate}
\end{rem}
\begin{thm}[The Dominated Convergence Theorem] Let $\{f_n\in L^1:n\in \mathbb{N}\}$ be a sequence of functions such that \textit{(a)} $f_n\rightarrow f$ almost everywhere and \textit{(b)} there exists a non-negative $g\in L^1$ such that $|f_n|\leqslant g$ almost everywhere for all $n\in \mathbb{N}$. Then, $f\in L^1$ and $\int f=\lim_{n\rightarrow\infty}\int f_n$.
\end{thm}
\begin{rem}
\begin{enumerate}[\itshape(i)]
\item $\int\lim_{n\rightarrow\infty} f_n=\lim_{n\rightarrow\infty}\int f_n$ is an equivalent statement.
\item Here $g$ \textbf{dominates} $f_n$s.
\end{enumerate}
\end{rem}
\begin{proof} Since $f$ is the limit of measurable functions $\{f_n\}$ almost every where, it is measurable. Since $|f_n|\leqslant g$ almost everywhere, $|f|=\lim_{n\rightarrow\infty}f_n\leqslant g$ almost everywhere, and hence, $f\in L^1$. Furthermore, $g+f_n\leqslant 0$ almost everywhere and $g-f_n\leqslant 0$ almost everywhere. By Corollary~\ref{fatou_ae}
\begin{align*}
\int g+\int f=\int (g+f)\leqslant \liminf \int (g+f_n)=\int g + \liminf \int f_n,
\end{align*}
or
\begin{align}
\int f\leqslant \liminf \int f_n.
\label{f_inf}
\end{align}
Using Corollary~\ref{fatou_ae} for $g-f_n$ we obtain
\begin{align*}
\int g-\int f=\int (g-f)\leqslant \liminf \int (g-f_n)=\int g - \liminf \int f_n,
\end{align*}
or
\begin{align}
\int f\geqslant \limsup \int f_n.
\label{f_sup}
\end{align}
Since $\liminf \int f_n\leqslant \limsup \int f_n$, using \ref{f_inf} and \ref{f_sup} we get
\begin{align*}
\int f\leqslant \liminf \int f_n\leqslant \limsup \int f_n\leqslant\int f,
\end{align*}
which forces
\begin{align*}
\int f= \liminf \int f_n= \limsup \int f_n\leqslant\int f=\lim_{n\rightarrow\infty}\int f_n.
\end{align*}
as claimed.
\end{proof}
\begin{thm}Let $\{f_n\in L^1:n\in\mathbb{N}\}$ be a sequence of functions such that $\sum_{n\in\mathbb{N}}\int |f_n|<\infty$. Then, $\sum_{n\in\mathbb{N}}f_n$ converges to a function in $L^1$ and $\int \sum_{n\in\mathbb{N}}f_n =\sum_{n\in\mathbb{N}}\int f_n$.
\end{thm}
\begin{proof}Recall that if $\{h_n\}$ is a finite of infinite sequence in $L^+$, then $\int \sum_nh_n=\sum_n\int h_n$. Set $h_n=|f_n|$ and let $g=\sum_{n\in\mathbb{N}}|f_n|$. Then, $\int g=\sum_{n\in\mathbb{N}}\int |f_n|<\infty$ and hence $g\in L^1$.
By Proposition~\ref{int_nll_fnt}, $g(x)(=\sum_{n\in\mathbb{N}}|f_n(x)|)$ is finite for all $\{x:g(x)>0\}$, and for each such $x$ $\sum_{n\in\mathbb{N}}f_n(x)$ converges. Furthermore, the partial sums $F_k\triangleq\sum_{n=1}^{k}f_n\leqslant g$ (by triangle inequality) for all $k$. We can now apply dominated convergence theorem to the sequence of partial sums $F_k$ to obtain
\begin{align*}
\int \lim_{k\rightarrow\infty} F_k=\lim_{k\rightarrow\infty}\int F_k,
\end{align*}
which can be simplified to (using linearity of $\int$)
\begin{align*}
\int \sum_{n\in\mathbb{N}}f_n =\sum_{n\in\mathbb{N}}\int f_n.
\end{align*}
\end{proof}
\begin{thm}If $f\in L^1$ and $\epsilon>0$, then there is an integrable simple function $\phi=\sum a_j\goodchi_{E_j}$ such that $\int |f-\phi|<\epsilon$. (That is, the integrable simple functions are dense in $L^1$ in the $L^1$ metric.)
\end{thm}
\begin{proof}Recall that for any real-valued measurable function $g$, there exists a sequence $\{\psi_n\}$ of simple functions such that $\psi_n\rightarrow g$ and $0\leqslant |\psi_1|\leqslant |\psi_2|\leqslant\ldots\leqslant |g|$ pointwise. Let $\{\phi_n\}$ be as above for $f$. Then, $\phi_n$s are integrable. Since $|\phi_n-f|\leqslant 2|f|$, $\int |\phi_n-f|<\epsilon$ for sufficiently large $n$ by the dominated convergence theorem.
