Home

Measure Theory

× These are my own notes, so errors and typos are bound to appear!

My work has become much more technical that I am used to, so I thought it would be good to take some notes on basic measure and probability theory in anticipation of working through several theoretical papers. A lot of the definitions below come from Wikipedia, Durrett1, and Axler2, but I’ve added some more intuitive ways of thinking about these concepts that I’ve come up with or collected from some really helpful sites I’ve found.

Note: Not all of the proofs are finished/included. I am hoping to find the time to return to thist post and complete them.


Building Blocks

Measure theory, in my mind, is just about sets, mappings, and ways to describe them. To formalize these ideas, however, we need to define some basic building blocks.

Sets

We’ll begin with a $\sigma$-field For the technical reader, a $\sigma$-field, also called a $\sigma$-algebra, is a generalization of the algebra, which has the same definition except must be closed under finite unions. . This is simply a collection of subsets of some other set.

Definition ($\sigma$-Field).
Let $X$ be a set, and let $\mathcal{P}(X)$ denote its power set (the set of all possible subsets of $X$). A $\sigma$-field is any subset $\mathcal{S} \subseteq \mathcal{P}(X)$ satisfying the following:
  1. $X \in \mathcal{S}$
  2. Closed Under Complementation: $A \in \mathcal{S} \implies A^c \in \mathcal{S}$
  3. Closed Under Countable Unions: $A_1, A_2, \dots \in \mathcal{S} \implies A = \bigcup_{i = 1}^\infty A_i \in \mathcal{S}$
Example ($\sigma$-Field).
Let $X = \{ 1, 2, 3\}$. One $\sigma$-field on $X$ is $\mathcal{S} = \{ \emptyset, \{ 1 \}, \{2, 3 \}, \{1, 2, 3 \} \}$. It is easy to see that the first two properties are satisfied. The entirety of $X$ is in $\mathcal{S}$ by construction. Any element in $\mathcal{S}$, its complement is also in $\mathcal{S}$ (the complement of $\emptyset$ is $X$, and the complement of $\{ 1\}$ is $\{ 2, 3 \}$). Taking any countable union of elements of $\mathcal{S}$ also yields an element of $\mathcal{S}$ (the union of any $A \in \mathcal{S}$ and the empty set is the set itself; the complement of the entirety of $X$ and any $A \in \mathcal{S}$ will just be $X$; and the union of $\{ 1\}$ and $\{ 2, 3\}$ is $X$). This satisfies the third property, which completes the proof that $\mathcal{S}$ is a $\sigma$-field on $X$.

Notice that the first and second properties in the above definition imply that $\emptyset \in \mathcal{S}$ as well. The properties also imply that a $\sigma$-field must be closed under countable intersection. That is, $\cap_{i = 1}^\infty A_i \in \mathcal{S}$ for some sequence of $A_1, A_2, \dots \in \mathcal{S}$.

A $\sigma$-field is a generalization of the concept of an algebra (also called a field).

Definition (Algebra).
Let $X$ be a set. A collection, $\mathcal{A}$, of subsets of $X$ is an algebra (or field) if the following are satisfied:
  1. $X \in \mathcal{A}$
  2. Closed Under Complementation: $A \in \mathcal{A} \implies A^c \in \mathcal{A}$
  3. Closed Under Finite Unions: $A, B \in \mathcal{A} \implies A \cup B \in \mathcal{A}$

Now we can define measurable spaces!

Definition (Measurable Space).
Let $X$ be some set, and let $\mathcal{S}$ be a $\sigma$-field on $X$. The tuple $(X, \mathcal{S})$ is called a measurable space, and any element of $\mathcal{S}$ is called an $\mathcal{S}$-measurable set.

We’ll come back to this definition later when we discuss measures, but a measurable space is just a space that could be assigned a measure.

Let’s finish up this sub-section by introducing topological spaces and Borel sets.

Definition (Topology).
Let $X$ be a non-empty space. A topology, $\tau$, on $X$ is any collection of subsets of $X$ that satisfy:
  1. The empty set, $\emptyset$, and the entirety of $X$ are in $\tau$.
  2. The union (finite or infinite) of any subset in $\tau$ is also in $\tau$.
  3. The intersection of a finite number of subsets in $\tau$ is also in $\tau$.
We call the tuple $(X, \tau)$ a topological space.
Definition (Borel Set).
A Borel set on a topological space, $X$, denoted by $\mathcal{B}(X)$, is any subset of $X$ that can be constructed from open sets on that space in $X$ via countable unions, countable intersections, and set differences.
The Borel $\sigma$-field (or Borel algebra) is the collection of all Borel sets on a space.

This definition is a bit tricky to develop intuition for. The Borel $\sigma$-field is just the collection of all possible open sets in a given space, $X$.

An important Borel $\sigma$-field that will come up again when we discuss measures and probability is the Borel $\sigma$-field on the real line. Several examples follow from our definition:

  1. Any closed subset of $\mathbb{R}$ is a Borel set because $\sigma$-fields are closed under complementation.
  2. Any countable subset of $\mathbb{R}$ is a Borel set because $\sigma$-fields are closed under countable unions, and a single point is a closed subset of $\mathbb{R}$.
  3. Any half-open interval is a Borel set because $\sigma$-fields are closed under countable intersections.

Borel sets on $\mathbb{R}$ can also be extended to $[-\infty, \infty]$.

Along with the Borel set and the $\sigma$-field is the semialgebra.

Definition (Semialgebra).
Let $\mathcal{S}$ be a collection of sets. $\mathcal{S}$ is called a semialgebra if it satisfies the following properties:
  1. Closed Under Intersection: $S,T \in \mathcal{S} \implies S \cap T \in \mathcal{S}$
  2. (Sort Of) Closed Under Complementation: $S \in \mathcal{S} \implies S^c$ is a finite disjoint union of $T \in \mathcal{S}$

This concept will not be as useful in later discussions, but we include it for completeness. An example of a semialgebra is the union of ${ \emptyset }$ and the collection of sets that can be written as:

\[(a_1, b_1] \times \dots \times (a_d, b_d] \subset \mathbb{R}^d \hspace{5mm} \text{for } -\infty \leq a_i < b_i \leq \infty \nonumber\]

Given a semialgebra, $\mathcal{S}$, the collection of finite disjoint unions of sets in $\mathcal{S}$ forms an algebra called the algebra generated by $\mathcal{S}$.


