Note.
By default, $(X,\sigma,\mu)$ is any measure space.
Indicator function
Given a subset $S$ of a set $X$, the indicator function of $S$ is $\mathcal X_S:X\to\{0,1\}$ such that
$\mathcal X_S(x)=1$ if $x\in S$ and $\mathcal X_S(x)=0$ otherwise.
Simple function
Let $U$ be measurable.
A simple function on $U$ is a function $f:U\to\overline R$ such that for some $n\in N$,
for some measurable $S_1,\ldots,S_n\subseteq U$ and $c_1,\ldots,c_n\in R$,
we have $$f=\sum_{k=1}^nc_k\mathcal X_{S_k}$$
An unsigned simple function requires that $c_1,\ldots,c_n\in[0,\infty]$ instead.
Lemma.
If $f$ is a simple function or unsigned simple function on some measurable $U$, then for some $l\in N^+$,
for some measurable partition $A_1,\ldots,A_l$ of $U$ and $c_1,\ldots,c_l\in R$ (for simple $f$) or $c_1,\ldots,c_l\in[0,\infty]$ (for unsigned simple $f$),
we have $$f=\sum_{k=1}^lc_k\mathcal X_{A_k}$$
(show proof)
Proof.
Suppose $f=\sum_{k=1}^nc_k\mathcal X_{S_k}$ where $n\in N$,
$S_1,\ldots,S_n\subseteq U$ are measurable, and $c_1,\ldots,c_n\in R$ (for simple $f$) or $c_1,\ldots,c_n\in[0,\infty]$ (for unsigned simple $f$).
We can partition $U$ into $2^{n}$ partitions depending on whether a point in $U$ is in $S_1,\ldots,S_n$.
Let $A_1,\ldots,A_l$ denote this partition, where $l=2^n\gt0$.
Note that each of $S_1,\ldots,S_n$ is the union of some of $A_1,\ldots,A_l$.
We define $J_k$ to be the set of indexes such that $S_k=\bigcup_{i\in J_k}A_i$.
And we define $g_k:\{1,\ldots,l\}\to\{0,1\}$ such that $g_k(i)=1$ if $i\in J_k$ and $g_k(i)=0$ otherwise.
Let $x\in U$.
Then
$$f(x)
=\sum_{k=1}^nc_k\mathcal X_{S_k}(x)
=\sum_{k=1}^nc_k\sum_{i\in J_k}\mathcal X_{A_i}(x)
=\sum_{k=1}^n\sum_{i\in J_k}c_k\mathcal X_{A_i}(x)
=\sum_{k=1}^n\sum_{i=1}^l\p{c_kg_k(i)}\mathcal X_{A_i}(x)
=\sum_{i=1}^l\sum_{k=1}^n\p{c_kg_k(i)}\mathcal X_{A_i}(x)
=\sum_{i=1}^l\p{\sum_{k=1}^nc_kg_k(i)}\mathcal X_{A_i}(x)$$
Note that each $A_i$ is the intersection of finite measurable sets (including $U$) or their complements with respect to $U$,
and is thus measurable.
$\blacksquare$
Lemma.
Suppose $c_1,\ldots,c_n,c'_1,\ldots,c'_{n'}\in[0,\infty]$.
If $\sum_{k=1}^nc_k\mathcal X_{S_k}=\sum_{k=1}^{n'}c'_k\mathcal X_{S'_k}$,
then $\sum_{k=1}^nc_k\mu(S_k)=\sum_{k=1}^{n'}c'_k\mu(S'_k)$.
(show proof)
Proof.
We can partition $U$ into $2^{n+n'}$ partitions depending on whether a point in $U$ is in $S_1,\ldots,S_n,S'_1,\ldots,S'_{n'}$.
Let $A_1,\ldots,A_l$ denote the non-empty partitions of this partition, and let $a_i\in A_i$.
Since each $A_i$ is the intersection of finite measurable sets (including $U$) or their complements with respect to $U$,
each $A_i$ is measurable.
Note that each of $S_1,\ldots,S_n,S'_1,\ldots,S'_{n'}$ is the union of some of $A_1,\ldots,A_l$.
We define $J_k$ to be the set of indexes such that $S_k=\bigcup_{i\in J_k}A_i$,
and $J'_k$ to be the set of indexes such that $S'_k=\bigcup_{i\in J'_k}A_i$.
Then
$$\sum_{k=1}^nc_k\mu(S_k)
=\sum_{k=1}^nc_km\p{\bigcup_{i\in J_k}A_i}
=\sum_{k=1}^nc_k\sum_{i\in J_k}\mu(A_i)
=\sum_{k=1}^n\sum_{i\in J_k}c_k\mu(A_i)
=\sum_{k=1}^n\sum_{i=1}^lc_k\mu(A_i)\mathcal X_{S_k}(a_i)
=\sum_{i=1}^l\sum_{k=1}^nc_k\mu(A_i)\mathcal X_{S_k}(a_i)
=\sum_{i=1}^l\mu(A_i)\p{\sum_{k=1}^nc_k\mathcal X_{S_k}}(a_i)$$
Similarly, we have $$\sum_{k=1}^{n'}c'_k\mu(S'_k)=\sum_{i=1}^l\mu(A_i)\p{\sum_{k=1}^{n'}c'_k\mathcal X_{S'_k}}(a_i)$$
Since $\sum_{k=1}^nc_k\mathcal X_{S_k}=\sum_{k=1}^{n'}c'_k\mathcal X_{S'_k}$,
we have $$\sum_{k=1}^nc_k\mu(S_k)=\sum_{k=1}^{n'}c'_k\mu(S'_k)$$
$\blacksquare$
Note.
Given measurable $U$, we can define a map $I$ that maps an unsigned simple function $f=\sum_{k=1}^nc_k\mathcal X_{S_k}$
from $U$ to $I(f)=\sum_{k=1}^nc_k\mu(S_k)$. By the above lemma, $I(f)$ is independent on the choice of $n$, $S_1,\ldots,S_n$ and $c_1,\ldots,c_n$.
Measurable function
Let $U$ be measurable.
A function $f:U\to\overline R$ is said to be measurable if there exists a sequence $(f_n)$ of simple functions on $U$
that converges pointwise to $f$.
$f$ is said to be unsigned measurable if every function in $(f_n)$ is an unsigned simple function instead.
Proposition.
Let $U$ be measurable. Then a function $f:U\to\overline R$ is unsigned measurable if and only if it is both unsigned and measurable.
(show proof)
Proof.
Suppose $f$ is unsigned measurable.
Then there exists a sequence $(f_n)$ of unsigned simple functions that converges to $f$.
Thus $f$ is trivially unsigned.
Define $(g_n)$ from $(f_n)$ such that for each $n$ and each $p\in U$,
$g_n(p)=f_n(p)$ if $f_n(p)\in R$ and $g_n(p)=n$ if $f_n(p)=\infty$.
Then each $g_n$ is a simple function and $(g_n)$ converges to $f$, implying $f$ is measurable.
Suppose $f$ is both unsigned and measurable.
Then there exists a sequence $(f_n)$ of simple functions that converges to $f$.
Define $(g_n)$ from $(f_n)$ such that for each $n$ and each $p\in U$,
$g_n(p)=f_n(p)$ if $f_n(p)\ge0$ and $g_n(p)=0$ if $f_n(p)\lt0$.
Then each $g_n$ is an unsigned simple function and $(g_n)$ converges to $f$, implying $f$ is unsigned measurable.
$\blacksquare$
Proposition.
Let $U$ be measurable.
If $f:U\to\overline R$ is measurable or unsigned measurable, then $f^{-1}(I)$ is measurable for every interval $I$,
where each endpoint may be open or closed, finite or infinite.
(show proof)
Proof.
We will first show that, given a simple or unsigned simple function $g:U\to\overline R$ and $c\in\overline R$,
$\{x\in U|g(x)\gt c\}$ is measurable.
Assume $g$ is a simple function first.
Note that for some $l\in N^+$,
for some measurable partition $A_1,\ldots,A_l$ of $U$ and $c_1,\ldots,c_l\in R$,
we have $g=\sum_{k=1}^lc_k\mathcal X_{A_k}$.
Then $$\{x\in U|g(x)\gt c\}=\bigcup\{A_k:c_k\gt c\}$$
which is measurable.
The case when $g$ is unsigned simple is proven similarly.
Let $I=(c,\infty]$ for some $c\in R$.
Then we have $f^{-1}(I)=\{x\in U|f(x)\gt c\}$.
Suppose $f$ is measurable, then there exists a sequence $(f_n)$ of simple functions on $U$, such that
for all $x\in U$, we have $$f(x)=\lim_{n\to\infty}f_n(x)=\inf_{n\in N}\sup_{m\ge n}f_m(x)$$
Let $x\in U$.
Suppose $x\in\{x\in U|f(x)\gt c\}$, then for some $k\in N^+$, $\inf_{n\in N}\sup_{m\ge n}f_m(x)=f(x)\gt c+\frac{1}{k}$,
thus $\sup_{m\ge n}f_m(x)\gt c+\frac{1}{k}$ for all $n\in N$.
Now suppose for some $k\in N^+$, for all $n\in N$, $\sup_{m\ge n}f_m(x)\gt c+\frac{1}{k}$.
Then $f(x)=\inf_{n\in N}\sup_{m\ge n}f_m(x)\ge c+\frac{1}{k}\gt c$. Thus $x\in\{x\in U|f(x)\gt c\}$.
Therefore,
$$\{x\in U|f(x)\gt c\}=\bigcup_{k\in N^+}\bigcap_{n\in N}\{x\in U|\sup_{m\ge n}f_m(x)\gt c+\frac{1}{k}\}
=\bigcup_{k\in N^+}\bigcap_{n\in N}\bigcup_{m\ge n}\{x\in U|f_m(x)\gt c+\frac{1}{k}\}$$
Note that each $\{x\in U|f_m(x)\gt c+\frac{1}{k}\}$ is measurable.
Hence $f^{-1}(I)$ is measurable.
The case when $f$ is unsigned measurable is proven similarly.
The following applies to both measurable and unsigned measurable $f$.
Since $f^{-1}((-\infty,\infty])=\bigcup_{k\in N}f^{-1}((-k,\infty])$ and $f^{-1}((\infty,\infty])=\emptyset$ are measurable,
given any $c\in\overline R$, $f^{-1}((c,\infty])$ is measurable, thus $f^{-1}([-\infty,c])=U\setminus f^{-1}((c,\infty])$ is measurable.
Let $c\in R$, then $f^{-1}([c,\infty])=\bigcap_{k\in N^+}f^{-1}((c-\frac{1}{k},\infty])$ is measurable.
Since $f^{-1}([-\infty,\infty])=U$ and $f^{-1}([\infty,\infty])=\bigcap_{k\in N}f^{-1}((k,\infty])$ are measurable,
given any $c\in\overline R$, $f^{-1}([c,\infty])$ is measurable, thus $f^{-1}([-\infty,c))=U\setminus f^{-1}([c,\infty])$ is measurable.
From here, we can see that given any interval $I$, $f^{-1}(I)$ is measurable.
$\blacksquare$
Proposition.
Let $U$ be measurable and $f:U\to\overline R$.
If any of the following conditions is satisfied:
- for all $c\in R$, $\{x\in U|f(x)\gt c\}$ is measurable;
- for all $c\in R$, $\{x\in U|f(x)\ge c\}$ is measurable;
- for all $c\in R$, $\{x\in U|f(x)\lt c\}$ is measurable;
- for all $c\in R$, $\{x\in U|f(x)\le c\}$ is measurable;
then $f$ is measurable.
(show proof)
Proof.
If any of these conditions is satisfied, then $f^{-1}(I)$ is measurable for any interval $I$.
Note that $U$ can be partitioned into $U_{-\infty}=f^{-1}(-\infty)$, $U_R=f^{-1}(R)$, and $U_\infty=f^{-1}(\infty)$,
and they are all measurable.
Let $n\in N^+$.
Then for $k\in\{1,\ldots,n^2\}$, $S^+_k=f^{-1}((\frac{k}{n},\infty))$ and $S^-_k=f^{-1}((-\infty,-\frac{k}{n}))$ are measurable subsets of $U_R$.
Thus $$f_n=\sum_{k=1}^{n^2}\p{\frac{1}{n}\mathcal X_{S^+_k}+\frac{-1}{n}\mathcal X_{S^-_k}}$$ defined on $U_R$ is a simple function.
Note that given $x\in U_R$, for all natural $n\gt\abs{f(x)}$, $\abs{f(x)-f_n(x)}\le\frac{1}{n}$.
We have shown that $(f_n)$ converges pointwise to $f$ on $U_R$.
Note that let $g_n=n\mathcal X_{U_\infty}$ and $h_n=-n\mathcal X_{U_{-\infty}}$ for $n\in N^+$,
then $(g_n)$ converges pointwise to $f$ on $U_\infty$ and $(h_n)$ converges pointwise to $f$ on $U_{-\infty}$.
If we extend the domains of $(f_n),(g_n),(h_n)$ to $U$, then $(f_n+g_n+h_n)$ converges pointwise to $f$ on $U$,
thus $f$ is measurable.
Suppose $f$ is signed. Note that for each $n\in N^+$, for some $l_n\in N^+$,
for some measurable partition $A_{n,1},\ldots,A_{n,l_n}$ of $U$ and $c_{n,1},\ldots,c_{n,l_n}\in R$,
we have $f_n+g_n+h_n=\sum_{k=1}^{l_n}c_{n,k}\mathcal X_{A_{n,k}}$.
Let $f^*_n=\sum_{k=1}^{l_n}\sup\{c_{n,k},0\}\mathcal X_{A_{n,k}}$, then $f^*_n$ is an unsigned simple function on $U$.
Let $x\in U_R$ and $\varepsilon\gt0$.
Then there exists $m\in N^+$ such that for all $n\ge m$, $d((f_n+g_n+h_n)(x),f(x))\lt\varepsilon$.
