index

Measure space (show)

Definition. We define $$0\times\pm\infty=0\quad\text{and}\quad\pm\infty\times0=0$$

Note. Let $(a_i)$ be a sequence of non-negative extended real numbers. Then $\sum_{i=0}^I a_i$ is either bounded, in which case $\sum a_i=\sup_I\sum_{i=0}^I a_i\in[0,\infty)$, or unbounded, in which case $\sum a_i=\infty$. If $\sup$ is taken on extended real numbers, then we can simply write $$\sum a_i=\sup_I\sum_{i=0}^I a_i$$ which is always a non-negative extended real number.

Note. Let $(a_i)$ and $(b_i)$ be sequences of non-negative extended real numbers, and let $c$ be a non-negative extended real number. Then it is easily verifiable that
  • $$\sum(ca_i)=c\sum a_i$$
  • $$\sum (a_i+b_i)=\sum a_i+\sum b_i$$

Note. Let $(a_i)$ and $(b_i)$ be sequences of non-negative extended real numbers such that $a_i\lt b_i$, we have $\sum a_i\le\sum b_i$, instead of $\sum a_i\lt\sum b_i$, since they can both be $\infty$.

Definition. Let $S$ be an arbitrary set and let $f:S\to[0,\infty]$. We define $$\sum_{s\in S} f(s)=\sup\left\{\sum_{a\in A} f(a):A\subseteq S,\abs{A}\in N\right\}$$ where $\sup$ is taken on extended real numbers.

Note. This definition clearly coincides with our previous definition (given in the "discrete mathematics" section of the "foundation of mathematics" chapter) when $S$ is finite.

Lemma. Let $S$ be a countably infinite set and let $f:S\to[0,\infty]$. Let $g:N\to S$ be a bijection and let $a_i=f(g(i))$, then $$\sum a_i=\sum_{s\in S} f(s)$$ (show proof)

Proposition. Let $S$ be a countably infinite set and let $f:S\to[0,\infty]$. Let $g:N\times N\to S$ be a bijection and let $a_{i,j}=f(g(i,j))$, then $$\sum_i\sum_j a_{i,j}=\sum_{s\in S} f(s)=\sum_j\sum_i a_{i,j}$$ (show proof)

Proposition. Let $(a_{i,j})$ be a double sequence of non-negative extended real numbers, and let $(b_k)$ be a reindex of $(a_{i,j})$, then $$\sum_k b_k=\sum_i\sum_j a_{i,j}$$ (show proof)

Proposition. Let $(a_{i,j})$ be a double sequence of real numbers. If $\sum_i\sum_j\abs{a_{i,j}}$ or $\sum_j\sum_i\abs{a_{i,j}}$ is real, then for any reindex $(b_k)$ of $(a_{i,j})$, $$\sum_i\sum_j a_{i,j}=\sum_k b_k=\sum_j\sum_i a_{i,j}$$ (show proof)

Note. From this point on, by default, indexes of sequences start from $1$.

Note. We will be working under Euclidean spaces with non-zero dimensionality in this section. That is, whenever we write $R^d$, we implicitly assume $d\neq0$.

Interval
An interval is a subset of $R$ that takes one of the following forms: $$(a,b),(a,b],[a,b),[a,b]$$ where $a$ and $b$ are extended real numbers. If $a$ and $b$ are not both $\infty$ or both $-\infty$, we define the length of an interval $I$ with endpoints $a$ and $b$ (in that order) to be $b-a$ if $a\le b$ and $0$ otherwise, denoted $\abs{I}$.

Box
A box in $R^d$ is a Cartesian product of $d$ intervals. We define the volume of $B=I_1\times\ldots\times I_d$ to be $\prod\abs{I_k}$ if well-defined, denoted $\abs{B}$. If all of the intervals have finite endpoints, we call the box bounded. A bounded box always has a non-negative real volume.

Elementary set
An elementary set in $R^d$ is a set that is the union of a finite number of bounded boxes in $R^d$.

Note. Obviously, given elementary sets $A$ and $B$, $A\cup B$, $A\cap B$, $A\setminus B$ are elementary sets.