Functions

We now need to define a concept that is at the crux of our discussions of mappings: the inverse image.

Definition (Inverse Image/Pre-Image).
Let $f: X \rightarrow Y$ be some function, and let $A \subseteq Y$. The inverse image, also called the pre-image, of subset $A$ is defined as the set: $$ f^{-1}(A) = \left\{x \in X \rvert f(x) \in A \right\} \nonumber $$ The inverse image satisfies nice properties:
  1. For any $A \subseteq Y$, $f^{-1}(Y \setminus A) = X \setminus f^{-1}(A)$
  2. For any set $\mathcal{A}$ of subsets of $Y$: $f^{-1}(\cup_{A \in \mathcal{A}} A) = \cup_{A \in \mathcal{A}}f^{-1}(A)$
  3. For any set $\mathcal{A}$ of subsets of $Y$: $f^{-1}(\cap_{A \in \mathcal{A}} A) = \cap_{A \in \mathcal{A}}f^{-1}(A)$
  4. For function $g: Y \rightarrow W$: $(g \circ f)^{-1}(A) = f^{-1}(g^{-1}(A))$s for any $A \subseteq W$

In words, the inverse image of a subset $A$ of $Y$ under function $f$ is the subset of elements in the domain $X$ that map to elements in $A$. It’s important to note that the inverse image of the whole of $Y$ does not necessarily have to be the whole of $X$!

We now introduce a definition that describes what it means for functions of a certain type to be “nice” with respect to a $\sigma$-field.

Definition (Measurable Function).
Let $(X, \mathcal{S})$ be a measurable space, and let $f: X \rightarrow [-\infty, \infty]$ be a function mapping to the extended real line. We say $f$ is $\mathcal{S}$-measurable if $f^{-1}(B) \in \mathcal{S}$ for every Borel set $B \subseteq [-\infty, \infty]$. Any function from $X$ to $\mathbb{R}$ is $\mathcal{S}$-measurable if $\mathcal{S} = \mathcal{P}(X)$, the power set of $X$.
Furthermore, for $\mathcal{S}$-measurable functions $f,g: X \rightarrow \mathbb{R}$:
  1. $f+g$, $f-g$, and $fg$ are $\mathcal{S}$-measurable.
  2. $f/g$ is $\mathcal{S}$-measurable if $g(x) \neq 0$ for all $x \in X$.
More generally, for measurable spaces $(X, \mathcal{S})$ and $(Y, \mathcal{S}')$, $f: X \rightarrow Y$ is $(\mathcal{S}, \mathcal{S}')$-measurable if, for all $E \in \mathcal{S}'$, we have $f^{-1}(E) \in \mathcal{S}$.

The basic idea behind an $\mathcal{S}$-measurable function is that we should be able to achieve any Borel set as output for some part of $\mathcal{S}$, which is in its domain (since $\text{dom}(f) = X$). It is important to remember that measurability is with respect to the $\sigma$-fields of the two measure spaces of interest.

To put it intuitively, a measurable function $f$ needs to take on values that “make sense” with respect to the $\sigma$-field of interest. For example, only constant functions are measurable with respect to the trivial $\sigma$-field ${ \emptyset, \Omega }$ for some $\Omega$. In addition, constant functions are measurable with respect to any $\sigma$-field.

Proof. Suppose we have measurable spaces $(X, \mathcal{S})$ and $(Y, \mathcal{S}')$. Let $\mathcal{S} = \{ \emptyset, \Omega \}$, and suppose we have non-constant function $f: X \rightarrow Y$. That is, there exist $a, b \in \Omega$ such that $f(a), f(b) \in \mathcal{S}'$ and $f(a) \neq f(b)$. Consider the pre-image of one of these points. We know that $f^{-1}(f(a)) = a \notin \mathcal{S}$ since $a$ is neither the null set nor the entirety of $\Omega$ (since we also have $b$ and, necessarily, $a \neq b$). To prove the second claim, consider $\mathcal{S} = \{ \emptyset, \Omega \}$ and arbitrary $\mathcal{S}'$ in the previous set-up. Since $f$ is constant, it must be the case that $f(x) = a$ for all $x \in X$ and some $a$. Pick any $s \in \mathcal{S}'$. If $a \in s$, then $f^{-1}(s) = \Omega$, since any input value maps to $a$ ($f$ is constant). If $a \notin s$, then $f^{-1}(s) = \Omega^c = \emptyset$ by the same argument. Thus, for any $s \in \mathcal{S}'$, $f^{-1}(s) \in \mathcal{S}$, implying that $f$ is $(\mathcal{S}, \mathcal{S}')$-measurable for any $\mathcal{S}'$.

To check whether a function is $\mathcal{S}$-measurable, it is sufficient to check whether \(f^{-1}((a, \infty]) = \{ x \in X \rvert f(x) > a \} \in \mathcal{S}\) for all $a \in \mathbb{R}$.

Furthermore, in the special case that $X \subseteq \mathbb{R}$ and $\mathcal{S}$ is the set of Borel subsets of $\mathbb{R}$ that are contained in $X$, then a function $f: X \rightarrow \mathbb{R}$ is called Borel measurable if $f^{-1}(B)$ is a Borel set for all Borel sets $B \subseteq \mathbb{R}$. It can be shown that any continuous or increasing function $f: X \rightarrow \mathbb{R}$ where $X$ is a Borel subset of $\mathbb{R}$ is Borel measurable.


Measures

We have finally come to the star of our discussion: the measure. A measure is a function that assigns a “size” to sets (it is similar to the idea of length for intervals or area for two dimensional regions).