Note that for some $j_n\in\{1,\ldots,l_n\}$, $x\in A_{n,j_n}$.
Thus $$d(f^*_n(x),f(x))=d(\sup\{c_{n,j_n},0\},f(x))\le d(c_{n,j_n},f(x))=d((f_n+g_n+h_n)(x),f(x))\lt\varepsilon$$
Hence $(f^*_n)$ converges pointwise to $f$ on $U_R$.
A similar argument goes for $x\in U_\infty$, and $U_{-\infty}=\emptyset$.
Therefore, $(f^*_n)$ converges pointwise to $f$ on $U$,
implying $f$ is unsigned measurable.
$\blacksquare$
Proposition.
Let $U$ be measurable, let $f$ and $g$ be measurable functions on $U$, and let $c\in R$,
then $f+g$, if well-defined, and $cf$ are measurable.
(show proof)
Proof.
There exist a sequence $(f_n)$ of simple functions that converges pointwise to $f$ and
a sequence $(g_n)$ of simple functions that converges pointwise to $g$.
Suppose $f+g$ is well-defined, then $(f_n+g_n)_n$ is a sequence of simple functions that converges pointwise to $f+g$, implying $f+g$ is measurable.
If $c\in R$, then $(cf_n)_n$ is a sequence of simple functions that converges pointwise to $cf$, implying $cf$ is measurable.
$\blacksquare$
Proposition.
Let $(f_n)$ be extended-real-valued functions on measurable $U$ that converge pointwise to an extended-real-valued function $f$.
If each $f_n$ is measurable, then $f$ is measurable.
(show proof)
Proof.
Suppose $(f_n)$ are extended-real-valued measurable functions on measurable $U$.
Let $c\in R$.
Then $\{x\in U|f_n(x)\gt c\}$ is measurable for each $n$.
Thus $\{x\in U|\sup_n f_n(x)\gt c\}=\bigcup_n\{x\in U|f_n(x)\gt c\}$ is measurable.
Thus $\sup_nf_n$ is measurable.
By a symmetric argument, $\inf_nf_n$ is also measurable.
Now suppose $(f_n)$ are extended-real-valued measurable functions on measurable $U$ that converge pointwise to an extended-real-valued function $f$.
For each $n$, $\sup_{m\ge n}f_m$ is measurable, and $\inf_n\sup_{m\ge n}f_m$ is measurable.
Thus $f=\inf_n\sup_{m\ge n}f_m$ is also measurable.
$\blacksquare$
Proposition.
Let $(f_n)$ be extended-real-valued functions on measurable $U$ such that $\sum f_n$ converges pointwise to an extended-real-valued function $f$.
If each $f_n$ is measurable, then $f$ is measurable.
(show proof)
Proof.
There is an implicit condition to this: for each $p\in U$,
there do not exist $i,j$ such that $f_i(p)=\infty$ and $f_j(p)=-\infty$, to ensure that
$\sum_{n\le m}f_n(p)$ is defined for any $m$.
Then we can define $(g_m)$ such that $g_m(p)=\sum_{n\le m}f_n(p)$ for all $p\in U$.
Then $(g_m)$ converges pointwise to $f$.
Since each $g_m$ is measurable, $f$ is measurable.
$\blacksquare$
Note.
Let $f$ be an extended-real-valued function defined on measurable $U$.
We use $U_+$ to denote $\{x\in U|f(x)\ge0\}$ and $U_-$ to denote $\{x\in U|f(x)\le0\}$.
It is easy to see that if $f$ is measurable on $U$, then
- $U_+$ and $U_-$ are measurable,
- $f$ is unsigned measurable on $U_+$, and
- $-f$ is unsigned measurable on $U_-$.
Lebesgue integral
Let $(X,\sigma,\mu)$ be a measure space.
Given an unsigned function $f$ on measurable $U$,
let $\mathcal S_U$ denote the set of unsigned simple functions on $U$,
we define $$I_*(f,U)=\sup\{I(g):g\in\mathcal S_U,g\le f\}$$
Note that $0\in\mathcal S_U$ and $0\le f$, thus $I_*(f,U)$ is unsigned.
Let $f$ be a measurable function on measurable $U$.
If not both $I_*(f,U_+)$ and $I_*(-f,U_-)$ are $\infty$,
then $f$ is said to be Lebesgue integrable on $U$ with respect to $(X,\sigma,\mu)$,
and we define the Lebesgue integral of $f$ on $U$ with respect to $(X,\sigma,\mu)$ to be
$$\int_U fd\mu=I_*(f,U_+)-I_*(-f,U_-)$$
Note.
Note that given a measure space $(X,\sigma,\mu)$ and a measurable function $f$ on $U\in\sigma$,
integrability and integral of $f$ with respect to $(X,\sigma,\mu)$ and the restriction $(U,\sigma_U,\mu_U)$ are equivalent.
Note.
Suppose $f$ is unsigned on measurable $U$.
Let $\mathcal S^*_U$ be the set of unsigned simple functions on $U$ with finite coefficients,
then $$\sup\{I(g):g\in\mathcal S^*_U,g\le f\}=\sup\{I(g):g\in\mathcal S_U,g\le f\}$$
(show proof)
Proof.
Clearly, $$\sup\{I(g):g\in\mathcal S^*_U,g\le f\}\le\sup\{I(g):g\in\mathcal S_U,g\le f\}$$
Suppose $r=\sup\{I(g):g\in\mathcal S_U,g\le f\}$ is finite.
Let $\varepsilon\gt0$, then for some $g\in\mathcal S_U$ with $g\le f$, $I(g)\in(r-\varepsilon,r]$.
Note that for some $l\in N^+$,
for some measurable partition $A_1,\ldots,A_l$ of $U$ and $c_1,\ldots,c_l\in[0,\infty]$ with at most one being $\infty$,
we have $g=\sum_{k=1}^lc_k\mathcal X_{A_k}$.
If $c_1,\ldots,c_l$ are finite, then $I(g)\in\{I(g):g\in\mathcal S^*_U,g\le f\}$.
If some $c_j$ is $\infty$, then $\mu(A_j)=0$, because otherwise $r\ge I(g)=\infty$.
Then $I(g)=I(\sum_{k\neq j}c_k\mathcal X_{A_k})\in\{I(g):g\in\mathcal S^*_U,g\le f\}$.
Thus $\sup\{I(g):g\in\mathcal S^*_U,g\le f\}\ge I(g)\gt r-\varepsilon$.
Since $\varepsilon$ is arbitrary, $\sup\{I(g):g\in\mathcal S^*_U,g\le f\}\ge\sup\{I(g):g\in\mathcal S_U,g\le f\}$.
Now suppose $\sup\{I(g):g\in\mathcal S_U,g\le f\}=\infty$.
Let $s\in R$, then for some $g\in\mathcal S_U$ with $g\le f$, $I(g)\in(s,\infty]$.
Again, for some $l\in N^+$,
for some measurable partition $A_1,\ldots,A_l$ of $U$ and $c_1,\ldots,c_l\in[0,\infty]$ with at most one being $\infty$,
we have $g=\sum_{k=1}^lc_k\mathcal X_{A_k}$.
If $c_1,\ldots,c_l$ are finite or if some $c_j$ is $\infty$ with $\mu(A_j)=0$,
then $\sup\{I(g):g\in\mathcal S^*_U,g\le f\}\ge I(g)\gt s$.
If some $c_j$ is $\infty$ with $\mu(A_j)\gt0$, then for some $t\in[0,\infty)$, $t\mu(A_j)\gt s$.
Note that $t\mathcal X_{A_j}\in\mathcal S^*_U$ and $t\mathcal X_{A_j}\le g\le f$.
Thus $\sup\{I(g):g\in\mathcal S^*_U,g\le f\}\ge I(t\mathcal X_{A_j})=t\mu(A_j)\gt s$.
Since $s$ is arbitrary, $\sup\{I(g):g\in\mathcal S^*_U,g\le f\}=\infty=\sup\{I(g):g\in\mathcal S_U,g\le f\}$.
$\blacksquare$
Note.
Note that If $f$ is unsigned measurable on measurable $U$, then it is integrable on $U$, with
$$\int_U fd\mu=I_*(f,U)$$
Note.
Suppose $f$ and $g$ are extended-real-valued functions defined on measurable $U$. Then clearly,
- if $\mu(U)=0$, then $f$ is integrable on $U$ and $$\int_U fd\mu=0$$
- if $f$ is an unsigned simple function on $U$, then $f$ is integrable on $U$ and $$\int_Ufd\mu=I(f)$$
- if $f$ is integrable on $U$, then $-f$ is integrable on $U$ and $$\int_U(-f)d\mu=-\int_U fd\mu$$
- if $f$ is unsigned and integrable on $U$, and $V\subseteq U$ is measurable, then $f$ is integrable on $V$ and $$\int_Vfd\mu\le\int_Ufd\mu$$
- if $f$ and $g$ are unsigned and integrable on $U$, and $f\le g$ on $U$, then $$\int_Ufd\mu\le\int_Ugd\mu$$
Proposition.
Suppose $f$ is an extended-real-valued function defined and unsigned on measurable $(U_i)$.
If $f$ is integrable on each $U_i$, then $f$ is integrable on $\bigcup U_i$ and
$$\int_{\bigcup U_i}fd\mu\le\sum\int_{U_i}fd\mu$$
If in addition, $\mu(U_i\cap U_j)=0$ whenever $i\neq j$, then
$$\int_{\bigcup U_i}fd\mu=\sum\int_{U_i}fd\mu$$
(show proof)
Proof.
Define $U^*_i=U_i\setminus\p{\bigcup_{j\lt i}U_j}$,
then $(U^*_i)$ is a measurable partition of $\bigcup U_i$.
Since $f$ is integrable on each $U_i$, it is measurable on each $U_i$, and thus each $U^*_i$.
Suppose $(f_{j})_i$ are simple functions that converge pointwise to $f$ on $U^*_i$.
Define $f^*_j=\sum_{i\le j}f_{j,i}$ such that each indicator function is extended to $\bigcup U^*_i$,
then $(f^*_j)$ are simple functions that converge pointwise to $f$ on $\bigcup U^*_i$.
Thus $f$ is measurable on $\bigcup U^*_i$. Since $f$ is also unsigned on $\bigcup U^*_i$, it is integrable on $\bigcup U^*_i$ (which is also $\bigcup U_i$).
If some $\int_{U_i}fd\mu$ is $\infty$, then both sides are $\infty$.
Now suppose all $\int_{U_i}fd\mu$ are finite, then all $\int_{U^*_i}fd\mu$ are also finite.
Let $J$ be any finite set of indexes.
Suppose $\abs{J}\gt0$. Let $\varepsilon\gt0$, then
for all $j\in J$, there exists an unsigned simple function $f_j$ on $U^*_j$ bounded above by $f$
such that $I_*(f,U^*_j)-I(f_j)\lt\frac{\varepsilon}{\abs{J}}$.
Then $\sum_{j\in J}f_j$ is an unsigned simple function on $\bigcup U^*_i$ bounded above by $f$. Thus
$$I_*\p{f,\bigcup U^*_i}\ge I\p{\sum_{j\in J}f_j}=\sum_{j\in J}I(f_j)\gt\sum_{j\in J}I_*(f,U^*_j)-\varepsilon$$
Since $\varepsilon$ is arbitrary, we have $$\int_{\bigcup U^*_i}fd\mu=I_*\p{f,\bigcup U^*_i}\ge\sum_{j\in J}I_*(f,U^*_j)=\sum_{j\in J}\int_{U^*_j}fd\mu$$
Trivially, if $\abs{J}=0$, this still holds.
Hence $$\int_{\bigcup U^*_i}fd\mu\ge\sup\brace{\sum_{j\in J}\int_{U^*_j}fd\mu:J\subseteq N,\abs{J}\in N}=\sum\int_{U^*_i}fd\mu$$
Suppose $\int_{\bigcup U^*_i}fd\mu$ is finite.
Again, let $\varepsilon\gt0$. Then there exists an unsigned simple function $f^*$ on $\bigcup U^*_i$ bounded above by $f$
such that $I_*\p{f,\bigcup U^*_i}-I(f^*)\lt\varepsilon$.
Note that $f^*|_{U^*_i}$ is an unsigned simple function on $U^*_i$ bounded above by $f$ for each $i$, and
$$\sum I_*(f,U^*_i)\ge\sum I(f^*|_{U^*_i})=I(f^*)\gt I_*\p{f,\bigcup U^*_i}-\varepsilon$$
Since $\varepsilon$ is arbitrary, $$\sum\int_{U^*_i}fd\mu=\sum I_*(f,U^*_i)\ge I_*\p{f,\bigcup U^*_i}=\int_{\bigcup U^*_i}fd\mu$$
By a similar argument, if $\int_{\bigcup U^*_i}fd\mu=\infty$, this still holds.
We have shown that $$\int_{\bigcup U_i}fd\mu=\int_{\bigcup U^*_i}fd\mu=\sum\int_{U^*_i}fd\mu\le\sum\int_{U_i}fd\mu$$
Note that if in addition, $(U_i)$ is pairwise disjoint, then $U^*_i=U_i$ for each $i$.
Then we have $$\int_{\bigcup U_i}fd\mu=\sum\int_{U_i}fd\mu$$
This can be easily reduced to the finite case.
Now suppose $\mu(U_i\cap U_j)=0$ whenever $i\neq j$.
Note that $U_i\cap\p{\bigcup_{j\lt i}U_j}$ and $U^*_i$ is a measurable partition of $U_i$ for each $i$.
Thus $f$ is integrable on both $U_i\cap\p{\bigcup_{j\lt i}U_j}$ and $U^*_i$.