Proposition. Every elementary set is the union of finite disjoint bounded boxes. (show proof)

Proposition. Given an elementary set, and two finite partitions into bounded boxes, denoted $B_1,\ldots,B_k$ and $B'_1,\ldots,B'_l$, we have $\sum\abs{B_i}=\sum\abs{B'_j}$. (show proof)

Elementary measure
Given an elementary set $E\subset R^d$, we define the elementary measure of $E$, denoted $m(E)$, to be $\sum\abs{B_i}$, where $B_1,\ldots,B_k$ is any finite partition of $E$ into bounded boxes.

Proposition. Suppose $E_1,E_2$ are disjoint elementary sets, then $m(E_1\cup E_2)=m(E_1)+m(E_2)$. (show proof)

Proposition. Suppose $E_1,E_2$ are elementary sets such that $E_1\subseteq E_2$, then $m(E_1)\le m(E_2)$. (show proof)

Proposition. Suppose $E_1,E_2$ are elementary sets, then $m(E_1\cup E_2)\le m(E_1)+m(E_2)$. (show proof)

Jordan measure
Given a bounded set $S\subset R^d$, let $\mathcal E$ denote the collection of elementary sets, we define
  • the Jordan inner measure of $S$ to be $$\sup\{m(E):E\in\mathcal E,E\subseteq S\}$$ denoted $m_{*,J}(S)$;
  • the Jordan outer measure of $S$ to be $$\inf\{m(E):E\in\mathcal E,S\subseteq E\}$$ denoted $m^{*,J}(S)$;
  • the Jordan measure of $S$ to be its Jordan inner/outer measure if they equal, in which case we call $S$ Jordan measurable.

Note. Obviously, every elementary set is Jordan measurable, and its Jordan measure equals its elementary measure. Hence we will denote the Jordan measure of a Jordan measurable set $S$ as $m(S)$.

Lebesgue outer measure
Given a set $S\subseteq R^d$, let $\mathcal B$ denote the collection of bounded boxes, we define the Lebesgue outer measure of $S$ to be $$\inf\brace{\sum\abs{B_i}:(\forall i\in N^+,B_i\in\mathcal B),S\subseteq\bigcup B_i}$$ denoted $m^*(S)$. Note that $m^*(S)\ge0$ (possibly $\infty$) given any $S\subseteq R^d$.

Note. Clearly,
  • $m^*(\emptyset)=0$;
  • given any set $S$, $m^*(S)\le m^{*,J}(S)$;
  • given sets $A\subseteq B$, $m^*(A)\le m^*(B)$;
  • given a bounded set $S$, $m^*(A)\neq\infty$.

Lebesgue measure
Given a set $S\subseteq R^d$, if for every $\varepsilon\gt0$, there exists an open subset $U$ of $R^d$ such that $S\subseteq U$ and $m^*(U\setminus S)\lt\varepsilon$, then $S$ is said to be Lebesgue measurable, and the Lebesgue measure of $S$ is defined to be $m^*(S)$.

Proposition. Every Jordan measurable set is Lebesgue measurable, and its Lebesgue measure equals its Jordan measure. (show proof)

Note. Due to the last proposition, we will denote the Lebesgue measure of a Lebesgue measurable set $S$ as $m(S)$. Note that $m^*$ is an extension of $m$ from Lebesgue measurable sets to $R^d$.

Proposition. $$m^*\p{\bigcup S_i}\le\sum m^*\p{S_i}$$ (show proof)

Note. By the above proposition, we easily find $$m^*\p{A\cup B}\le m^*\p{A}+m^*\p{B}$$

Definition. Two sets are said to be almost disjoint if their interiors are disjoint.

Proposition. Every open set is the union of a sequence of almost disjoint closed bounded boxes. (show proof)

Proposition. Every open set can be partitioned into a sequence of bounded boxes. (show proof)

Lemma. Let $B_1,\ldots,B_n$ be almost disjoint bounded boxes, then $m^*\p{\bigcup B_i}=\sum\abs{B_i}$. (show proof)

Lemma. Let $(B_i)$ be a sequence of almost disjoint bounded boxes, then $m^*\p{\bigcup B_i}=\sum\abs{B_i}$. (show proof)

Lemma. Let $\tau$ denote the collection of open sets, then $$m^*(S)=\inf\brace{m^*(U):U\in\tau,S\subseteq U}$$ (show proof)

Note. By definition, open sets are Lebesgue measurable.