Definition (Measure).
Let $(X, \mathcal{S})$ be a measure space. A function $\mu: \mathcal{S} \rightarrow [0, \infty]$ is called a measure on $(X, \mathcal{S})$ if:
  1. $\mu(\emptyset) = 0$
  2. $\mu\left(\bigcup_{i = 1}^\infty A_i \right) = \sum_{i = 1}^\infty \mu(A_i)$ for every disjoint (i.e. $A_i \cap A_j = \emptyset$ for all $i \neq j$) sequence $A_1, A_2, \dots$ of sets in $\mathcal{S}$

With this definition, we define a measure space, which is the tuple $(X, \mathcal{S}, \mu)$. For measure space $(X, \mathcal{S}, \mu)$ and $A, B \in \mathcal{S}$ such that $A \subseteq B$, we have that $\mu(A) \leq \mu(B)$ and $\mu(B \setminus A) = \mu(B) - \mu(A)$ (assuming that $\mu(A)$ is finite). We also have the additional property of countable subadditivity, which is basically a generalization of Boole’s inequality:

\[\mu\left(\bigcup_{i = 1}^\infty A_i \right) \leq \sum_{i = 1}^\infty \mu(A_i) \nonumber\]

for any sequence of sets $A_1, A_2, \dots \in \mathcal{S}$. Measures also satisfy $\mu(A \cup B) = \mu(A) + \mu(B) - \mu(A \cap B)$ (assuming that $\mu(A \cap B)$ is finite).

If we have two $\sigma$-finite (see below) measure spaces, $(X, \mathcal{S}, \mu_1)$ and $(Y, \mathcal{T}, \mu_2)$, we can define two addition sets:

\[\Omega = X \times Y = \{ (x, y): x \in X, y \in Y\} \hspace{5mm} \text{and} \hspace{5mm} \mathcal{U} = \{ S \times T: S \in \mathcal{S}, T \in \mathcal{T}\} \nonumber\]

Sets $U \in \mathcal{U}$ are rectangles. Let $\mathcal{F} = \mathcal{S} \times \mathcal{T}$ be the $\sigma$-filed generated by $\mathcal{U}$. The unique measure $\mu = \mu_1 \times \mu_2$ on $\mathcal{F}$ defined as $\mu(S \times T) = \mu_1(S) \mu_2(T)$ is called a product measure. This result can be extended to finitely many $\sigma$-finite measurable spaces.

Characteristics

Measures can be characterized in a variety of ways. First, consider the $\sigma$-finite measure.

Definition ($\sigma$-Finite).
Let $(X, \mathcal{S})$ be a measure space, and let $\mu$ be a measure defined on it. We call $\mu$ a $\sigma$-finite measure if any of the following are satisfied:
  • There exist countably many $A_1, A_2, \dots \in \mathcal{S}$ with $\mu(A_n) < \infty$ for all $n \in \mathbb{N}$ such that $\bigcap_{n \in \mathbb{N}} = X$. That is, $X$ can be covered with the intersection of countably many measurable sets in $\mathcal{S}$.
  • There exist disjoint and countably many $B_1, B_2, \dots \in \mathcal{S}$ with $\mu(B_n) < \infty$ for all $n \in \mathbb{N}$ such that $\bigcup_{n \in \mathbb{N}} = X$. That is, $X$ can be covered by the union of countably many disjoint measurable sets in $\mathcal{S}$.
  • There exist countably many $C_1, C_2, \dots \in \mathcal{S}$ with $C_1 \subset C_2 \subset \dots$ with $\mu(C_n) < \infty$ for all $n \in \mathbb{N}$ such that $\bigcup_{n \in \mathbb{N}} C_n = X$. That is, $X$ can be covered with the union of a countable monotone sequence of measurable sets in $\mathcal{S}$.
  • There exists a function $f$ such that $f(x) > 0$ for all $x \in X$ and $\int f(x) \mu(dx) < \infty$. That is, there exists a strictly positive function with finite integral that is measurable with respect to $\mu$.

We can also define a sense of continuity to measures.

Definition (Absolute Continuity).
Let $\mu$ be a measure on the Borel subsets of $\mathbb{R}$. We call $\mu$ absolutely continuous with respect to the Lebesgue measure, $\lambda$, (see below for definition), if, for every $\lambda$-measurable set $A$, $\lambda(A) = 0$ implies $\mu(A) = 0$. This condition is denoted by $\mu << \lambda$, and we say that $\mu$ is dominated by $\lambda$.

Measures can also be “coarsened” by restricting the $\sigma$-field on which they operate.

Definition (Restricted Measure)3.
Let $(\Omega, \mathcal{F}, \mu)$ be a measure space. Let $\mathcal{F}'$ be a sub-$\sigma$-field of $\mathcal{F}$. The restricted measure of $\mu$ to $\mathcal{F}'$ is the mapping $\nu: \mathcal{F}' \rightarrow \mathbb{R} \cup \{ -\infty, +\infty\}$ such that $\nu(E') = \mu(E')$ for all $E' \in \mathcal{F}'$.

A restricted measure is basically the original measure but its domain is shrunken to whatever sub-$\sigma$-field it is restricted to.

Measures also satisfy several properties.

Theorem (Theorem 1.1.11).
Let $\mu$ be a measure on $(\Omega, \mathcal{F})$, and let $A_i \uparrow A$ denote $A_1 \subset A_2 \subset \dots$ with $\cup_i A_i = A$. The measure $\mu$ satisfies the following:
  • Monotonicity: $A \subset B \implies \mu(A) \leq \mu(B)$
  • Subadditivity: $A \subset \cup_{m = 1}^\infty A_m \implies \mu(A) \leq \sum_{m= 1}^\infty \mu(A_m)$
  • Continuity From Below: $A_i \uparrow A \implies \mu(A_i) \uparrow \mu(A)$
  • Continuity From Above: $A_i \downarrow A \implies \mu(A_i) \downarrow \mu(A)$
Proof. TODO: FINISH PROOF.

A sense of “convergence” with respect to a measure can be defined for measurable functions.

Definition (In Measure).
Let $\mu$ be a $\sigma$-finite probability measure, and let $f$ be a measurable function and let $\{ f_n \}_{n = 1}^\infty$ be a sequence of measurable functions. We say that $f_n \rightarrow f$ in measure if, for any $e > 0$, we have that: $$ \mu\left(\{ x : \rvert f_n(x) - f(x) \rvert > e \right) \rightarrow 0 \hspace{5mm} \text{ as } n \rightarrow \infty \nonumber $$

Before we can move on to some of the core concepts in probability theory, we need one more definition.

Definition (Measurable Map).
Let $(\Omega, \mathcal{F})$ and $(S, \mathcal{S})$ be two measurable spaces. The function $X: \Omega \rightarrow S$ is a measurable map (from $(\Omega, \mathcal{F})$ to $(S, \mathcal{S})$) if, for all $B \in \mathcal{S}$: $$ X^{-1}(B) = \{ \omega: X(\omega) \in B \} \in \mathcal{F} \nonumber $$

Probability

With our building blocks in place, we can move on to probability theory. We’ll start with a fundamental definition: the probability space, which is just a special measure space!