Also note that $$\mu\p{U_i\cap\p{\bigcup_{j\lt i}U_j}}=\mu\p{\bigcup_{j\lt i}\p{U_i\cap U_j}}\le\sum_{j\lt i}\mu(U_i\cap U_j)=0$$
Therefore $\int_{U_i}fd\mu=\int_{U^*_i}fd\mu$ for each $i$,
and we have $$\int_{\bigcup U_i}fd\mu=\int_{\bigcup U^*_i}fd\mu=\sum\int_{U^*_i}fd\mu=\sum\int_{U_i}fd\mu$$
$\blacksquare$
Lemma.
Suppose $f$ is measurable on measurable $U$, then $\abs{f}$ is integrable on $U$, and
$$\int_U\abs{f}d\mu=I_*(f,U_+)+I_*(-f,U_-)$$
(show proof)
Proof.
Suppose $(f_n)$ are simple functions that converge to $f$ on $U$,
then $(\abs{f_n})$ are simple functions that converge to $\abs{f}$ on $U$.
Thus $\abs{f}$ is measurable on $U$.
Since $\abs{f}$ is also unsigned, it is integrable on $U$, and
$$\int_U\abs{f}d\mu=I_*(\abs{f},U)$$
Suppose $I_*(f,U_+),I_*(-f,U_-),I_*(\abs{f},U)$ are finite.
Let $\varepsilon\gt0$. Then there exists an unsigned simple function $g\le\abs{f}$ on $U$ such that $I(g)\gt I_*(\abs{f},U)-\varepsilon$.
Note that $g|_{U_+}$ and $g|_{U_-}$ are unsigned simple functions such that $g|_{U_+}\le\abs{f}|_{U_+}=f|_{U_+}$, $g|_{U_-}\le\abs{f}|_{U_-}=(-f)|_{U_-}$,
and $I(g)=I(g|_{U_+})+I(g|_{U_-})$.
Thus $$I_*(f,U_+)+I_*(-f,U_-)\ge I(g|_{U_+})+I(g|_{U_-})=I(g)\gt I_*(\abs{f},U)-\varepsilon$$
Since $\varepsilon$ is arbitrary, we have $$I_*(f,U_+)+I_*(-f,U_-)\ge I_*(\abs{f},U)$$
Let $\varepsilon\gt0$. Then there exists an unsigned simple function $g_+\le f$ on $U_+$
and an unsigned simple function $g_-\le -f$ on $U_-$ such that $I(g_+)+I(g_-)\gt I_*(f,U_+)+I_*(-f,U_-)-\varepsilon$.
Let $g$ be the sum of $g_+$ and $g_-$ extended to $U$ with zeros,
then $g$ is an unsigned simple function on $U$ such that $g\le\abs{f}$ and $I(g)=I(g_+)+I(g_-)$.
Thus $$I_*(\abs{f},U)\ge I(g)=I(g_+)+I(g_-)\gt I_*(f,U_+)+I_*(-f,U_-)-\varepsilon$$
Since $\varepsilon$ is arbitrary, we have $$I_*(\abs{f},U)\ge I_*(f,U_+)+I_*(-f,U_-)$$
Hence $$\int_U\abs{f}d\mu=I_*(\abs{f},U)=I_*(f,U_+)+I_*(-f,U_-)$$
The general case where $I_*(f,U_+),I_*(-f,U_-),I_*(\abs{f},U)$ may be infinite can be proven similarly.
$\blacksquare$
Proposition.
Suppose $f$ is integrable on measurable $U$, then
$$\int_U\abs{f}d\mu\ge\abs{\int_Ufd\mu}$$
(show proof)
Proof.
By integrability, $f$ is measurable on $U$, thus $\abs{f}$ is integrable on $U$, and
$$\int_U\abs{f}d\mu=I_*(f,U_+)+I_*(-f,U_-)\ge\abs{I_*(f,U_+)-I_*(-f,U_-)}=\abs{\int_Ufd\mu}$$
$\blacksquare$
Proposition.
Suppose $(f_n)$ is an increasing sequence of unsigned measurable functions on measurable $U$ that converges to an unsigned measurable function $f$, then $$\lim_{n\to\infty}\int_Uf_nd\mu=\int_Ufd\mu$$
(show proof)
Proof.
Since $(f_n)$ is increasing, $\int_Uf_nd\mu$ is also increasing, implying $\lim_{n\to\infty}\int_Uf_nd\mu$ exists.
Since $\int_Uf_nd\mu\le\int_Ufd\mu$ for each $n$, we have $$\lim_{n\to\infty}\int_Uf_nd\mu\le\int_Ufd\mu$$
Let $f^*$ be an unsigned simple function on $U$ with finite coefficients bounded by $f$,
then there exists a measurable partition $A_1,\ldots,A_l$ of $U$ and finite unsigned coefficients $c_1,\ldots,c_l$ such that $f^*=\sum_{k=1}^lc_k\mathcal X_{A_k}$.
Let $\varepsilon\in(0,1)$. Define $A_{k,n}=\{p\in A_k|f_n(p)\gt(1-\varepsilon)c_k\}$ for each $k\le l$ and each $n$, then each $A_{k,n}$ is measurable, $A_{k,n}\subseteq A_{k,n+1}$ for each $n$,
and $\bigcup_nA_{k,n}=A_k$.
Then
$$
\lim_{n\to\infty}\int_Uf_nd\mu
=\lim_{n\to\infty}\sum_k\int_{A_k}f_nd\mu
=\sum_k\lim_{n\to\infty}\int_{A_k}f_nd\mu
\ge\sum_k\lim_{n\to\infty}\int_{A_{k,n}}(1-\varepsilon)c_kd\mu
=(1-\varepsilon)\sum_kc_k\lim_{n\to\infty}\mu(A_{k,n})
=(1-\varepsilon)\sum_kc_k\mu\p{\bigcup_nA_{k,n}}
=(1-\varepsilon)\sum_kc_k\mu(A_k)
=(1-\varepsilon)I(f^*)
$$
Since $\varepsilon$ is arbitrary in $(0,1)$, we have $\lim_{n\to\infty}\int_Uf_nd\mu\ge I(f^*)$.
Since $f^*$ is arbitrary in $\{g\in\mathcal S^*_U|g\le f\}$, we have
$$
\lim_{n\to\infty}\int_Uf_nd\mu
\ge\sup\{I(g):g\in\mathcal S^*_U,g\le f\}
=\sup\{I(g):g\in\mathcal S_U,g\le f\}
=\int_Ufd\mu
$$
$\blacksquare$
Note.
Suppose $f$ is unsigned measurable on measurable $U$.
For all $c\in R$, let $U_c=\{x\in U|f(x)\gt c\}$, then $U_c$ is measurable.
For all $n\in N$, let $$f_n=\sum_{k=1}^{n2^n}\frac{1}{2^n}\mathcal X_{U_{\frac{k}{2^n}}}$$
then $(f_n)$ are clearly increasing unsigned simple functions bounded above by $f$ that converge pointwise to $f$.
Lemma.
Suppose $f$ and $g$ are unsigned simple functions defined on measurable $U$. Then
$\int_U(f+g)d\mu=\int_Ufd\mu+\int_Ugd\mu$.
(show proof)
Proof.
Trivially, both $f$ and $g$ are measurable and thus integrable.
Thus $f+g$ is measurable and integrable.
Clearly, $f+g$ is also an unsigned simple function, and
$$\int_U(f+g)d\mu=I(f+g)=I(f)+I(g)=\int_Ufd\mu+\int_Ugd\mu$$
$\blacksquare$
Lemma.
Suppose $f$ and $g$ are extended-real-valued functions defined and unsigned on measurable $U$.
If $f$ and $g$ are integrable on $U$, then $f+g$ is integrable on $U$ and
$\int_U(f+g)d\mu=\int_Ufd\mu+\int_Ugd\mu$.
(show proof)
Proof.
It is obvious that $f+g$ is integrable on $U$.
Let $(f_n)$ be a sequence of increasing unsigned simple functions bounded above by $f$ that converge pointwise to $f$,
and define $(g_n)$ similarly for $g$. Then
$$
\int_U(f+g)d\mu
=\lim_{n\to\infty}\int_U(f_n+g_n)d\mu
=\lim_{n\to\infty}\p{\int_Uf_nd\mu+\int_Ug_nd\mu}
=\lim_{n\to\infty}\int_Uf_nd\mu+\lim_{n\to\infty}\int_Ug_nd\mu
=\int_Ufd\mu+\int_Ugd\mu
$$
$\blacksquare$
Proposition.
Suppose $(f_n)$ are extended-real-valued functions defined and unsigned on measurable $U$.
If each $f_n$ is integrable on $U$, then $\sum f_n$ is integrable on $U$ and
$$\int_U\sum f_nd\mu=\sum\int_Uf_nd\mu$$
(show proof)
Proof.
Since each $f_n$ is unsigned on $U$, $\sum f_n$ is defined and unsigned on $U$.
Since each $f_n$ is integrable on $U$, each $f_n$ is measurable on $U$.
Then $\sum f_n$ is measurable and thus integrable on $U$.
And we have
$$
\sum\int_Uf_nd\mu
=\lim_{m\to\infty}\sum_{n=1}^m\int_Uf_nd\mu
=\lim_{m\to\infty}\int_U\sum_{n=1}^mf_nd\mu
=\int_U\sum f_nd\mu
$$
$\blacksquare$
Note.
Given $a,b\in\overline R$, $a+b$ is defined if and only if $a$ and $b$ are not both $\infty$ and $-\infty$.
Note.
Suppose $f$ is an extended-real-valued function defined on measurable $U$.
Define $f^+=\sup(f,0)$ and $f^-=\inf(f,0)$.
If $f$ is integrable on $U$, then clearly $f^+$ and $f^-$ are integrable on $U$ with $\int_Uf^+d\mu=I_*(f,U_+)$ and $\int_Uf^-d\mu=-I_*(-f,U_-)$, thus $$\int_Ufd\mu=\int_Uf^+d\mu+\int_Uf^-d\mu$$
Proposition.
Suppose $f$ is an extended-real-valued function defined on measurable $U$ and $V$.
If $f$ is integrable on both $U$ and $V$,
then $$\int_{U\cup V} fd\mu+\int_{U\cap V} fd\mu=\int_{U} fd\mu+\int_{V} fd\mu$$
if the additions are defined.
(show proof)
Proof.
Since $f$ is integrable on $U$ and $V$, $f$ is measurable on $U$ and $V$.
Then $f$ is measurable on $U\cap V$, $U\setminus V$ and $V\setminus U$, which are a measurable partition of $U\cup V$.
Thus $f$ is measurable on $U\cup V$.
Hence both $f^+$ and $f^-$ are measurable on all the sets above.
Suppose both $I_*(f,(U\cap V)_+)$ and $I_*(-f,(U\cap V)_-)$ are $\infty$,
then
$$I_*(f,U_+)=\int_{U} f^+d\mu\ge\int_{U\cap V} f^+d\mu=I_*(f,(U\cap V)_+)=\infty$$
$$I_*(-f,U_-)=-\int_{U}f^-d\mu=\int_{U}-f^-d\mu\ge\int_{U\cap V}-f^-d\mu=I_*(-f,(U\cap V)_-)=\infty$$
implying $f$ is not integrable on $U$, a contradiction.
Therefore, $f$ is integrable on $U\cap V$.
Now suppose both $I_*(f,(U\cup V)_+)$ and $I_*(-f,(U\cup V)_-)$ are $\infty$,
then
$$I_*(f,U_+)+I_*(f,V_+)=\int_{U} f^+d\mu+\int_{V} f^+d\mu\ge\int_{U\cup V} f^+d\mu=I_*(f,(U\cup V)_+)=\infty$$
$$I_*(-f,U_-)+I_*(-f,V_-)=\p{-\int_{U}f^-d\mu}+\p{-\int_{V}f^-d\mu}=\int_{U}-f^-d\mu+\int_{V}-f^-d\mu\ge\int_{U\cup V}-f^-d\mu=I_*(-f,(U\cup V)_-)=\infty$$
Thus $I_*(f,U_+)=\infty$ or $I_*(f,V_+)=\infty$, and $I_*(-f,U_-)=\infty$ or $I_*(-f,V_-)=\infty$.
Then $f$ is not integrable on $U$, or $f$ is not integrable on $V$, or $\int_{U} fd\mu+\int_{V} fd\mu$ is not defined,
a contradiction.
Hence $f$ is integrable on $U\cup V$.
Note that $$\int_{U\cup V} f^+d\mu=\int_{U\cap V} f^+d\mu+\int_{U\setminus V} f^+d\mu+\int_{V\setminus U} f^+d\mu$$
$$\int_{U\cup V} f^+d\mu+\int_{U\cap V} f^+d\mu=\int_{U} f^+d\mu+\int_{V} f^+d\mu$$
And similarly, $$\int_{U\cup V} f^-d\mu+\int_{U\cap V} f^-d\mu=\int_{U} f^-d\mu+\int_{V} f^-d\mu$$
Since $f$ is integrable on both $U$ and $V$, and $\int_{U} fd\mu+\int_{V} fd\mu$ is defined,
if $\int_{U} f^+d\mu=\infty$ or $\int_{V} f^+d\mu=\infty$, then $\int_{U} f^-d\mu\neq-\infty$ and $\int_{V} f^-d\mu\neq-\infty$.
Then $$\int_{U} f^+d\mu+\int_{V} f^+d\mu+\int_{U} f^-d\mu+\int_{V} f^-d\mu$$ is defined.
If both $\int_{U} f^+d\mu$ and $\int_{V} f^+d\mu$ are finite, then this still holds.
Similarly, $$\int_{U\cup V} f^+d\mu+\int_{U\cap V} f^+d\mu+\int_{U\cup V} f^-d\mu+\int_{U\cap V} f^-d\mu$$
is also defined.