Proposition. Let $(S_i)$ be a sequence of Lebesgue measurable sets, then $\bigcup S_i$ is Lebesgue measurable. (show proof)

Proposition. Closed sets are Lebesgue measurable. (show proof)

Lemma. Sets with Lebesgue outer measure $0$ is measurable. (show proof)

Lemma. The complement of Lebesgue measurable sets are Lebesgue measurable. (show proof)

Definition. In the context of a base set $X$, the intersection of an empty collection of subsets of $X$ is defined to be $X$.

Proposition. Let $(S_i)$ be a sequence of Lebesgue measurable sets, then $\bigcap S_i$ is Lebesgue measurable. (show proof)

Proposition. Let $A$ and $B$ be Lebesgue measurable sets, then $A\setminus B$ is Lebesgue measurable. (show proof)

Isometry
An isometry is a map from $R^d$ to $R^d$ that preserves distance.

Translation
A translation is map from $R^d$ to $R^d$ in the form $T(x)=x+p$, where $p\in R^d$.

Affine transformation
An affine transformation is map from $R^d$ to $R^d$ in the form $A=T\circ L$, where $T$ is a translation and $L$ is a linear transformation. We can also write $A(x)=L(x)+p$ for some $p\in R^d$.

Lemma. Let $I$ be an isometry, then $I'(x)=I(x)-I(\vb0)$ is a linear isometry. (show proof)

Proposition. Every isometry is an affine transformation. (show proof)

Proposition. Every isometry is bijective. (show proof)

Lemma. Let $L:R^d\to R^d$ be an isomorphism and $U\subseteq R^d$. Then $U$ is open if and only if $L(U)$ is open. (show proof)

Lemma. Let $I$ be a linear isometry and $A$ its standard matrix representation, then $\abs{\det(A)}=1$. (show proof)

Lemma. Let $L:R^d\to R^d$ be a linear map and $A$ its standard matrix representation, then $$m^*(L(S))=\abs{\det(A)}m^*(S)$$ (show proof)

Proposition. Let $L:R^d\to R^d$ be a linear map and $A$ its standard matrix representation. If $S$ is a Lebesgue measurable set, then $L(S)$ is Lebesgue measurable, and $$m(L(S))=\abs{\det(A)}m(S)$$ (show proof)

Proposition. Let $I$ be an isometry and $S$ be a Lebesgue measurable set, then $I(S)$ is Lebesgue measurable, and $$m(I(S))=m(S)$$ (show proof)

Proposition. Let $(S_i)$ be a sequence of disjoint Lebesgue measurable sets, then $$m\p{\bigcup S_i}=\sum m(S_i)$$ (show proof)

Measure space
Let $X$ be a set. A collection $\sigma$ of subsets of $X$ such that:
  • $\emptyset\in\sigma$;
  • if $S\in\sigma$ then $S^c\in\sigma$;
  • $\bigcup S_n\in\sigma$ given any sequence $(S_n)$ of $\sigma$;
is called a $\sigma$-algebra over $X$. Such $(X,\sigma)$ is called a measurable space and the elements of $\sigma$ are called measurable sets with respect to $(X,\sigma)$. Given a measurable space $(X,\sigma)$, a function $\mu:\sigma\to[0,\infty]$ such that
  • $\mu(\emptyset)=0$;
  • $\mu\p{\bigcup S_n}=\sum\mu(S_n)$ given any disjoint sequence $(S_n)$ of measurable sets;
is called a measure on $(X,\sigma)$. Such $(X,\sigma,\mu)$ is called a measure space. Clearly, $R^d$, the set of Lebesgue measurable sets, and the Lebesgue measure together form a measure space, called Lebesgue measure space.