Definition (Probability Space).
A probability space is a measure space $(\Omega, \mathcal{F}, P)$ such that $P(\Omega) = 1$. We have the following conventions:
  1. The set $\Omega$ is called the sample space. It is the set of all possible outcomes.
  2. The $\sigma$-field $\mathcal{F}$ over $\Omega$ is called the event space. It is a measurable set of subsets of the sample space.
  3. The measure $P$ is called the probability measure. It assigns a value (in $[0, 1]$) to give an event to give a sense of that event's likelihood of occurring.

Random Variables

Using the above, we can define random variables and vectors in a rigorous way. Note that the following can be generalized to the extended real line (i.e. $\mathbb{R} \cup {-\infty, \infty }$).

Definition (Random Variable/Vector).
Let $(\Omega, \mathcal{F}, P)$ be a probability space, and let $(\mathbb{R}^d, \mathcal{B}^d)$ be our measurable space, where $\mathcal{B}^d(\mathbb{R}^d)$ is the $\sigma$-field on $\mathbb{R}^d$. A measurable map $X: \Omega \rightarrow \mathbb{R}^d$ is called a random vector if $d > 1$ and a random variable otherwise.

Random variables map each element in the sample space to an element in $H$, which is the set of all possible values the variable can take on. Naturally, we need the pre-image of all elements in $\mathcal{H}$ to be in $\mathcal{F}$. When we refer to a random variable being measurable with respect to some $\mathcal{F}’$ (a sub-$\sigma$-field of $\mathcal{F}$), we mean that it is $(\mathcal{F}’, \mathcal{B}(\mathbb{R}))$-measurable.

Theorem (Theorems 1.3.5, 1.3.6, 1.3.71).
Let $X_1, X_2, \dots$ be random variables, and let $f: (\mathbb{R}^n, \mathcal{B}^n) \rightarrow (\mathbb{R}, \mathcal{B})$ be a measurable function. Then the following are also random variables:
  • $f(X_1, \dots, X_n)$
  • $X_1 + \dots + X_n$
  • $\underset{n}{\inf} X_n$
  • $\underset{n}{\sup} X_n$
  • $\underset{n}{\lim \inf} X_n$
  • $\underset{n}{\lim \sup} X_n$

The distribution of a random variable can also be defined from a measure theoretic perspective.

Definition (Distribution).
Let $(\Omega, \mathcal{F}, P)$ be a probability space. The probability measure induced by a random variable, $X$, defined on this space is called its distribution and is defined as: $\mu(A) = \mathbb{P}_P(X \in A)$ for all Borel sets $A \in \mathcal{B}$.
The distribution function, $F(x) = \mathbb{P}_P(X \leq x)$, describes the distribution of $X$ and satisfies the following properties:
  1. $F$ is non-decreasing
  2. $\underset{x \rightarrow \infty}{\lim} F(x) = 1$ and $\underset{x \rightarrow -\infty}{\lim} F(x) = 0$
  3. $\underset{y \downarrow x}{\lim} F(x) = 0$ (right continuous)
  4. $F(x-) = \mathbb{P}_P(X < x)$
  5. $\mathbb{P}_P(X = x) = F(x) - F(x-)$
where $F(x-) = \underset{y \uparrow x}{\lim} F(y)$.
Proof. TODO: Add Proof (p 11)

Durrett provides the best intuition for the distribution of a random variable: “In words, we pull $A \in \mathbb{B}$ back to $X^{-1}(A) \in \mathcal{F}$ and then take $P$ of that set”1.

It’s important to remember that two different random variables can induce the same distribution. In this case, we say that the random variables (denote them by $X$ and $Y$) are equal in distribution, which we denote with $X \overset{d}{=} Y$.

A distribution function with the form $F(x) = \int_{-\infty}^x f(y) dy$) can also be described by its density function, $f$, satisfying: \(\mathbb{P}(X = x) = \underset{e \rightarrow 0}{\lim} \int_{x - e}^{x + e} f(y) dy = 0 \nonumber\) In this case, we say that $F$ is absolutely continuous. Integrating the density function over the entire sample space/support will equal $1$, and the density function will always be non-negative.

Similarly, we can define a discrete distribution function (i.e. an induced probability measure) as one in which there exists a countable set $S$ such that $P(S^c) = 0$.

Example (Discrete Distribution).
Suppose a random variable on $(\mathbb{R}, \mathcal{B})$ induces distribution $F$ such that $F(x) = 1$ for $x \geq 0$ and $F(x) = 0$ for $x < 0$. This measure is discrete, and we call it a point mass at $0$.

Now, since random variables are just measurable functions, we can use its mapping to define special $\sigma$-fields.

Definition (Generated $\sigma$-Field - Function).
Let $(\Omega, \mathcal{A})$ and $(S, \mathcal{B})$ be measurable spaces and let $f: \Omega \rightarrow S$ be a measurable function, then the $\sigma$-field $f^{-1}(\mathcal{A})$ is called the $\sigma$-field generated by $f$. We denote it by $\sigma(f)$. More intuitively, $\sigma(f)$ is the smallest $\sigma$-field on which $f$ is measurable.

Put intuitively, the $\sigma$-field generated by random variable $X$ is the collection of all possible subsets of the set of values $X$ can take on such that the probability of the event that $X$ takes on that value can be determined (i.e. is measurable).

We can also define $\sigma$-fields generated by arbitary subsets. This is the smallest $\sigma$-field containing a given collection of subsets.