Therefore,
$$
\int_{U\cup V} fd\mu+\int_{U\cap V} fd\mu
=\int_{U\cup V} f^+d\mu+\int_{U\cap V} f^+d\mu+\int_{U\cup V} f^-d\mu+\int_{U\cap V} f^-d\mu
=\int_{U} f^+d\mu+\int_{V} f^+d\mu+\int_{U} f^-d\mu+\int_{V} f^-d\mu
=\int_{U} fd\mu+\int_{V} fd\mu
$$
$\blacksquare$
Proposition.
Suppose $f$ and $g$ are extended-real-valued functions defined on measurable $U$.
If both $f$ and $g$ are integrable on $U$, and $f\le g$ on $U$, then
$$\int_Ufd\mu\le\int_Ugd\mu$$
(show proof)
Proof.
Since $f\le g$ on $U$, we have $f^+\le g^+$ and $f^-\le g^-$ on $U$, thus $-f^-\ge-g^-$ on $U$.
Note that $$-\int_Ug^-d\mu=\int_U-g^-d\mu\le\int_U-f^-d\mu=-\int_Uf^-d\mu$$
Hence $$\int_Ufd\mu=\int_Uf^+d\mu+\int_Uf^-d\mu\le\int_Ug^+d\mu+\int_Ug^-d\mu=\int_Ugd\mu$$
$\blacksquare$
Proposition.
Suppose $f$ and $g$ are extended-real-valued functions defined on measurable $U$.
If both $f$ and $g$ are integrable on $U$, and $c\in R$, then
- $$\int_Ucfd\mu=c\int_Ufd\mu$$
- $$\int_U(f\pm g)d\mu=\int_Ufd\mu\pm\int_Ugd\mu$$
if the addition/subtraction on both sides are defined.
(show proof)
Proof.
Since $f$ and $g$ are integrable on $U$,
- $f$ and $g$ are measurable on $U$,
- not both $I_*(f,U_+)$ and $I_*(-f,U_-)$ are $\infty$, and
- not both $I_*(g,U_+)$ and $I_*(-g,U_-)$ are $\infty$.
Now let $U_{c+}=\{x\in U|cf\ge0\}$ and $U_{c-}=\{x\in U|cf\le0\}$.
Suppose $c\in(0,\infty)$. Then $cf$ is clearly measurable on $U$.
It is also clear that $I_*(cf,U_{c+})=cI_*(f,U_+)$ and $I_*(-(cf),U_{c-})=cI_*(-f,U_-)$, and not both are $\infty$.
Therefore $\int_Ucfd\mu=c\int_Ufd\mu$.
Suppose $c=0$, this trivially holds.
Suppose $c\in(-\infty,0)$, we have $$\int_Ucfd\mu=\int_U-\abs{c}fd\mu=-\int_U\abs{c}fd\mu=-\abs{c}\int_Ufd\mu=c\int_Ufd\mu$$
Clearly, if $f+g$ is defined, then $f+g$ is measurable on $U$. Then $(f+g)^+$ and $(f+g)^-$ are measurable and thus integrable on $U$.
Note that, when restricted on $U$, $$(f+g)^++(f+g)^-=f+g=f^++f^-+g^++g^-$$
Let $x\in U$.
- If $(f+g)(x)=\infty$, then $f^-(x),g^-(x),(f+g)^-(x)$ are finite, thus
$$(f+g)^+(x)-f^-(x)-g^-(x)=(f+g)^+(x)+(f+g)^-(x)-(f^-(x)+g^-(x)+(f+g)^-(x))=f^+(x)+f^-(x)+g^+(x)+g^-(x)-(f^-(x)+g^-(x)+(f+g)^-(x))=f^+(x)+g^+(x)-(f+g)^-(x)$$
- If $(f+g)(x)=-\infty$, then $f^+(x),g^+(x),(f+g)^+(x)$ are finite, thus
$$(f+g)^-(x)-f^+(x)-g^+(x)=(f+g)^+(x)+(f+g)^-(x)-(f^+(x)+g^+(x)+(f+g)^+(x))=f^+(x)+f^-(x)+g^+(x)+g^-(x)-(f^+(x)+g^+(x)+(f+g)^+(x))=f^-(x)+g^-(x)-(f+g)^+(x)$$
- If $(f+g)(x)=R$, then all terms are finite, and we have $$(f+g)^+(x)+(-f^-)(x)+(-g^-)(x)=f^+(x)+g^+(x)+(-(f+g)^-)(x)$$
Therefore, when restricted on $U$, $$(f+g)^++(-f^-)+(-g^-)=f^++g^++(-(f+g)^-)$$
Note that each term is unsigned on $U$, thus
$$\int_U(f+g)^+d\mu+\int_U-f^-d\mu+\int_U-g^-d\mu=\int_U\p{(f+g)^++(-f^-)+(-g^-)}d\mu=\int_U\p{f^++g^++(-(f+g)^-)}d\mu=\int_Uf^+d\mu+\int_Ug^+d\mu+\int_U-(f+g)^-d\mu$$
Note that
$$f^++g^+\ge(f+g)^+$$
$$f^-+g^-\le(f+g)^-$$
Suppose $\int_Ufd\mu+\int_Ugd\mu$ is defined.
- If $\int_Ufd\mu=\infty$ or $\int_Ugd\mu=\infty$, then $\int_Uf^+d\mu=\infty$ or $\int_Ug^+d\mu=\infty$, and $\int_U-f^-d\mu$ and $\int_U-g^-d\mu$ are finite, implying $\int_U(f+g)^+d\mu=\infty$.
Since $0\le\int_U-(f+g)^-d\mu\le\int_U-f^-d\mu+\int_U-g^-d\mu$, $\int_U-(f+g)^-d\mu$ is finite,
then $f+g$ is integrable on $U$, and both $\int_U(f+g)d\mu$ and $\int_Ufd\mu+\int_Ugd\mu$ are $\infty$.
- If $\int_Ufd\mu=-\infty$ or $\int_Ugd\mu=-\infty$, then $\int_U-f^-d\mu=\infty$ or $\int_U-g^-d\mu=\infty$, and $\int_Uf^+d\mu$ and $\int_Ug^+d\mu$ are finite, implying $\int_U-(f+g)^-d\mu=\infty$.
Since $0\le\int_U(f+g)^+d\mu\le\int_Uf^+d\mu+\int_Ug^+d\mu$, $\int_U(f+g)^+d\mu$ is finite,
then $f+g$ is integrable on $U$, and both $\int_U(f+g)d\mu$ and $\int_Ufd\mu+\int_Ugd\mu$ are $-\infty$.
- If $\int_Ufd\mu$ and $\int_Ugd\mu$ are finite, then all terms are finite, thus $f+g$ is integrable on $U$, and we have
$$\int_U(f+g)d\mu=\int_U(f+g)^+d\mu-\int_U-(f+g)^-d\mu=\int_Uf^+d\mu+\int_Ug^+d\mu-\int_U-f^-d\mu-\int_U-g^-d\mu=\int_Ufd\mu+\int_Ugd\mu$$
Now suppose $f-g$ and $\int_Ufd\mu-\int_Ugd\mu$ are defined, then we have $$\int_U(f-g)d\mu=\int_U(f+(-g))d\mu=\int_Ufd\mu+\int_U-gd\mu=\int_Ufd\mu-\int_Ugd\mu$$
$\blacksquare$
Proposition.
Suppose $(f_n)$ are real-valued functions defined on measurable $U$ such that $\sum\abs{f_n}$ converges pointwise to a real-valued function on $U$.
If each $f_n$ is integrable on $U$, and either $\int_U\sum\abs{f_n}d\mu$ or $\sum\int_U\abs{f_n}d\mu$ is real,
then $\sum f_n$ is integrable on $U$, $\sum\int_Uf_nd\mu$ converges to a real number, and
$$\int_U\sum f_nd\mu=\sum\int_Uf_nd\mu$$
(show proof)
Proof.
Note that $(\abs{f_n})$ are extended-real-valued functions defined and unsigned on measurable $U$,
and each $\abs{f_n}$ is integrable on $U$.
Thus $\sum\abs{f_n}$ is integrable on $U$ and
$$\int_U\sum\abs{f_n}d\mu=\sum\int_U\abs{f_n}d\mu$$
Hence either is real implies the other is also real.
Since $\sum\int_U\abs{f_n}d\mu$ is real, each $\int_U\abs{f_n}d\mu$ is real.
Note that $$\sum\abs{\int_Uf_nd\mu}\le\sum\int_U\abs{f_n}d\mu$$
implying $\sum\abs{\int_Uf_nd\mu}$ is real and each $\int_Uf_nd\mu$ is real.
Thus $\sum\int_Uf_nd\mu$ converges to a real number.
Since $\sum\abs{f_n}$ converges pointwise to a real-valued function on $U$,
$\sum f_n$ also converges pointwise to a real-valued function on $U$.
Since each $f_n$ is integrable, it is measurable.
Thus $\sum f_n$ is measurable on $U$.
Then $\abs{\sum f_n}$ is integrable on $U$.
Let $p\in U$, then $\abs{\sum f_n(p)}\le\sum\abs{f_n(p)}$.
Thus $$\int_U\abs{\sum f_n}d\mu\le\int_U\sum\abs{f_n}d\mu$$
implying $\int_U\abs{\sum f_n}d\mu$ is real.
Hence $\sum f_n$ is integrable on $U$.
Let $f_n^+$ denote the function $\sup(f_n,0)$ and let $f_n^-$ denote the function $\inf(f_n,0)$.
Then $f_n^+$ and $-f_n^-$ are clearly integrable.
Note that $0\le f_n^+\le\abs{f_n}$ and $0\le f_n^-\le\abs{f_n}$.
Thus
$$
\int_U\sum f_nd\mu
=\int_U\sum(f_n^+-(-f_n^-))d\mu
=\int_U\sum f_n^+d\mu-\int_U\sum -f_n^-d\mu
=\sum\int_U f_n^+d\mu-\sum\int_U -f_n^-d\mu
=\sum\int_U(f_n^+-(-f_n^-))d\mu
=\sum\int_U f_nd\mu
$$
$\blacksquare$
Definition.
Let $X$ be a set. A collection $\mathcal F$ of subsets of $X$ is said to be a
$d$-system on $X$ if
- $X\in\mathcal F$,
- if $A,B\in\mathcal F$ and $B\subseteq A$, then $A\setminus B\in\mathcal F$,
- if $(S_n)$ is an increasing sequence of $\mathcal F$, then $\bigcup S_n\in\mathcal F$.
Note that, given a collection of subsets $\tau$ of $X$, $\mathcal P(X)$ is a $d$-system on $X$ that contains $\tau$,
and the intersection of all $d$-systems on $X$ that contain $\tau$ is the minimum $d$-system on $X$ that contains $\tau$, called the $d$-system generated by $\tau$, denoted $d(\tau)$.
Definition.
Let $X$ be a set. A collection $\mathcal F$ of subsets of $X$ is said to be a $\pi$-system on $X$ if
given any $A,B\in\mathcal F$, $A\cap B\in\mathcal F$.
Lemma.
Let $X$ be a set and let $\tau$ be a $\pi$-system on $X$, then $d(\tau)=\sigma(\tau)$.
(show proof)
Proof.
Since $\sigma(\tau)$ is a $d$-system of $X$, $d(\tau)\subseteq\sigma(\tau)$.
Define $$\mathcal A=\{U\in d(\tau)|\forall S\in\tau, U\cap S\in d(\tau)\}$$
Trivially, $X\in\mathcal A$.
If $A,B\in\mathcal A$ and $B\subseteq A$,
then $A\setminus B\in d(\tau)$ and given $S\in\tau$, $B\cap S\subseteq A\cap S$ and
$(A\setminus B)\cap S=(A\cap S)\setminus(B\cap S)\in d(\tau)$, implying $A\setminus B\in\mathcal A$.
If $(U_n)$ is an increasing sequence of $\mathcal A$,
then $\bigcup U_n\in d(\tau)$ and given $S\in\tau$, for each $n$, $U_n\cap S\in d(\tau)$, and $(U_n\cap S)_n$ is increasing,
thus $(\bigcup U_n)\cap S=\bigcup_n(U_n\cap S)\in d(\tau)$, implying $\bigcup U_n\in\mathcal A$.
We have shown that $\mathcal A$ is a $d$-system on $X$.
Since $\tau$ is a $\pi$-system, $\tau\subseteq\mathcal A$, implying $d(\tau)=\mathcal A$.
Define $$\mathcal B=\{U\in d(\tau)|\forall S\in d(\tau), U\cap S\in d(\tau)\}$$
Trivially, $X\in\mathcal B$.
If $A,B\in\mathcal B$ and $B\subseteq A$,
then $A\setminus B\in d(\tau)$ and given $S\in d(\tau)$, $B\cap S\subseteq A\cap S$ and
$(A\setminus B)\cap S=(A\cap S)\setminus(B\cap S)\in d(\tau)$, implying $A\setminus B\in\mathcal B$.
If $(U_n)$ is an increasing sequence of $\mathcal B$,
then $\bigcup U_n\in d(\tau)$ and given $S\in d(\tau)$, for each $n$, $U_n\cap S\in d(\tau)$, and $(U_n\cap S)_n$ is increasing,
thus $(\bigcup U_n)\cap S=\bigcup_n(U_n\cap S)\in d(\tau)$, implying $\bigcup U_n\in\mathcal B$.
We have shown that $\mathcal B$ is a $d$-system on $X$.
Since $d(\tau)=\mathcal A$, $\tau\subseteq\mathcal B$, implying $d(\tau)=\mathcal B$.
Trivially, $\emptyset=X\setminus X\in d(\tau)$, and
if $S\in d(\tau)$, then $S^c=X\setminus S\in d(\tau)$.
Let $(S_n)$ be any sequence of $d(\tau)$.
Since $d(\tau)=\mathcal B$, given $A,B\in d(\tau)$, $A\cap B\in d(\tau)$.
Thus $\bigcup_{k=1}^nS_k=(\bigcap_{k=1}^n(S_k)^c)^c\in d(\tau)$.