\end{proof}
\section{Modes of Convergence}
Let $(X,\mathcal{M},\mu)$ be a measure space. Let $\{f_n\}$ be a sequence of functions in $L^1$ and $f\in L^1$.
\begin{defn}[Convergence in $L^1$] If $f_n\rightarrow f$ in the metric $\rho(f,g)=\int |f-g|\,d\mu$, then $\{f_n\}$ is said to converge to $f$ in $L^1(\mu)$.
\end{defn}
\begin{lem}$f_n\rightarrow f$ in $L^1$ iff $\lim_{n\in\mathbb{N}}\int |f_n-f|\,d\mu=0$.
\end{lem}
\begin{defn}[Pointwise Convergence] $\{f_n\}$ is said to converge to $f$ pointwise if $f_n(x)$ converges to $f(x)$ for all $x\in X$. In other words, for every $\epsilon>0$ and $x$, there exists an $N_{\epsilon,x}$ such that $|f_n(x)-f(x)|\leqslant\epsilon$ for all $n\geqslant N_{\epsilon,x}$.
\end{defn}
\begin{defn}[Uniform Convergence] $\{f_n\}$ is said to converge to $f$ uniformly if for every $\epsilon>0$, there exists an $N_{\epsilon}$ such that $|f_n(x)-f(x)|\leqslant\epsilon$ for all $n\geqslant N_{\epsilon}$ and $x\in X$.
\end{defn}
\begin{defn}[Almost Everywhere Convergence] $\{f_n\}$ is said to converge to $f$ almost everywhere if $\mu(\{x\in X:\lim_{n\rightarrow\infty}f_n(x)\neq f(x)\})=0$.
\end{defn}
\begin{defn}[Convergence in Measure] $\{f_n\}$ is said to converge to $f$ in measure if for every $\epsilon>0$, $\lim_{n\rightarrow\infty}\mu(\{x\in X:|f_n(x)-f(x)|\geqslant\epsilon\})=0$.
\end{defn}
\begin{defn}[Cauchy Convergence] $\{f_n\}$ is said to be Cauchy in measure if for every $\epsilon >0$, $\mu(\{x\in X:|f_n(x)-f_m(x)|\geqslant 0\})\rightarrow 0$ as $m,n\rightarrow \infty$.
\end{defn}
\begin{thm}If $f_n\rightarrow f$ almost everywhere and $f_n\leqslant g$ for all $n\in \mathbb{N}$ and some $g\in L^1$, then $f_n\rightarrow f$ in $L^1$.
\end{thm}
\begin{proof}Follows from the dominated convergence theorem since $|f_n-f|\leqslant 2g$.
\end{proof}
\begin{prop}If $f_n\rightarrow f$ in $L^1$, then $f_n\rightarrow f$ in measure.
\end{prop}
\begin{proof}Let $E_{n,\epsilon}=\{x:|f_n(x)-f(x)|\geqslant\epsilon\}$. Then, $\int |f_n-f|\geqslant\int_{E_{n,\epsilon}}|f_n-f|\geqslant\epsilon\mu(E_{n,\epsilon})$, and hence $\mu(E_{n,\epsilon})\leqslant\epsilon^{-1}\int|f_n-f|\rightarrow 0$ as $n\rightarrow\infty$.
\end{proof}
\begin{thm}[Erogoff's Theorem, Almost Uniform Convergence]Let $\mu(X)<\infty$ and $f_1,f_2,\ldots,f$ be measurable real-valued functions on $X$ such that $f_n\rightarrow f$ almost everywhere. Then, for every $\epsilon>0$, there exists $E\subseteq X$ such that $\mu(E)<\epsilon$ and $f_n\rightarrow f$ uniformly on $E^{\mathsf{c}}$.
\end{thm}
\begin{rem}$\{f_n\}$ is said to converge to $f$ almost uniformly if for every $\epsilon>0$, there exists $E\in\mathcal{M}$ of measure $\mu(E)<\epsilon$ such that $f_n\rightarrow f$ uniformly on $E^{\mathsf{c}}$.
\end{rem}
\begin{proof}Without loss of generality, assume that $f_n\rightarrow f$ everywhere on $X$. For $k,n\in\mathbb{N}$, let
\begin{align*}
E_n(k)\;=\; \cup_{m=n}^{\infty} \{x\in X: |f_m(x)-f(x)|\geqslant \,^1/_k\}.
\end{align*}
Then, for fixed $k$, $E_n(k)$ decreases as $n$ increases and $\cap_{n=1}^{\infty}E_n(k)=\emptyset$. Since $\mu(X)\infty$, we conclude that $\lim_{n\rightarrow\infty}\mu(E_n(k))=0$. Given $\epsilon>0$ and $k\in\mathbb{N}$, choose $n_k$ large enough that $\mu(E_n(k))<\epsilon 2^{-k}$ and let $E=\cup_{k=1}^{\infty}E_{n_k}(k)$. Then, $\mu(E)<\epsilon$ and we have $|f_n(x)-f(x)|<\,^1/_k$ for $n>n_k$ and $x\notin E$. Thus, $f_n\rightarrow f$ uniformly on $E^{\mathsf{c}}$.
\end{proof}
\end{document}