Definition (Generated $\sigma$-Field - Family).
Let $X$ be a set, and let $A$ be a collection of subsets of $X$. The $\sigma$-field generated by $A$, notated as $\sigma(A)$, is the collection of all subsets of $X$ that can be constructed from elements of $A$ under countable unions, intersections, and complementations. If $A = \emptyset$, then $\sigma(A) = \{ \emptyset, X\}$. If $A$ is a single event, then $\sigma(A) = \{ \emptyset, A, A^c, X \}$.
Example (Generated $\sigma$-Field).
Consider the Borel $\sigma$-field. This $\sigma$-field can be generated by any of the following sets:
  • $\{ (a, b) \rvert a, b \in \mathbb{R}, a < b \}$ or $\{ (a, \infty) \rvert a \in \mathbb{R} \}$
  • $\{ (a, b] \rvert a, b \in \mathbb{R}, a < b \}$ or $\{ (-\infty, a) \rvert a \in \mathbb{R} \}$
  • $\{ [a, b) \rvert a, b \in \mathbb{R}, a < b \}$ or $\{ [a, \infty) \rvert a \in \mathbb{R} \}$
  • $\{ [a, b] \rvert a, b \in \mathbb{R}, a < b \}$ or $\{ (-\infty, a] \rvert a \in \mathbb{R} \}$

Stochastic Processes

We can also think of having many random variables, each associated with some step in a sequence (perhaps time or space). We call this a stochastic process.

Definition (Stochastic Process).
Let $(\Omega, \mathcal{F}, P)$ be a probability space, and let $(S, \Sigma)$ be a measurable space. Let $T$ be an index set. We call a collection of random variables, $\{ X(t, \omega): t \in T, \omega \in \Omega \}$, taking on values in $S$ a stochastic process on $(\Omega, \mathcal{F}, P)$ with state space $(S, \Sigma)$. In other words, a stochastic process is a random function $X: T \times \Omega \rightarrow S$. We will sometimes alternatively denote a stochastic process with $\{ X(t): t \in T\}$ and $(X_t)_{t \in T}$.

Stochastic processes can be characterized by their continuity (or lack thereof).

Definition (Right-Continuous).
Let $X: T \times \Omega \rightarrow S$ be a stochastic process on probability space $(\Omega, \mathcal{F}, P)$ with (measurable) state space $(S, \Sigma)$. If, for all $\omega \in \Omega$, there exists $\epsilon > 0$ such that $X_s(\omega) = X_t(\omega)$ for all $s, t$ such that $t \leq s \leq t + \epsilon$, then we call $(X_t)_{t \in T}$ a right-continuous stochastic process.
Note that this definition requires the index set to be over the non-negative reals.

Integration

Before we can look at random variables any further, we need to discuss a very important concept in mathematics. In the following, we will restrict our discussion to $\mathbb{R}$, but the definitions can easily be generalized to higher dimensions by exchanging lengths for volumes via Cartesian products.

First, we define a special indicator function that got a fancy name (not sure why).

Definition (Characteristic Function).
Let $E \subseteq X$. We define the characteristic function of $E$ as the function $\chi_E: X \rightarrow \mathbb{R}$ defined by: $$ \chi_E(x) = \begin{cases} 1 & \text{if } x \in E \\ 0 & \text{if } x \notin E \end{cases} \nonumber $$

Though not very useful for our discussion, we’ll define the outer measure of a set $A \subseteq \mathbb{R}$. The outer measure formalizes the size of a set by using the lengths of open intervals.

Definition (Outer Measure).
Let $I$ be some open interval on the real line, and let $\ell(I)$ denote the length of $I$ defined as: $$ \ell(I) = \begin{cases} b - a & \text{if } I = (a, b) \text{ for some } a, b \in \mathbb{R} \text{ such that } a < b \\ 0 & \text{if } I = \emptyset \\ \infty & \text{otherwise } \end{cases} \nonumber $$ We define the outer measure of $A \subseteq \mathbb{R}$ as: $$ \rvert A \rvert = \inf\left\{ \sum_{i = 1}^\infty \ell(I_i) \bigg\rvert I_1, I_2, \dots \text{ are open intervals such that } A \subseteq \bigcup_{i = 1}^\infty I_i \right\} \nonumber $$ The outer measure satisfies:
  1. $\rvert A \rvert \leq \rvert B \rvert$ for $A \subseteq B \subseteq \mathbb{R}$
  2. $\rvert \{ t + a \rvert a \in A \}\rvert = \rvert A \rvert$ for $t \in \mathbb{R}$ and $A \subseteq \mathbb{R}$
  3. $\rvert \cup_{i = 1]^\infty} A_i \rvert \leq \sum_{i = 1}^\infty \rvert A_i \rvert$ for $A_1, A_2, \dots \subseteq \mathbb{R}$

In words, the outer measure of a set is the smallest total length of some sequence of open intervals of $\mathbb{R}$ that, together, contain $A$. Finite sets have outer measure $0$ because we can make our open intervals arbitrarily “short” (i.e. force them to have length approaching $0$). By similar reasoning, any countable subset of $\mathbb{R}$ also has outer measure $0$.

It’s important to remember that the outer measure is not a true measure in the sense that we defined. However, the outer measure allows us to define a special (and true) measure called the Lebesgue measure.

Definition (Lebesgue Measure).
Let $\mathcal{B}$ be the $\sigma$-field of Borel subsets of $\mathbb{R}$, and let $(\mathbb{R}, \mathcal{B})$ be our measurable space. The Lebesgue measure on $(\mathbb{R}, \mathcal{B})$ is the measure such that $\mu(B) = \rvert B \rvert$ for any Borel set $B \in \mathcal{B}$.

In words, the outer measure becomes a true measure if we restrict ourselves to only Borel sets. The Lebesgue measure leads to a refinement of the idea of a measurable set. A set $A \subseteq \mathbb{R}$ is called Lebesgue measurable if it is really “close” to being a Borel set. Put formally, $A$ is Lebesgue measurable if there exists a Borel set $B \subseteq A$ such that $\rvert A \setminus B \rvert = 0$. There are also many equivalent definitions (see pg. 52 of Axler (2025)).

Note that sometimes the definition of the Lebesgue measure is alteredThe change is limited to the function’s domain (Borel vs. Lebesgue measurable sets). to mean the measure on $(\mathbb{R}, \mathcal{L})$ where $\mathcal{L}$ is the $\sigma$-field of Lebesgue measurable subsets of $\mathbb{R}$.

A function $f: A \rightarrow \mathbb{R}$ for $A \subseteq \mathbb{R}$ is Lebesgue measurable if $f^{-1}(B)$ is a Lebesgue measurable set for every Borel set $B \subseteq \mathbb{R}$.

A lot of things in probability depend upon integration. For example, expectation, variance, cumulative probability, and many more things can all be stated as some type of integral. Thus, it’s important we have a solid understanding of the integral.