Then $(\bigcup_{k=1}^nS_k)_n$ is an increasing sequence of $d(\tau)$,
and we have $\bigcup S_n=\bigcup_n\bigcup_{k=1}^nS_k\in d(\tau)$.
We have shown that $d(\tau)$ is a $\sigma$-algebra over $X$, thus $\sigma(\tau)\subseteq d(\tau)$.
$\blacksquare$
Lemma.
Let $(X,\sigma)$ be a measurable space. Let $\tau$ be a $\pi$-system on $X$ such that $\sigma=\sigma(\tau)$.
If $\mu$ and $\nu$ are measures on $\sigma$ that agree on $\tau$, and $\mu(X)=\nu(X)\lt\infty$, then $\mu=\nu$.
(show proof)
Proof.
Let $\mathcal F=\{U\in\sigma|\mu(U)=\nu(U)\}$.
Trivially, $X\in\mathcal F$.
If $A,B\in\mathcal F$ and $B\subseteq A$, then $\mu(A\setminus B)=\mu(A)-\mu(B)=\nu(A)-\nu(B)=\nu(A\setminus B)$, implying $A\setminus B\in\mathcal F$.
Let $(S_n)$ be an increasing sequence of $\mathcal F$, then $\mu(\bigcup S_n)=\lim_{n\to\infty}\mu(S_n)=\lim_{n\to\infty}\nu(S_n)=\nu(\bigcup S_n)$, implying $\bigcup S_n\in\mathcal F$.
We have shown that $\mathcal F$ is a $d$-system of $X$ that contains $\tau$, thus $\sigma=\sigma(\tau)=d(\tau)\subseteq\mathcal F$,
implying $\mu=\nu$.
$\blacksquare$
Lemma.
Let $(X,\sigma)$ be a measurable space. Let $\tau$ be a $\pi$-system on $X$ such that $\sigma=\sigma(\tau)$.
If $\mu$ and $\nu$ are measures on $\sigma$ that agree on $\tau$, and there exists an increasing sequence $(S_n)$ of $\tau$ with finite measures with respect to $\mu$ and $\nu$,
such that $\bigcup S_n=X$, then $\mu=\nu$.
(show proof)
Proof.
Let $(S_n)$ be the sequence of $\tau$ as described.
For each $n$, we can define measures $\mu_n$ and $\nu_n$ on $\sigma$ by $\mu_n(U)=\mu(U\cap S_n)$ and $\nu_n(U)=\nu(U\cap S_n)$.
Note that, since $\tau$ is a $\pi$-system, given $U\in\tau$, $U\cap S_n\in\tau$, thus $\mu_n(U)=\nu_n(U)$.
Also note that $\mu_n(X)=\mu(S_n)=\nu(S_n)=\nu_n(X)\lt\infty$.
Thus $\mu_n=\nu_n$.
Let $U\in\sigma$, then $$\mu(U)=\mu\p{\p{\bigcup S_n}\cap U}=\mu\p{\bigcup(S_n\cap U)}=\lim_{n\to\infty}\mu(S_n\cap U)=\lim_{n\to\infty}\mu_n(U)=\lim_{n\to\infty}\nu_n(U)=\nu(U)$$
$\blacksquare$
Definition.
Let $(A,\sigma),(B,\tau)$ be measurable spaces.
Let $\mathcal R$ denote $\{S\times T:S\in\sigma,T\in\tau\}$,
then $\mathcal R$ is a collection of subsets of $A\times B$.
Thus $\mathcal R$ generates a $\sigma$-algebra on $A\times B$, denoted $\sigma\otimes\tau$
and called the product $\sigma$-algebra of $\sigma$ and $\tau$.
Notation.
Let $S\subseteq A\times B$, we define
$$S_x=\{y\in Y|(x,y)\in S\}$$
$$S^y=\{x\in X|(x,y)\in S\}$$
Let $f:X\times Y\to\overline R$, we define
$$f_x(y)=f(x,y)$$
$$f^y(x)=f(x,y)$$
Lemma.
Let $(X,\sigma)$ and $(Y,\tau)$ be measurable spaces.
- Given $S\in\sigma\otimes\tau$, for all $x\in X$, $S_x\in\tau$, and for all $y\in Y$, $S^y\in\sigma$.
- Given $(\sigma\otimes\tau)$-measurable $f:X\times Y\to\overline R$, for all $x\in X$, $f_x$ is $\tau$-measurable, and for all $y\in Y$, $f^y$ is $\sigma$-measurable.
(show proof)
Proof.
Let $x\in X$. Define $$\mathcal F_x=\{S\subseteq X\times Y|S_x\in\tau\}$$
Then trivially, $X\times Y\in\mathcal F_x$.
Let $S\in\mathcal F_x$, then $S_x\in\tau$, thus $(S^c)_x=(S_x)^c\in\tau$, implying $S^c\in\mathcal F_x$.
Let $(S_n)$ be a sequence of $\mathcal F_x$, then for each $n$, $(S_n)_x\in\tau$,
thus $(\bigcup S_n)_x=\bigcup_n((S_n)_x)\in\tau$, implying $\bigcup S_n\in\mathcal F_x$.
We have shown that $\mathcal F_x$ is a $\sigma$-algebra on $X\times Y$.
Note that $\{S\times T:S\in\sigma,T\in\tau\}\subseteq\mathcal F_x$, thus $\sigma\otimes\tau\subseteq\mathcal F_x$.
Hence given $S\in\sigma\otimes\tau$, for all $x\in X$, $S_x\in\tau$.
Similarly, given $S\in\sigma\otimes\tau$, for all $y\in Y$, $S^y\in\sigma$.
Let $f:X\times Y\to\overline R$ be $(\sigma\otimes\tau)$-measurable, then for all $c\in R$, $\{(x,y)\in X\times Y|f(x,y)\gt c\}\in\sigma\otimes\tau$.
Let $c\in R$ and $x\in X$, then $\{y\in Y|f_x(y)\gt c\}=\{(x,y)\in X\times Y|f(x,y)\gt c\}_x\in\tau$.
Thus for all $x\in X$, $f_x$ is $\tau$-measurable.
Similarly, for all $y\in Y$, $f^y$ is $\sigma$-measurable.
$\blacksquare$
Notation.
Let $(X,\sigma,\mu)$ and $(Y,\tau,\nu)$ be measure spaces and $S\in\sigma\otimes\tau$.
Then for all $x\in X$, $S_x\in\tau$, thus $\nu(S_x)\in[0,\infty]$, hence we can define a function $f:X\to[0,\infty]$ such that $f(x)=\nu(S_x)$.
Similarly, we can define a function $g:Y\to[0,\infty]$ such that $g(y)=\mu(S^y)$.
We will denote $f$ as $\nu[S]$ and $g$ as $\mu[S]$.
Lemma.
Let $(X,\sigma,\mu)$ and $(Y,\tau,\nu)$ be $\sigma$-finite measure spaces.
If $S\in\sigma\otimes\tau$, then $\nu[S]$ is $\sigma$-measurable and $\mu[S]$ is $\tau$-measurable.
(show proof)
Proof.
Suppose $\nu$ is finite.
Define $\mathcal F$ to be the collection of sets $U$ in $\sigma\otimes\tau$ such that $\nu[U]$ is $\sigma$-measurable.
Trivially, $X\times Y\in\mathcal F$.
Let $A,B\in\mathcal F$ with $B\subseteq A$.
Then $$\nu[A\setminus B](x)=\nu((A\setminus B)_x)=\nu((A_x)\setminus(B_x))=\nu(A_x)-\nu(B_x)=\nu[A](x)-\nu[B](x)$$
Since $\nu[A]$ and $\nu[B]$ are $\sigma$-measurable functions, $\nu[A\setminus B]$ is $\sigma$-measurable, implying $A\setminus B\in\mathcal F$.
Let $(S_n)$ be an increasing sequence of $\mathcal F$.
Then $$\nu\left[\bigcup S_n\right](x)=\nu\p{\p{\bigcup S_n}_x}=\nu\p{\bigcup_n((S_n)_x)}=\lim_{n\to\infty}\nu((S_n)_x)=\lim_{n\to\infty}\nu[S_n](x)$$
Since each $\nu[S_n]$ is $\sigma$-measurable, $\nu[\bigcup S_n]$ is $\sigma$-measurable, implying $\bigcup S_n\in\mathcal F$.
We have shown that $\mathcal F$ is a $d$-system on $X\times Y$.
Note that $\{S\times T:S\in\sigma,T\in\tau\}$ is a $\pi$-system on $X\times Y$,
thus $$\sigma\otimes\tau=\sigma\p{\{S\times T:S\in\sigma,T\in\tau\}}=d\p{\{S\times T:S\in\sigma,T\in\tau\}}\subseteq\mathcal F$$
Hence given $S\in\sigma\otimes\tau$, $\nu[S]$ is $\sigma$-measurable.
Now let $\nu$ be $\sigma$-finite.
Then $Y$ is the union of a sequence $(V_n)$ of disjoint $\tau$-measurable sets with finite $\nu$-measures.
For each $n$, we can define a finite measure $\nu_n$ on $\tau$ by $\nu_n(U)=\nu(U\cap V_n)$.
Then given $S\in\sigma\otimes\tau$, $\nu_n[S]$ is $\sigma$-measurable.
Let $S\in\sigma\otimes\tau$, then
$$\nu[S](x)=\nu(S_x)=\nu\p{\bigcup_n(S_x\cap V_n)}=\sum_n\nu(S_x\cap V_n)=\sum_n\nu_n(S_x)=\sum_n\nu_n[S](x)$$
Since each $\nu_n[S]$ is $\sigma$-measurable, $\nu[S]$ is $\sigma$-measurable.
Similarly, if $S\in\sigma\otimes\tau$, then $\mu[S]$ is $\tau$-measurable.
$\blacksquare$
Proposition.
Let $(X,\sigma,\mu)$ and $(Y,\tau,\nu)$ be $\sigma$-finite measure spaces.
Then there exists a unique measure $\mu\otimes\nu$ on $\sigma\otimes\tau$ such that for all $A\in\sigma$ and $B\in\tau$,
$$(\mu\otimes\nu)(A\times B)=\mu(A)\nu(B)$$
Given $S\in\sigma\otimes\tau$, we have
$$\int_X\nu[S]d\mu=(\mu\otimes\nu)(S)=\int_Y\mu[S]d\nu$$
(show proof)
Proof.
Let $S\in\sigma\otimes\tau$, then $\nu[S]$ is $\sigma$-measurable and $\mu[S]$ is $\tau$-measurable.
Since they are also unsigned, they are integrable, thus $\int_X\nu[S]d\mu$ and $\int_Y\mu[S]d\nu$ are well-defined and unsigned.
Then we can define functions $f:\sigma\otimes\tau\to[0,\infty]$ and $g:\sigma\otimes\tau\to[0,\infty]$
by $f(S)=\int_X\nu[S]d\mu$ and $g(S)=\int_Y\mu[S]d\nu$.
Trivially, $\int_X\nu[\emptyset]d\mu=0$.
Let $(S_n)$ be a disjoint sequence of $\sigma\otimes\tau$,
then given $x\in X$, $((S_n)_x)_n$ is a disjoint sequence of $\tau$,
thus
$$
f\p{\bigcup S_n}
=\int_X\nu\p{\p{\bigcup S_n}_x}d\mu
=\int_X\nu\p{\bigcup_n((S_n)_x)}d\mu
=\int_X\sum_n\nu[S_n]d\mu
=\sum_n\int_X\nu[S_n]d\mu
=\sum_nf(S_n)
$$
We have shown that $f$ is a measure on $\sigma\otimes\tau$.
Similarly, $g$ is a measure on $\sigma\otimes\tau$.
Note that given $A\in\sigma$ and $B\in\tau$,
$$f(A\times B)=\int_X\nu((A\times B)_x)d\mu=\int_X\nu(B)\mathcal X_Ad\mu=I(\nu(B)\mathcal X_A)=\mu(A)\nu(B)$$
Similarly, $g(A\times B)=\mu(A)\nu(B)$.
We have shown the existence of the desired measure on $\sigma\otimes\tau$.
Suppose we have measures $\alpha,\beta$ of $\sigma\otimes\tau$ with the desired property.
Note that $\{S\times T:S\in\sigma,T\in\tau\}$ is a $\pi$-system on $X\times Y$, $\sigma\otimes\tau=\sigma\p{\{S\times T:S\in\sigma,T\in\tau\}}$,
and $\alpha$ and $\beta$ agree on $\{S\times T:S\in\sigma,T\in\tau\}$.
Since $(X,\sigma,\mu)$ and $(Y,\tau,\nu)$ are $\sigma$-finite,
there exist an increasing sequence $(U_n)$ of $\sigma$-measurable sets with finite $\mu$-measure such that $\bigcup U_n=X$,
and an increasing sequence $(V_n)$ of $\tau$-measurable sets with finite $\nu$-measure such that $\bigcup V_n=Y$.
Then $(U_n\times V_n)_n$ is an increasing sequence of $\sigma\otimes\tau$ with finite measures with respect to $\alpha$ and $\beta$,
such that $\bigcup_n(U_n\times V_n)=X\times Y$.
By a lemma above, we have $\alpha=\beta$.
We have shown that the desired measure on $\sigma\otimes\tau$ uniquely exists.
Let $\mu\times\nu$ denote this measure, then $f=\mu\times\nu=g$,
thus given $S\in\sigma\otimes\tau$, we have
$$\int_X\nu[S]d\mu=f(S)=(\mu\otimes\nu)(S)=g(S)=\int_Y\mu[S]d\nu$$
$\blacksquare$
Tonelli's Theorem
Let $(X,\sigma,\mu)$ and $(Y,\tau,\nu)$ be $\sigma$-finite measure spaces.
Let $f$ be an unsigned $(\sigma\otimes\tau)$-measurable function on $X\times Y$.