Note. Given a measure space $(X,\sigma,\mu)$, it is trivial that:
  • $\bigcup S_n\in\sigma$ given any tuple $(S_n)$ of $\sigma$;
  • $\bigcap S_n\in\sigma$ given any tuple or sequence $(S_n)$ of $\sigma$;
  • $\mu\p{\bigcup S_n}=\sum\mu(S_n)$ given any disjoint tuple $(S_n)$ of measurable sets; in particular, $$\mu(A\cup B)=\mu(A)+\mu(B)$$ given disjoint $A,B\in\sigma$.

Definition. Given a measurable space $(X,\sigma)$ and $U\in\sigma$, define $\sigma_U=\{S\in\sigma|S\subseteq U\}$, then $(U,\sigma_U)$ is a measurable space, called the restriction of $(X,\sigma)$ on $U$. Let $\mu$ be a measure on $\sigma$, define $\mu_U$ to be the restriction of $\mu$ on $\sigma_U$, then $(U,\sigma_U,\mu_U)$ is a measure space, called the restriction of $(X,\sigma,\mu)$ on $U$.

Definition. A measure space $(X,\sigma,\mu)$ is called $\sigma$-finite if $X$ is the union of a sequence $(U_i)$ of measurable sets such that $\mu(U_i)$ is finite for each $i$. Trivially, a Lebesgue measure space is $\sigma$-finite.

Proposition. Let $(X,\sigma,\mu)$ be a measure space. If $A,B\in\sigma$, and $A\setminus B\in\sigma$. (show proof)

Proposition. Let $(X,\sigma,\mu)$ be a measure space. If $A,B\in\sigma$ and $A\subseteq B$, then $\mu(A)\le\mu(B)$. (show proof)

Proposition. Let $(X,\sigma,\mu)$ be a measure space. Then $$\mu\p{\bigcup S_n}\le\sum\mu(S_n)$$ given any sequence $(S_n)$ of measurable sets. (show proof)

Note. It is trivial that the above proposition applies when $(S_n)$ is a tuple of measurable sets. In particular, $$\mu(A\cup B)\le\mu(A)+\mu(B)$$ given $A,B\in\sigma$.

Proposition. Let $(X,\sigma,\mu)$ be a measure space. If $A$ and $B$ are measurable and have finite measure, then $$\mu(A\cup B)=\mu(A)+\mu(B)-\mu(A\cap B)$$ (show proof)

Proposition. Let $(S_i)$ be a sequence of measurable sets such that $\mu(S_j\cap S_k)=0$ whenever $j\neq k$, then $$\mu\p{\bigcup S_i}=\sum\mu(S_i)$$ (show proof)

Proposition. Let $(S_i)$ be a sequence of measurable sets such that $S_i\subseteq S_{i+1}$, then $$\lim_{i\to\infty}\mu(S_i)=\mu\p{\bigcup S_i}$$ (show proof)

Definition. Let $X$ be a set and let $\tau$ be a collection of subsets of $X$, then there exists a minimal $\sigma$-algebra $\sigma$ over $X$ such that $\tau\subseteq\sigma$ (show proof), called the $\sigma$-algebra generated by $\tau$, denoted $\sigma(\tau)$.

Borel $\sigma$-algebra
Let $(X,\tau)$ be a topological space. Then $\sigma(\tau)$ is called the Borel $\sigma$-algebra of $(X,\tau)$. And the elements of $\sigma(\tau)$ are called Borel measurable sets.

Proposition. Borel measurable subsets of $R^d$ are Lebesgue measurable. (show proof)

Lebesgue integral (show)

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)

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)

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)

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)

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)

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)

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)

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)

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)

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)

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)

Proposition. Suppose $f$ is integrable on measurable $U$, then $$\int_U\abs{f}d\mu\ge\abs{\int_Ufd\mu}$$ (show proof)

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)

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)

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)

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)

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)

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)

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)

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)

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)

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)

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)

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)

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)

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)

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)

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)

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)

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)

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)

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)

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)

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)

Lemma. Suppose $U\subseteq R^d$ and $V\subseteq R^{d'}$ are open, then $m(U\times V)=m(U)m(V)$. (show proof)

Proposition. Suppose $U\subseteq R^d$ and $V\subseteq R^{d'}$ are measurable, then $$m(U\times V)=m(U)m(V)$$ (show proof)

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)

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)