Non-Negative Functions

We start with the integral of the characteristic function:

\[\int \chi_E d\mu = \mu(E) \hspace{5mm} \forall E \in \mathcal{S} \nonumber\]

Recall that a simple function is any function that takes on finitely many values. Any piecewise function with finitely many pieces is simple. We can use the integral of the characteristic function to that of simple functions by taking a linear combination.

Let $(X, \mathcal{S}, \mu)$ be a measure space, let $A_1, \dots, A_n$ be disjoint set in $\mathcal{S}$, and let $c_1, \dots, c_n \in [0, \infty]$. Then:

\[\int \left(\sum_{i = 1}^n c_i \chi_{A_i} \right) d\mu = \sum_{i = 1}^n c_i \mu(A_i) \nonumber\]

With these definitions in mind, we can define the integral of any non-negative function.

Definition (Integral of a Non-Negative Function).
Let $(X, \mathcal{S}, \mu)$ be a measure space, and let $f: X \rightarrow [0, \infty]$ be an $\mathcal{S}$-measurable function. Its integral with respect to $\mu$ is defined as: $$ \int f d\mu = \sup \left\{ \sum_{i = 1}^n c_i \mu(A_i) \bigg\rvert \text{ disjoint } A_1, \dots, A_n \in \mathcal{S}; \hspace{2mm} c_1, \dots, c_n \in [0, \infty); \hspace{2mm} f(x) \geq \sum_{i = 1}^n c_i \chi_{A_{i}}(x) \text{ for all } x \in X \right\} \nonumber $$

Real-Valued Functions

We begin with a definition.

Definition ($f^+$ and $f^-$).
Let $f: X \rightarrow [-\infty, \infty]$ be a function. We have that $f = f^+ - f^-$ and $\rvert f \rvert = f^+ + f^-$ for piecewise functions: $$ f^+ = \begin{cases} f(x) & \text{ if } f(x) \geq 0 \\ 0 & \text{ if } f(x) < 0 \end{cases} \hspace{10mm} \text{and} \hspace{10mm} f^- = \begin{cases} 0 & \text{ if } f(x) \geq 0 \\ f(x) & \text{ if } -f(x) < 0 \end{cases} \nonumber $$

Notice that if $f(x) \geq 0$, then $f^+(x) \geq 0$ and $f^-(x) = 0$. Alternatively, if $f(x) < 0$, then $f^+(x) = 0$ and $f^-(x) = -f(x) > 0$. Thus, $f^+$ and $f^-$ are both non-negative functions. This allows us extend the definition of the integral to real-valued functions.

Definition (Integral of a Real-Valued Function).
Let $(X, \mathcal{S}, \mu)$ be a measure space, and let $f: X \rightarrow [-\infty, \infty]$ be an $\mathcal{S}$-measurable function such that $\int f^+ d\mu < \infty$, $\int f^- d\mu < \infty$, or both. The integral of $f$ with respect to $\mu$ is defined as: $$ \int f d\mu = \int f^+ d\mu - \int f^- d\mu \nonumber $$ The integral is homogeneous (i.e. $\int c f d\mu = c \int f d\mu$ for any $c \in \mathbb{R}$) and additive (i.e. $\int (f + g) d\mu = \int f d\mu + \int g d\mu$ for $\mathcal{S}$-measurable $f$ and $g$ satisfying $\int \rvert f \rvert d\mu < \infty$ and $\int \rvert g \rvert d \mu < \infty$).

If we have $(\Omega, \mathcal{F}, \mu) = (\mathbb{R}^d, \mathcal{B}^d, \lambda)$, then we denote $\int f d\lambda$ with $\int f(x) dx$, and if $(\Omega, \mathcal{F}, \mu) = (\mathbb{R}, \mathcal{B}, \lambda)$ and we have some interval $E = [a, b]$, we write $\int_a^b f(x) dx$ instead of $\int_E f d\lambda$.

Integration can be restricted to a subset of the domain of a function. That is, for $E \in \mathcal{S}$:

\[\int_E f d\mu = \int f \chi_E d \mu \nonumber\]

It can also be restricted to an interval of the extended real line. First, we call a bounded function $f: [a, b] \rightarrow \mathbb{R}$ Riemann integrable if the set of points in $[a, b]$ at which $f$ is not continuous has length $0$. If we have Lebesgue measure on $\mathbb{R}$, $\lambda$, and $f: (a, b) \rightarrow \mathbb{R}$ is a Lebesgue measurable function, then for $-\infty \leq a < b \leq \infty$ we let $\int_a^b f(x) dx = \int_{(a,b)} f d\lambda$.

Radon-Nikodym

Two different measures can be related via the Radon-Nikodym Theorem, which states that (under certain conditions), there exists a function such that one measure is equivalent to the integral of the function with respect to a second measure.

Radon-Nikodym Theorem4.
Let $(X, \mathcal{S})$ be a measurable space, and let $\mu$ and $\nu$ denote two $\sigma$-finite measures on this space such that $\nu << \mu$ ($\nu$ is absolutely continuous with respect to $\mu$). Then there existgs a $\mathcal{S}$-measurable function, $f: X \rightarrow [0, \infty)$ such that, for any measurable $A \subset \mathcal{S}$: $$ \nu(A) = \int_A f d\mu \nonumber $$

A fun fact is that $f$ is unique up to some set of measure $0$ with respect to $0$. That is, for any other $g$ that satisfies the definition, $f(x) = g(x)$ for all $x \in X$ except some $x \in X’ \subset X$ such that $\mu(X’) = 0$. Such a function, $f$, is called the Radon-Nikodym derivative and can be denoted by $\frac{d \nu}{d \mu}$.

Integral Properties

Here we list and prove several properties of integrals that are ubiquitous in theoretical statistics.