Then
$$\int_X\int_Yf_xd\nu d\mu=\int_{X\times Y}fd(\mu\otimes\nu)=\int_Y\int_Xf^yd\mu d\nu$$
(show proof)
Proof.
Let $S\in\sigma\otimes\tau$, then for all $x\in X$, $$\int_Y(\mathcal X_S)_xd\nu=\nu[S](x)$$
thus $\int_Y(\mathcal X_S)_xd\nu$ is integrable with respect to $(X,\sigma,\mu)$,
and $$\int_X\int_Y(\mathcal X_S)_xd\nu d\mu=\int_X\nu[S](x)d\mu=(\mu\otimes\nu)(S)=\int_{X\times Y}\mathcal X_Sd(\mu\otimes\nu)$$
Now suppose $f$ is a simple function with non-negative coefficients,
then $f=\sum_{k=1}^nc_k\mathcal X_{S_k}$ for some $n\in N$, $c_1,\ldots,c_n\in[0,\infty)$, and $S_1,\ldots,S_n\in\sigma\otimes\tau$.
For all $x\in X$,
$$
\int_Yf_xd\nu
=\int_Y\p{\sum_{k=1}^nc_k\mathcal X_{S_k}}_xd\nu
=\int_Y\sum_{k=1}^nc_k(\mathcal X_{S_k})_xd\nu
=\sum_{k=1}^nc_k\int_Y(\mathcal X_{S_k})_xd\nu
=\sum_{k=1}^nc_k\nu[S_k](x)
$$
Thus $\int_Yf_xd\nu$ is integrable with respect to $(X,\sigma,\mu)$,
and $$
\int_X\int_Yf_xd\nu d\mu
=\int_X\sum_{k=1}^nc_k\nu[S_k]d\mu
=\sum_{k=1}^nc_k\int_X\nu[S_k]d\mu
=\sum_{k=1}^nc_k\int_{X\times Y}\mathcal X_{S_k}d(\mu\otimes\nu)
=\int_{X\times Y}\sum_{k=1}^nc_k\mathcal X_{S_k}d(\mu\otimes\nu)
=\int_{X\times Y}fd(\mu\otimes\nu)
$$
Now suppose $f$ is unsigned measurable, then there exists an increasing sequence $(f_n)$ of simple functions with non-negative coefficients that converges pointwise to $f$.
For all $x\in X$, $((f_n)_x)_n$ is an increasing sequence of unsigned $\tau$-measurable functions.
Thus
$$
\int_Yf_xd\nu
=\int_Y\p{\lim_{n\to\infty}f_n}_xd\nu
=\int_Y\lim_{n\to\infty}(f_n)_xd\nu
=\lim_{n\to\infty}\int_Y(f_n)_xd\nu
$$
implying $\int_Yf_xd\nu$ is integrable with respect to $(X,\sigma,\mu)$.
Note that $(\int_Y(f_n)_xd\nu)_n$ is an increasing sequence of unsigned $\sigma$-measurable functions.
Thus
$$
\int_X\int_Yf_xd\nu d\mu
=\lim_{n\to\infty}\int_X\int_Y(f_n)_xd\nu d\mu
=\lim_{n\to\infty}\int_{X\times Y}f_nd(\mu\otimes\nu)
=\int_{X\times Y}fd(\mu\otimes\nu)
$$
The other equality can be proven by a symmetric argument.
$\blacksquare$
Fubini's Theorem
Let $(X,\sigma,\mu)$ and $(Y,\tau,\nu)$ be $\sigma$-finite measure spaces.
Let $f$ be a $(\sigma\otimes\tau)$-measurable function on $X\times Y$ such that $$\int_{X\times Y}\abs{f}d(\mu\otimes\nu)\lt\infty$$
Then
$$\int_X\int^*_Yf_xd\nu d\mu=\int_{X\times Y}fd(\mu\otimes\nu)=\int_Y\int^*_Xf^yd\mu d\nu$$
where $\int^*$ evaluates to $0$ if the function is not integrable.
(show proof)
Proof.
Since $f$ is measurable, $\abs{f}$ is unsigned measurable and thus integrable.
And we have $$\int_X\int_Y\abs{f}_xd\nu d\mu=\int_{X\times Y}\abs{f}d(\mu\otimes\nu)\lt\infty$$
Since $f$ is measurable and $\int_{X\times Y}\abs{f}d(\mu\otimes\nu)\lt\infty$, $f$ is integrable.
Since $f$ is measurable, $f^+$ and $-f^-$ are unsigned measurable,
then $\int_Y(f^+)_xd\nu$ and $\int_Y(-f^-)_xd\nu$, as functions of $x$, are unsigned measurable.
Let $X^*$ denote the set of $x\in X$ such that $f_x$ is not integrable.
Since $(f_x)^+=(f^+)_x$ and $-(f_x)^-=(-f^-)_x$, we have
$$X^*=\left\{x\in X\bigg|\int_Y(f^+)_xd\nu=\int_Y(-f^-)_xd\nu=\infty\right\}$$
thus $X^*$ is measurable.
Since $\int_X\int_Y\abs{f}_xd\nu d\mu\lt\infty$, $X^*$ has zero measure.
Define $g:X\times Y\to\overline R$ by $g(x,y)=0$ if $x\in X^*$ and $g(x,y)=f(x,y)$ otherwise, then
$$
\int_{X\times Y}fd(\mu\otimes\nu)
=\int_{X\times Y}f^+d(\mu\otimes\nu)-\int_{X\times Y}-f^-d(\mu\otimes\nu)
=\int_X\int_Y(f^+)_xd\nu d\mu-\int_X\int_Y(-f^-)_xd\nu d\mu
=\int_X\int_Y(g^+)_xd\nu d\mu-\int_X\int_Y(-g^-)_xd\nu d\mu
=\int_X\int_Y((g^+)_x-(-g^-)_x)d\nu d\mu
=\int_X\int_Yg_xd\nu d\mu
=\int_X\int^*_Yf_xd\nu d\mu
$$
The other equality can be proven by a symmetric argument.
$\blacksquare$
Lemma.
With respect to Lebesgue measure space,
if $f:U\to R$ is continuous, where $U\subseteq R^d$ is measurable, then $f$ is measurable.
(show proof)
Proof.
Let $A$ be an open subset of $U$, as a topological subspace of $R^d$, then $A$ is open with respect to the standard topology of $U$, as a metric subspace of $R^d$.
Let $\varepsilon\gt0$. Since $U$ is measurable, there exists an open set $B$ of $R^d$ containing $U$ such that $m^*(B\setminus U)\lt\varepsilon$.
Let $p\in A$, then there exists $r\gt0$ such that $\{x\in U|d(x,p)\lt r\}\subseteq A$ and $\{x\in R^d|d(x,p)\lt r\}\subseteq B$.
Note that let $x\in R^d$ with $d(x,p)\lt r$, if $x\notin A$, then $x\notin\{x\in U|d(x,p)\lt r\}$, thus $x\notin U$, and hence $x\in B\setminus U$.
This implies that for all $x\in R^d$ with $d(x,p)\lt r$, either $x\in A$ or $x\in B\setminus U$, thus $$\{x\in R^d|d(x,p)\lt r\}\subseteq A\cup(B\setminus U)$$
And $p$ is thus an interior point of $A\cup(B\setminus U)$.
Since $p$ is arbitrarily chosen from $A$, $A$ is a subset of the interior of $A\cup(B\setminus U)$, which is open in $R^d$.
Denote the interior of $A\cup(B\setminus U)$ by $C$, then
$$m^*(C\setminus A)\le m^*((A\cup(B\setminus U))\setminus A)=m^*(B\setminus U)\lt\varepsilon$$
Therefore $A$ is measurable.
Since $f$ is continuous and real-valued, for all $c\in R$, $f^{-1}((c,\infty])=f^{-1}((c,\infty))$, which is an open subset of $U$, as a topological subspace of $R^d$,
and is thus measurable. Hence $f$ is measurable.
$\blacksquare$
Note.
For the discussion below on Riemann integral, assume $B=\bigtimes_{i=1}^d[a_i,b_i]$ where $a_i,b_i\in R$ and $a_i\le b_i$, and $B\subseteq U\subseteq R^d$.
Also, we will be working under Lebesgue measure space.
Proposition.
Suppose $f:U\to R$, then $f$ is Riemann integrable on $B$ if and only if
$f$ is bounded on $B$ and the set of discontinuity of $f$ on $B$ has measure zero.
(show proof)
Proof.
Define $g:U\to[0,\infty]$ such that
$$g(p)=\inf_{\delta\gt0}\sup_{x,y\in U\cap B_\delta(p)}d(f(x),f(y))$$
Clearly, given $p\in U$, $g(p)\ge0$, and $f$ is continuous at $p$ if and only if $g(p)=0$.
Thus the set of discontinuity of $f$ on $B$ is $G=\{p\in B|g(p)\gt0\}$.
Suppose $f$ is bounded on $B$ and $m^*(G)\gt0$.
Let $G_k=\{x\in B|g(x)\ge\frac{1}{k}\}$ for $k\in N^+$,
then $G=\bigcup_{k\in N^+}G_k$.
Note that $$\sum_{k\in N^+} m^*\p{G_k}\ge m^*\p{\bigcup_{k\in N^+}G_k}=m^*(G)\gt0$$
Thus for some $k\in N^+$, $m^*\p{G_k}\gt0$.
Suppose $P(j_1,\ldots,j_d)$ is a $d$-partition on $B$.
Let $P_1,\ldots,P_l$ denote $P(j_1,\ldots,j_d)$ reindexed into a tuple.
Let $P^*$ denote the union of boundaries of $P_1,\ldots,P_l$, then clearly $m^*(P^*)=0$.
Thus $$m^*\p{G_k}\le m^*\p{G_k\setminus P^*}+m^*(G_k\cap P^*)\le m^*\p{G_k\setminus P^*}+m^*(P^*)=m^*\p{G_k\setminus P^*}\le m^*\p{G_k}$$
implying $m^*\p{G_k\setminus P^*}=m^*\p{G_k}$.
Let $P^*_j$ denote the interior of $P_j$.
Let $J=\{j\in\{1,\ldots,l\}|P^*_j\cap(G_k\setminus P^*)\neq\emptyset\}$.
Then $\{P^*_j:j\in J\}$ covers $G_k\setminus P^*$.
Thus $$U(P,f)-L(P,f)
\ge\sum_{j\in J}\p{\sup_{x\in P_j}f(x)-\inf_{x\in P_j}f(x)}m(P_j)
\ge\sum_{j\in J}\frac{1}{k}m(P_j)
=\frac{1}{k}\sum_{j\in J}m^*(P^*_j)
\ge\frac{1}{k}m^*\p{\bigcup_{j\in J}P^*_j}
\ge\frac{1}{k}m^*\p{G_k\setminus P^*}
=\frac{1}{k}m^*\p{G_k}$$
Note that $\varepsilon=\frac{1}{k}m^*\p{G_k}\gt0$ is independent on the $d$-partition $P$.
Suppose for contradiction that $\inf_PU(P,f)=\sup_PL(P,f)$.
Then for some $d$-partitions $P_1,P_2$, $U(P_1,f)-L(P_2,f)\lt\varepsilon$.
Then their common refinement $P^*$ has $U(P^*,f)-L(P^*,f)\le U(P_1,f)-L(P_2,f)\lt\varepsilon$,
a contradiction.
Hence $f$ is not Riemann integrable on $U$.
Taking the contraposition, if $f$ is Riemann integrable on $U$, then
either $f$ is not bounded on $B$ or $m^*(G)=0$.
But $f$ is bounded on $B$ by definition of Riemann integrability.
Thus $m^*(G)=0$, and hence $m(G)=0$.
Now suppose $f$ is bounded on $B$ and $m(G)=0$.
Note that given $r\gt0$, $G_r=\{p\in B|g(p)\ge r\}$ contains its limit points, and is thus closed.
Since $G_r$ is also bounded, it is compact.
Let $\varepsilon\gt0$ and $$\varepsilon^*=\inf\brace{\frac{\varepsilon}{2\p{\sup_{x\in B}f(x)-\inf_{x\in B}f(x)}+1},\frac{\varepsilon}{2m(B)+1}}$$
Since $m^*(G_{\varepsilon^*})\le m^*(G)=m(G)=0$, there exists a sequence of bounded open boxes $(B_n)$ covering $G_{\varepsilon^*}$ such that $\sum\abs{B_n}\lt\varepsilon^*$.
Since $G_{\varepsilon^*}$ is compact, there exists a finite subcover $B^*_1,\ldots,B^*_k$ of $G_{\varepsilon^*}$.
Let $S_j=\overline{B^*_j}\cap B$ for $j\in\{1,\ldots,k\}$.
Then $S_1,\ldots,S_k$ are closed subboxes of $B$ that cover $G_{\varepsilon^*}$.
Note that $$m\p{\bigcup S_j}\le\sum m(S_j)\le\sum m(B^*_j)\le\sum\abs{B_n}\lt\varepsilon^*$$
Now define a $d$-partition $P$ by the endpoints of the intervals that form each of $S_j$.
Then there are two kinds of partitions in $P$: those that are subboxes of some $S_j$, denoted $P_1$,
and the others, denoted $P_2$.
Note that
- $\bigcup P_1=\bigcup S_j$,
- $\p{\bigcup P_2}\cap\p{\bigcup B^*_j}=\emptyset$ and thus $\bigcup P_2\subseteq\{p\in B|g(p)\lt\varepsilon^*\}$, and
- $\bigcup P_2$ is closed and bounded and thus compact.
Let $p\in\bigcup P_2$, then there exists $\delta_p\gt0$ such that $\sup_{x,y\in U\cap B_{\delta_p}(p)}d(f(x),f(y))\lt\varepsilon^*$.
Note that $\{B_{\delta_p/2}(p):p\in\bigcup P_2\}$ is an open cover of $\bigcup P_2$ and thus has a finite subcover $\{B_{\delta_{p_i}/2}(p_i):i\in\{1,\ldots,l\}\}$ for some $p_1,\ldots,p_l\in\bigcup P_2$.