Jensen's Inequality.
Let $\phi$ be a convex function (i.e. $\lambda \phi(x) + (1- \lambda)\phi(y) \geq \phi(\lambda x + (1-\lambda)y)$ for all $\lambda \in (0, 1)$, $x,y \in \mathbb{R}$). Let $\mu$ be a probability measure, and let $f$ and $\phi(f)$ be integrable. Jensen's inequality states: $$ \phi\left(\int f d\mu \right) \leq \int \phi(f)d\mu \nonumber $$
Proof.
Hölder's Inequality.
Let $\mu$ be a probability measure, and let $p, q \in (1, \infty)$ such that $\frac{1}{p} + \frac{1}{q} = 1$. Hölder's inequality states: $$ \int \rvert fg \rvert d\mu \leq \rvert \rvert f \rvert \rvert_p \rvert \rvert g \rvert \rvert_1 \nonumber $$ where $\rvert \rvert f \rvert \rvert_p = (\int \rvert f \rvert^p d\mu)^{\frac{1}{p}}$ for $1 \leq p < \infty$.
Proof.

If $p = q = 2$, the above is called the Cauchy-Schwarz inequality.

Bounded Convergence Theorem.
Let $E$ be a set of finite measure (i.e. $\mu(E) < \infty$), and let $\{ f_n \}$ be a sequence of functions that vanish on $E^c$, are uniformly pointwise bounded (i.e. $\rvert f_n(x) \rvert \leq M$), and $f_n \rightarrow f$ in measure. Then: $$ \int f d\mu = \underset{n \rightarrow \infty}{\lim} \int f_n d\mu $$
Proof.
Monotone Convergence Theorem.
Let $(\Omega, \mathcal{F}, \mu)$ be a measure space, and let $X \in \mathcal{F}$ be a measurable set. Let $\{ f_k \}_{k = 0}^\infty$ be a pointwise non-decreasing sequence of $(\mathcal{F}, \mathbb{B}(\bar{\mathbb{R}}_{\geq 0})$-measurable, non-negative functions (i.e. $0 \leq \dots \leq f_k(x) \leq f_{k+1}(x) \leq \dots \leq \infty$ for every $k \geq 1$ and $x \in X$). Then the pointwise supremum, defined as the function: $$ \underset{k}{\sup} f_k: x \rightarrow \underset{k}{\sup} f_k(x) \nonumber $$ is $(\mathcal{F}, \mathbb{B}(\bar{\mathbb{R}}_{\geq 0})$-measurable and satisfies: $$ \underset{k}{\sup} \int_X f_k d\mu = \int_X \underset{k}{\sup} f_k d\mu \nonumber $$
Proof.
Dominated Convergence Theorem.
Let $(\Omega, \mathcal{F}, \mu)$ be a measure space, and let $\{ f_k \}_{k \in T}$ be a sequence of measurable functions (with index set $T$) on this space such that $\underset{n \rightarrow \infty}{\lim} f_n(x) = f(x)$ for some function $f$ for all $x \in \Omega$ (i.e. $\{ f_k \}_{k \in T}$ converges pointwise to $f$). Suppose that our sequence is dominated by some other integrable function, $g$; that is: $$ \rvert f_n(x) \rvert \leq g(x) \hspace{5mm} \forall x \in \Omega, \hspace{2mm} \forall n \in T \nonumber $$ The Dominated Convergence Theorem states that $f_n$ and $f$ are both (Lebesgue) integrable and: $$ \underset{n \rightarrow \infty}{\lim} \int_\Omega f_n d\mu = \int_\Omega \underset{n \rightarrow \infty}{\lim} f_n d\mu = \int_\Omega f d\mu \nonumber $$
Proof.
Fatou's Lemma5.
Let $(\Omega, \mathcal{F}, \mu)$ be a measure space, and let $\{ f_n: \Omega \rightarrow [0, \infty]\}$ be a sequence of non-negative measurable functions. Then: $$ \int_X \underset{n \rightarrow \infty}{\lim} \underset{m \geq n}{\inf} f_n d\mu \leq \underset{n \rightarrow \infty}{\lim} \underset{m \geq n}{\inf} \int_X f_n d\mu \nonumber $$
Proof.
Fubini's Theorem.
Let $(X, \mathcal{S}, \mu_1)$ and $(Y, \mathcal{T}, \mu_2)$ be $\sigma$-finite measurable spaces, and let $\mu = \mu_1 \times \mu_2$ (the product meeasure). If we have a function $f$ such that $f \geq 0$ or $\int \rvert f \rvert d \mu$, then: $$ \int_X \int_Y f(x,y) \mu_2(dy) \mu_1(dx) = \int_{X \times Y} f d \mu = \int_Y \int_X f(x,y) \mu_1(dx) \mu_2(dy) \nonumber $$
Proof.

Fubini’s Theorem tells us when it is okay to exchange the order of a double integral and to compute a double integral as an interated integral.


Expectation

For a random variable, $X$ on probability space $(\Omega, \mathcal{F}, P)$, how can we describe its central tendency (i.e. what values $X$ usually takes on)? We answer this question with the following definitions, which use ideas from integration (see later in this post).

Definition (Expectation).
Let $X$ be a real-valued random variable on probability space $(\Omega, \mathcal{F}, P)$. Its expectation or expected value is defined as the following Lebesgue integral: $$ \mathbb{E}[X] = \int_\Omega X dP \nonumber $$

The expected value or expectation of a random variable is basically just integration with respect to the probability measure of the space the variable is defined on. It can be any real number or even $\infty$. Since it is just an integral, we can extend all of the results in the previous section to the expectation. The results are the same, just rewritten with $\mathbb{E}[X]$ instead of $\int_\Omega X dP$.

We can also define the conditional expectation of a random variable with respect to a particular sub-$\sigma$-field.

Definition (Conditional Expectation).
Let $(\Omega, \mathcal{F}, P)$ be a probability space, let $X: \Omega \rightarrow \mathbb{R}^n$ be a real-valued random variable with finite expectation, and let $\mathcal{H} \subseteq \mathcal{F}$ be a sub-$\sigma$-field of $\mathcal{F}$. A conditional expectation of $X$ given $\mathcal{H}$ is any $\mathcal{H}$-measurable function, $\mathbb{E}(X \rvert \mathcal{H}): \Omega \rightarrow \mathbb{R}^n$, satisfying: $$ \int_H \mathbb{E}[X \rvert \mathcal{H}] dP = \int_H X dP \hspace{5mm} \forall H \in \mathcal{H} \nonumber $$ This function exists and is unique.
Definition (Conditional Expectation).
Suppose $X \in \mathcal{F}$. Then $\mathbb{E}_P[X \rvert \mathcal{F}] = X$ itself. In this scenario, we have perfect information. Since $X \in \mathcal{F}$, we have complete knowledge of whether it occurred or not when we are given all of $\mathcal{F}$.