Let $\delta=\frac{1}{2}\inf\{\delta_{p_i}:i\in\{1,\ldots,l\}\}$.
Note that for all $x,y\in\bigcup P_2$ with $d(x,y)\lt\delta$, $x\in B_{\delta_{p_i}/2}(p_i)$ for some $i\in\{1,\ldots,l\}$.
Then $d(y,p_i)\le d(x,p_i)+d(x,y)\lt\delta_{p_i}$. Thus $x,y\in U\cap B_{\delta_{p_i}}(p_i)$, implying $d(f(x),f(y))\lt\varepsilon^*$.
Now let $P^*$ be a refinement of $P$ such that each partition $A$ in $P^*$ has $d(x,y)\lt\delta$ for all $x,y\in A$.
Let $P^*_1$ denote partitions of $P^*$ in $\bigcup P_1$, and $P^*_2$ denote partitions of $P^*$ in $\bigcup P_2$,
then we have
$$\inf_PU(P,f)-\sup_PL(P,f)
\le U(P^*,f)-L(P^*,f)
=\sum_{A\in P^*_1}\p{\sup_{x\in A}f(x)-\inf_{x\in A}f(x)}m(A)+\sum_{A\in P^*_2}\p{\sup_{x\in A}f(x)-\inf_{x\in A}f(x)}m(A)
$$ $$
\le\sum_{A\in P^*_1}m(A)\p{\sup_{x\in B}f(x)-\inf_{x\in B}f(x)}+\sum_{A\in P^*_2}m(A)\varepsilon^*
=m\p{\bigcup P_1}\p{\sup_{x\in B}f(x)-\inf_{x\in B}f(x)}+m\p{\bigcup P_2}\varepsilon^*
\lt\varepsilon^*\p{\sup_{x\in B}f(x)-\inf_{x\in B}f(x)}+m(B)\varepsilon^*
\lt\frac{\varepsilon}{2}+\frac{\varepsilon}{2}
=\varepsilon$$
Since $\varepsilon$ is arbitrary, $f$ is integrable on $B$.
$\blacksquare$
Riemann integral and Lebesgue integral
Suppose $f:U\to R$ is Riemann integrable on $B$, then $f$ is Lebesgue integrable on $B$,
and the Lebesgue integral of $f$ on $B$ agrees with the Riemann integral of $f$ on $B$.
(show proof)
Proof.
Suppose $f$ is Riemann integrable on $B$.
Then the set $D$ of discontinuity of $f$ on $B$ has measure zero,
thus the set of continuity of $f$ on $B$, which is $C=B\setminus D$, is measurable.
Note that $f|_C$ is a continuous function on a measurable set, and is thus measurable.
Then both $C_+$ and $C_-$ are measurable.
Since $D$ has measure zero, $f|_D$ is measurable, and both $D_+$ and $D_-$ has measure zero.
Hence both $B_+=C_+\cup D_+$ and $B_-=C_-\cup D_-$ are measurable,
and $f|_B$ is measurable. Since $B$ is bounded and $f$ is bounded on $B$, both $I_*(f,B_+)$ and $I_*(-f,B_-)$ are finite,
hence $f$ is Lebesgue integrable on $B$.
Let $\varepsilon\gt0$.
Since $f$ is Riemann integrable on $B=\bigtimes_{i=1}^d[a_i,b_i]$, there exists a $d$-partition $P$ of $B$, reindexed into $P_1,\ldots,P_k$,
such that $U(P,f)-L(P,f)\lt\varepsilon$. For each partition $P_j=\bigtimes_{i=1}^d[a_{i,j},b_{i,j}]$,
let $P^*_j=\bigtimes_{i=1}^d(a_{i,j},b_{i,j}]$.
Define $B^*=\bigtimes_{i=1}^d(a_i,b_i]$, then $P^*_1,\ldots,P^*_k$ is a measurable partition of $B^*$.
If we further define $P_0=P^*_0=B\setminus B^*$, then $P^*_0,\ldots,P^*_k$ is a measurable partition of $B$.
Now define ${P^*_\gt}_j=P^*_j\cap\p{B\setminus B_-}$ and ${P^*_\le}_j=P^*_j\cap B_-$ for $j\in\{0,\ldots,k\}$,
then they are both measurable.
Note that ${P^*_\gt}_0,\ldots,{P^*_\gt}_k$ is a measurable partition of $B\setminus B_-$,
${P^*_\le}_0,\ldots,{P^*_\le}_k$ is a measurable partition of $B_-$,
and $\{{P^*_\gt}_j,{P^*_\le}_j\}$ is a partition of $P^*_j$ for $j\in\{0,\ldots,k\}$.
Define $$L_\gt=\sum_{j=0}^k\sup\{0,\inf_{x\in P_j}f(x)\}\mathcal X_{{P^*_\gt}_j}$$ on $B_+$, then $L_\gt$ is an unsigned simple function on $B_+$ with $L_\gt\le f|_{B_+}$.
Thus $$I_*(f,B_+)\ge I(L_\gt)$$
Define $$L_\le=\sum_{j=0}^k\inf\{0,\inf_{x\in P_j}f(x)\}\mathcal X_{{P^*_\le}_j}$$ on $B_-$, then $-L_\le$ is an unsigned simple function on $B_-$.
Suppose $g$ is an unsigned simple function on $B_-$ such that $g\le(-f)|_{B_-}$,
then for some $l\in N^+$, $c_1,\ldots,c_l\in[0,\infty]$ and measurable partition $A_1,\ldots,A_l$ of $B_-$,
$g=\sum_{j=1}^lc_j\mathcal X_{A_j}$.
Let $S_{i,j}={P^*_\le}_i\cap A_j$ for $i\in\{0,\ldots,k\}$ and $j\in\{1,\ldots,l\}$,
then the double tuple $(S_{i,j})$ is a measurable partition of $B_-$.
Also, given $i$, $S_{i,j}$ is a measurable partition of ${P^*_\le}_i$,
and given $j$, $S_{i,j}$ is a measurable partition of $A_j$.
Thus
$$-L_\le
=\sum_{i=0}^k\p{-\inf\{0,\inf_{x\in P_i}f(x)\}}\mathcal X_{{P^*_\le}_i}
=\sum_{i=0}^k\p{-\inf\{0,\inf_{x\in P_i}f(x)\}}\sum_{j=1}^l\mathcal X_{S_{i,j}}
=\sum_{i=0}^k\sum_{j=1}^l\p{-\inf\{0,\inf_{x\in P_i}f(x)\}}\mathcal X_{S_{i,j}}$$
and
$$g
=\sum_{j=1}^lc_j\mathcal X_{A_j}
=\sum_{j=1}^lc_j\sum_{i=0}^k\mathcal X_{S_{i,j}}
=\sum_{j=1}^l\sum_{i=0}^kc_j\mathcal X_{S_{i,j}}
=\sum_{i=0}^k\sum_{j=1}^lc_j\mathcal X_{S_{i,j}}$$
Note that for all $i\in\{0,\ldots,k\}$ and $j\in\{1,\ldots,l\}$,
either $S_{i,j}=\emptyset$, implying $m(S_{i,j})=0$,
or $S_{i,j}\neq\emptyset$, in which case let $p\in S_{i,j}$, then
$$-\inf\{0,\inf_{x\in P_{i}}f(x)\}\ge-\inf_{x\in P_{i}}f(x)\ge(-f)(p)\ge g(p)=c_j$$
Hence
$$I(-L_\le)
=\sum_{i=0}^k\sum_{j=1}^l\p{-\inf\{0,\inf_{x\in P_i}f(x)\}}m(S_{i,j})
\ge\sum_{i=0}^k\sum_{j=1}^lc_jm(S_{i,j})
=I(g)$$
And we thus have $$I(-L_\le)\ge I_*(-f,B_-)$$
Note that let $J_-=\{j\in\{1,\ldots,k\}|\inf_{x\in P_j}f(x)\le0\}$, then if $j\in\{1,\ldots,k\}\setminus J_-$,
we have $\inf_{x\in P_j}f(x)\gt0$, implying ${P^*_\le}_j=\emptyset$.
Therefore
$$\int_Bf(\vb x)dm
=I_*(f,B_+)-I_*(-f,B_-)
\ge I(L_\gt)-I(-L_\le)
$$ $$
=\p{\sum_{j=0}^k\sup\{0,\inf_{x\in P_j}f(x)\}m({P^*_\gt}_j)}-\p{\sum_{j=0}^k\p{-\inf\{0,\inf_{x\in P_j}f(x)\}}m({P^*_\le}_j)}
=\p{\sum_{j=1}^k\sup\{0,\inf_{x\in P_j}f(x)\}m({P^*_\gt}_j)}+\p{\sum_{j\in J_-}\inf_{x\in P_j}f(x)m({P^*_\le}_j)}
$$ $$
\ge\p{\sum_{j=1}^k\inf_{x\in P_j}f(x)m({P^*_\gt}_j)}+\p{\sum_{j=1}^k\inf_{x\in P_j}f(x)m({P^*_\le}_j)}
=\sum_{j=1}^k\inf_{x\in P_j}f(x)\p{m({P^*_\gt}_j)+m({P^*_\le}_j)}
=\sum_{j=1}^k\inf_{x\in P_j}f(x)m(P^*_j)
=\sum_{j=1}^k\inf_{x\in P_j}f(x)m(P_j)
=L(P,f)$$
By a symmetric argument, $\int_Bf(\vb x)dm\le U(P,f)$.
Let $\int_{[\vb a,\vb b]}f(\vb x)d\vb x$ denote the Riemann integral of $f$ on $B$.
Note that $$L(P,f)\le \int_{[\vb a,\vb b]}f(\vb x)d\vb x\le U(P,f)$$
therefore, $$\abs{\int_Bf(\vb x)dm-\int_{[\vb a,\vb b]}f(\vb x)d\vb x}\le U(P,f)-L(P,f)\lt\varepsilon$$
Since $\varepsilon$ is arbitrary, we have
$$\int_Bf(\vb x)dm=\int_{[\vb a,\vb b]}f(\vb x)d\vb x$$
$\blacksquare$
Notation.
In the context of integrating an integrable function $f$ on measurable $U$ with respect to Lebesgue measure space,
we may denote the integral as $$\int_Uf$$
Proposition.
Let $B=\bigtimes_{j=1}^dI_j$ with $\abs{I_j}\gt0$ for each $j\in\{1,\ldots,d\}$.
Let $f:U\to\overline R$, where $B\subseteq U\subseteq R^d$, be unsigned and measurable on $B$.
Then
$$\int_B f=\lim_{\forall j\in\{1,\ldots,d\},a'_j\to {a_j}^+,b'_j\to {b_j}^-}\int_{B'} f$$
where $a_j,b_j$ are endpoints (in that order) of $I_j$,
$B'=\bigtimes_{j=1}^d[a'_j,b'_j]$, and the limit is defined so that for each $\varepsilon$, one $\delta$ is taken for each variable.
(show proof)
Proof.
Since $\abs{I_j}\gt0$ for each $j\in\{1,\ldots,d\}$, we have $a_j\lt b_j$, and the interior of $B$ is non-empty.
Let $(c_1,\ldots,c_d)$ be a point in the interior of $B$, denoted $B^*$.
For all $n\in N^+$ and $j\in\{1,\ldots,d\}$, define
- $a_{n,j}=a_j+\frac{1}{n}(c_j-a_j)$ if $a_j\in R$;
- $a_{n,j}=c_j-n$ if $a_j=-\infty$;
- $b_{n,j}=b_j-\frac{1}{n}(b_j-c_j)$ if $b_j\in R$;
- $b_{n,j}=c_j+n$ if $b_j=\infty$;
- $B_n=\bigtimes_{j=1}^d(a_{n,j},b_{n,j})$.
- $B^*_n=B_n\setminus\bigcup_{m=1}^{n-1}B_m$.
Then $(B^*_n)$ is a measurable partition of $B^*$.
And we have $$\int_B f=\int_{B^*} f=\sum\int_{B^*_n} f=\sup_k\sum_{n=1}^k\int_{B^*_n} f=\sup_k\int_{\bigcup_{n=1}^kB^*_n} f=\sup_k\int_{B_k} f$$
From here, it is clear that the limit on the right-hand-side converges to $\int_B f$.
$\blacksquare$
Proposition.
Suppose $f:(a,b)\to R$, where $a$ and $b$ are extended-reals with $a\lt b$, is unsigned and continuous.
If $F$ is an antiderivative of $f$, then $$\int_{(a,b)}f=\lim_{s\to b^-}F(s)-\lim_{r\to a^+}F(r)$$
(show proof)
Proof.
Since $f$ is continuous, it is measurable on $(a,b)$ and integrable on any $[r,s]$ where $a\lt r\le s\lt b$.
Since $f$ is also unsigned, it is integrable on $(a,b)$.
Note that given $a\lt r\le s\lt b$,
$$F(s)-F(r)=\int_r^sf(x)dx=\int_{[r,s]}f\ge0$$
thus $F$ is increasing, which implies that
$$\lim_{r\to a^+}F(r)=\inf_rF(r)\neq\infty$$
$$\lim_{s\to b^-}F(r)=\sup_rF(r)\neq-\infty$$
Then $$
\int_{(a,b)}f
=\lim_{r\to a^+,s\to b^-}\int_{[r,s]}f
=\lim_{r\to a^+,s\to b^-}\int_r^sf(x)dx
=\lim_{r\to a^+,s\to b^-}(F(s)-F(r))
=\lim_{r\to a^+,s\to b^-}F(s)-\lim_{r\to a^+,s\to b^-}F(r)
=\lim_{s\to b^-}F(s)-\lim_{r\to a^+}F(r)$$
$\blacksquare$
Note.
We may denote $\int_{(a,b)}f$ as $\int_a^bf(x)dx$, so that notations like $$\int_{-\infty}^\infty f(x)dx$$
makes sense now.
Lemma.
Suppose $U\subseteq R^d$ and $V\subseteq R^{d'}$,
then $m^*(U\times V)\le m^*(U)m^*(V)$.
(show proof)
Proof.
Let $\varepsilon\in(0,1)$ and let $\delta=\frac{\varepsilon}{m^*(U)+m^*(V)+1}$, then $\delta\in(0,1)$.
Suppose $(B_i)$ is a sequence of bounded boxes covering $U$ with $\sum\abs{B_i}\lt m^*(U)+\delta$,
and $(B'_j)$ is a sequence of bounded boxes covering $V$ with $\sum\abs{B'_j}\lt m^*(V)+\delta$.
Note that for each $i$ and $j$, $B^*_{i,j}=B_i\times B'_j$ is a bounded box with $\abs{B^*_{i,j}}=\abs{B_i}\abs{B'_j}$.
Also note that $(B^*_{i,j})$ covers $U\times V$.
Thus
$$
m^*(U\times V)
\le\sum_i\sum_j\abs{B^*_{i,j}}
=\sum_i\sum_j\abs{B_i}\abs{B'_j}
=\sum_i\abs{B_i}\sum_j\abs{B'_j}
\le\sum_i\abs{B_i}(m^*(V)+\delta)
\lt(m^*(U)+\delta)(m^*(V)+\delta)
$$ $$
=m^*(U)m^*(V)+(m^*(U)+m^*(V))\delta+\delta^2
\lt m^*(U)m^*(V)+(m^*(U)+m^*(V)+1)\delta
=m^*(U)m^*(V)+\varepsilon
$$
Since $\varepsilon$ is arbitrary, we have $m^*(U\times V)\le m^*(U)m^*(V)$.
$\blacksquare$
Lemma.
Suppose $U\subseteq R^d$ and $V\subseteq R^{d'}$ are open,
then $m(U\times V)=m(U)m(V)$.
(show proof)
Proof.
Since $U\times V$ is open, it is measurable.
Note that $U$ can be partitioned into a sequence $(B_i)$ of bounded boxes,
and $V$ can be partitioned into a sequence $(B'_j)$ of bounded boxes.
Note that for each $i$ and $j$, $B^*_{i,j}=B_i\times B'_j$ is a bounded box with $\abs{B^*_{i,j}}=\abs{B_i}\abs{B'_j}$.
And $(B^*_{i,j})$ is clearly a partition of $U\times V$.
Thus
$$m(U\times V)
=\sum_i\sum_j\abs{B^*_{i,j}}
=\sum_i\sum_j\abs{B_i}\abs{B'_j}
=\sum_i\abs{B_i}\sum_j\abs{B'_j}
=m(U)m(V)$$
$\blacksquare$
Proposition.
Suppose $U\subseteq R^d$ and $V\subseteq R^{d'}$ are measurable,
then $$m(U\times V)=m(U)m(V)$$
(show proof)
Proof.
Suppose first that $U$ and $V$ are bounded, then both $m(U)$ and $m(V)$ are finite.
Let $\varepsilon\in(0,1)$ and let $\delta=\frac{\varepsilon}{m(U)+m(V)+2}$, then $\delta\in(0,\frac{1}{2})$.
Since $U$ and $V$ are measurable, there exists an open set $S_U$ containing $U$ with $m^*(S_U\setminus U)\lt\delta$,
and an open set $S_V$ containing $V$ with $m^*(S_V\setminus V)\lt\delta$.
Note that $S_U\times S_V$ is open, contains $U\times V$, and
$$(S_U\times S_V)\setminus(U\times V)=((S_U\setminus U)\times S_V)\cup(S_U\times(S_V\setminus V))$$
Thus
$$
m^*((S_U\times S_V)\setminus(U\times V))
\le m^*((S_U\setminus U)\times S_V)+m^*(S_U\times(S_V\setminus V))
\le m(S_U\setminus U)m(S_V)+m(S_U)m(S_V\setminus V)
\le (m(S_U)+m(S_V))\delta
\lt (m(U)+m(V))\delta+2\delta^2
\lt (m(U)+m(V)+1)\delta
\lt\varepsilon
$$
Since $\varepsilon$ is arbitrary, $U\times V$ is measurable.
Again, let $\varepsilon\in(0,1)$ and define $\delta, S_U, S_V$ as above, then
$$
m(U\times V)
=m(S_U\times S_V)-m((S_U\times S_V)\setminus(U\times V))
\gt m(S_U\times S_V)-\varepsilon
=m(S_U)m(S_V)-\varepsilon
\ge m(U)m(V)-\varepsilon
$$
Since $\varepsilon$ is arbitrary, $$m(U\times V)\ge m(U)m(V)$$
Since we also have $$m(U\times V)=m^*(U\times V)\le m^*(U)m^*(V)=m(U)m(V)$$
We conclude that $$m(U\times V)=m(U)m(V)$$
Now for general $U$ and $V$, define
$$B_i=B_i(\vb0_d)\setminus\p{\bigcup_{k\lt i}B_k(\vb0_d)}$$
$$U_i=U\times B_i$$
$$B'_j=B_j(\vb0_{d'})\setminus\p{\bigcup_{k\lt j}B_k(\vb0_{d'})}$$
$$V_j=V\times B'_j$$
Then $(U_i\times V_j)$ is a measurable partition of $U\times V$.
Thus $U\times V$ is measurable and
$$
m(U\times V)
=\sum_i\sum_jm(U_i\times V_j)
=\sum_i\sum_jm(U_i)m(V_j)
=\sum_im(U_i)\sum_jm(V_j)
=m(U)m(V)
$$
$\blacksquare$
Note.
Let $(R^d,\sigma^d,m^d)$ denote the $d$-dimensional Lebesgue measure space.
Then $\sigma^d\otimes\sigma^{d'}\subseteq\sigma^{d+d'}$ (if we implicitly map subsets of $R^d\times R^{d'}$ into subsets of $R^{d+d'}$ by the natural isomorphism).
Note that $(R^{d+d'},\sigma^d\otimes\sigma^{d'},m^{d+d'})$ (with $m^{d+d'}$ implicitly restricted) is a measure space that agrees with $(R^{d+d'},\sigma^d\otimes\sigma^{d'},m^d\otimes m^{d'})$ on
$$m^{d+d'}(U\times V)=m^d(U)m^{d'}(V)=m^d\otimes m^{d'}(U\times V)$$
for all $U\in\sigma^d$ and $V\in\sigma^{d'}$.
Thus $$m^{d+d'}=m^d\otimes m^{d'}$$
Notation.
Suppose $f$ is an extended-real-valued function defined and unsigned on $U\subseteq R^d$.
We denote $$\{(\vb x,y)\in R^{d+1}|\vb x\in U,y\in[0,f(\vb x)]\}$$ by $$U\times f$$
Measure interpretation of integral
Suppose $f$ is an extended-real-valued function defined and unsigned on measurable $U\subseteq R^d$.
If $f$ is integrable on $U$, then
$$m(U\times f)=\int_Uf$$
(show proof)
Proof.
Suppose $U$ is bounded and $f$ is bounded by $1$.
Since $f$ is integrable on $U$, $f$ is measurable on $U$.
Let $U_c=\{x\in U|f(x)\gt c\}$ for $c\in R$, and define
$${f_*}_k=\sum_{j=1}^{k-1}\frac{1}{k}\mathcal X_{U_{\frac{j}{k}}}$$
$${f^*}_k=\sum_{j=0}^{k-1}\frac{1}{k}\mathcal X_{U_{\frac{j}{k}}}$$
on $U$ for each $k\in N^+$.
Note that
- if $x\in U$, $j\lt k$, and $f(x)\in(\frac{j}{k},\frac{j+1}{k}]$, then ${f_*}_k(x)=\frac{j}{k}$ and ${f^*}_k(x)=\frac{j+1}{k}$;
- let ${P_*}_{k,j}=\{x\in U|f(x)\in(\frac{j}{k},\frac{j+1}{k}]\}\times[0,\frac{j}{k}]$ for $j\lt k$,
then these ${P_*}_{k,j}$, together with $\{x\in U|f(x)=0\}\times\{0\}$, is a measurable partition of $U\times {f_*}_k$;
- let ${P^*}_{k,j}=\{x\in U|f(x)\in(\frac{j}{k},\frac{j+1}{k}]\}\times[0,\frac{j+1}{k}]$ for $j\lt k$,
then these ${P^*}_{k,j}$, together with $\{x\in U|f(x)=0\}\times\{0\}$, is a measurable partition of $U\times {f^*}_k$.
Let $U_+$ denote $\{x\in U|f(x)\gt 0\}$, then $U_+\times {f_*}_k=\bigcup_{j\lt k}{P_*}_{k,j}$ and $U_+\times {f^*}_k=\bigcup_{j\lt k}{P^*}_{k,j}$.
Let $\varepsilon\gt0$. Let $n\in N^+$ such that $n\gt\frac{2m(U)}{\varepsilon}$.
Since each ${P^*}_{n,j}$ is measurable, we can find an open superset $S_{n,j}$ for each ${P^*}_{n,j}$
such that $m^*(S_{n,j}\setminus{P^*}_{n,j})\lt\frac{\varepsilon}{2n}$.
Then $S_n=\bigcup_{j\lt n}S_{n,j}$ is open,
$U_+\times {f_*}_n\subseteq U_+\times f\subseteq U_+\times {f^*}_n\subseteq S_n$,
and
$$
m^*(S_n\setminus U_+\times f)
\le m(S_n\setminus U_+\times {f^*}_n)+m(U_+\times {f^*}_n\setminus U_+\times {f_*}_n)
\le m\p{\bigcup_{j\lt n}\p{S_{n,j}\setminus{P^*}_{n,j}}}+m\p{\bigcup_{j\lt n}\p{{P^*}_{n,j}\setminus{P_*}_{n,j}}}
$$ $$
\le \sum_{j\lt n}m(S_{n,j}\setminus{P^*}_{n,j})+\sum_{j\lt n}m({P^*}_{n,j}\setminus{P_*}_{n,j})
\lt \sum_{j\lt n}\frac{\varepsilon}{2n}+\sum_{j\lt n}m\p{\brace{x\in U|f(x)\in\left(\frac{j}{n},\frac{j+1}{n}\right]}}\frac{1}{n}
\le\frac{\varepsilon}{2}+\frac{m(U)}{n}
\lt\varepsilon
$$
Since $\varepsilon$ is arbitrary, $U_+\times f$ is measurable.
Since $\{x\in U|f(x)=0\}\times f$ has measure zero, $U\times f$ is measurable.
Note that, if we let $U_j=\{x\in U|f(x)\in(\frac{j}{k},\frac{j+1}{k}]\}$ for $j\lt k$,
then
$${f_*}_k=\sum_{j\lt k}\frac{j}{k}\mathcal X_{U_j}$$
$${f^*}_k=\sum_{j\lt k}\frac{j+1}{k}\mathcal X_{U_j}$$
Let $U^0$ denote $\{x\in U|f(x)=0\}$, then
$$m(U\times {f_*}_k)=m(U^0\times\{0\})+\sum_{j\lt k}m\p{U_j\times\left[0,\frac{j}{k}\right]}=\sum_{j\lt k}m(U_j)\frac{j}{k}=I({f_*}_k)$$
$$m(U\times {f^*}_k)=m(U^0\times\{0\})+\sum_{j\lt k}m\p{U_j\times\left[0,\frac{j+1}{k}\right]}=\sum_{j\lt k}m(U_j)\frac{j+1}{k}=I({f^*}_k)$$
Note that
$$I({f_*}_k)\le \int_Uf\le I({f^*}_k)$$
$$m(U\times {f_*}_k)\le m(U\times f)\le m(U\times {f^*}_k)$$
Since $$I({f^*}_k)-I({f_*}_k)=\sum_{j\lt k}\frac{1}{k}m(U_j)\le\frac{m(U)}{k}$$
and $k$ is arbitrary, we have $$I({f^*}_k)=I({f_*}_k)$$
Therefore $$\int_Uf=m(U\times f)$$
Now for bounded $U$ and general $f$,
let $f_n=\inf(\sup(f-n,0),1)$ for each $n\in N$.
Then
- $f=\sum f_n$,
- each $f_n$ is unsigned, bounded by $1$ and integrable on $U$, and
- for each $n$, the isometry $I_n:R^{d+1}\to R^{d+1}$ such that $I_n(\vb x,y)=(\vb x,y+n)$
maps $(U\times f_n)\cap(R^d\times(0,1])$ to $(U\times f)\cap(R^d\times(n,n+1])$.
Since $R^d\times\{0\}$ has measure zero,
$$
\int_U f
=\int_U\sum f_n
=\sum\int_Uf_n
=\sum m(U\times f_n)
=\sum m((U\times f_n)\cap(R^d\times(0,1]))
=\sum m((U\times f)\cap(R^d\times(n,n+1]))
=m(U\times f)
$$
Now for general $U$ and general $f$,
let $U_i=U\cap\p{B_i(\vb0_d)\setminus\bigcup_{j\lt i}B_j(\vb0_d)}$ for each $i\in N$.
Then
- each $U_i$ is bounded,
- $(U_i)$ is a measurable partition of $U$, and
- $(U_i\times f)$ is a measurable partition of $U\times f$.
Therefore,
$$
\int_U f
=\sum\int_{U_i}f
=\sum m(U_i\times f)
=m(U\times f)
$$
$\blacksquare$
Lemma.
Let $U$ be an open subset of $R^d$.
Let $f:U\to R^d$ be differentiable.
Let $x,y\in U$ such that $(1-t)x+ty\in U$ for all $t\in[0,1]$.
Let $C\in R$.
If $\norm{DF((1-t)x+ty)}\le C$ for all $t\in[0,1]$, then
$$\sup$$
(show proof)
Proof.
$\blacksquare$