These definitions are a bit confusing, so let’s parse them by coming at the topic from a different angle (see this post).

Let’s say we have a random variable $X$ on some probability space $(\Omega, \mathcal{F}, P)$. We don’t know anything about it, so our best guess at its value would be some sort of weighted average over all of the possible values it could take on. These weights are determined by the probability measure, $P$, since a good guess should be closer to the more likely outcomes.

Now, suppose we know some information about $X$’s outcome (i.e. we can answer some set of questions about $X$). We could formulate this as a collection of subsets of $\Omega$. For example, if we were rolling dice, the question “Is $X$ odd?” could be contained in the set \(\{1, 3, 5\}\) or \(\{2, 4, 6\}\). We could imagine outputting a different best guess depending upon what set of information we are given, which is basically what the conditional expectation does.

In one way of thinking, $\mathbb{E}[X \rvert \mathcal{H}]$ is a random variable mapping from the possible values of $X$ to the best guesses. The condition $\int_H \mathbb{E}[X \rvert \mathcal{H}] dP = \int_H X dP$ for all $H \in \mathcal{H}$ can be thought of as enforcing the idea that, if we only are guessing values that are consistent with $H$, then our best guess using only the information in $H$ should be the same as the weighted average of $X$ itself over $H$. More concretely, if \(H = \{ 1, 3, 5\}\) in our dice rolling example, having $\int_H \mathbb{E}[X \rvert \mathcal{H} dP = \int_H X dP$ implies that, given $H$, we can guess the average of $X$ perfectly.

We can relate this measure theoretic definition with the more common ones learned in statistics courses. First, partition the sample space, $\Omega$, into disjoint sets $\Omega_1, \Omega_2, \dots$ such that $\mu(\Omega_i) > 0$ for all $i$. Let $\mathcal{F} = \sigma(\Omega_1, \Omega_2, \dots)$ be the $\sigma$-field generated by this collection of sets. For random variable $X$ defined on $(\Omega, \mathcal{F}, \mu)$, we have:

\[\mathbb{E}_\mu[X \rvert \mathcal{F}] = \frac{\mathbb{E}_\mu[X \rvert \Omega_i]}{\mu(\Omega_i)} \hspace{5mm} \text{on } \Omega_i\]

When we are given some information, $\Omega_i$, about which set in our partition $X$ can be found in, our best guess at $X$ becomes the average of $X$ over that set.

In most probability courses, we also learn about conditional expectations with respect to some other random variable. In this case, we write $\mathbb{E}[X \rvert Y]$ to mean $\mathbb{E}[X \rvert \sigma(Y)]$.


Misc

Vector Spaces

We now introduce the idea of vector spaces. We begin with the definition of a field.

Definition (Field).
A field is a set $F$ with the operations of addition ($+$) and multiplication ($\cdot$), which satisfy for any $a, b, c \in F$:
  1. Associativity: $a + (b + c) = (a + b) + c$ and $a \cdot (b \cdot c) = (a \cdot b) \cdot c$
  2. Communtatitvity: $a + b = b + a$ and $a \cdot b = b \cdot a$
  3. Identity: $\exists 0, 1 \in F$ such taht $a + 0 = a$ and $a \cdot 1 = a$
  4. Additive Inverses: $\forall a \in F$, $\exists -a \in F$ such taht $a + (-a) = 0$
  5. Multiplicative Inverses: $\forall a \in F$ such that $a \neq 0$, $\exists a^{-1} \in F$ such that $a \cdot a^{-1} = 1$
  6. Distributivity: $a \cdot (b + c) = (a \cdot b) + (a \cdot c)$

A vector space is defined with respect to a field. In generality, it is a set of elements that satisfy some special properties in relation to some field.

Definition (Vector Space).
Let $F$ be a field. A vector space, $V$, is some (non-empty) set with the operation of vector addition ($+$) and the function of scalar multiplication. Vector addition takes two vectors in $V$ and assigns them a sum, which is just a third vector in $V$. Scalar multiplication takes any a vector in $V$ and any scalar $a$ in $F$ and assigns it a product, which is another vector in $V$. This operation and function satisfy the following for any $\mathbf{u}, \mathbf{v}, \mathbf{w} \in V$ and any $a, b \in F$:
  1. Associativity: $\mathbf{u} + (\mathbf{v} + \mathbf{w}) = (\mathbf{u} + \mathbf{v}) + \mathbf{w}$
  2. Commutativity: $\mathbf{u} + \mathbf{v} = \mathbf{v} + \mathbf{u}$
  3. Vector Addition Identity: $\exists \mathbf{0} \in V$ such that $\mathbf{v} + \mathbf{0} = \mathbf{v}$ for all $\mathbf{v} \in V$
  4. Scalar Multiplication Identity: $1 \mathbf{v} = \mathbf{v}$ where $1$ is the multiplicative identity in $F$
  5. Inverses: $\exists -\mathbf{v} \in V$ for every $ \mathbf{v} \in V$ such that $\mathbf{v} + (-\mathbf{v}) = \mathbf{0}$
  6. Compatibility: $a(b \mathbf{v}) = (ab) \mathbf{v}$
  7. Distributivity: $a(\mathbf{u} + \mathbf{v}) = a \mathbf{u} + a \mathbf{v}$ and $(a + b) \mathbf{v} = a \mathbf{v} + b \mathbf{v}$

Many concepts in linear algebra and general mathematics are derived from the vector space, including linear combinations, subspaces, and bases. It’s important to note that, though we usually think of vectors as tuples, they don’t need to be. You could define a vector to be different cheeses, and as long as the definition is satisfied, it will be a valid vector space.

If we equip a vector space with a special type of map, then we get an inner product space.

Definition (Inner Product Space).
An inner product space is a vector space, $V$, over the field, $F$, of real numbers or complex numbers with the map $\langle, \cdot, \cdot, \rangle: V \times V \rightarrow F$, called an inner product, which satisfies the following for all $\mathbf{u}, \mathbf{v}, \mathbf{w} \in V$ and all $a, b \in F$:
  1. Conjugate Symmetry: $\langle \mathbf{u}, \mathbf{v} \rangle = \overline{\langle \mathbf{v}, \mathbf{u} \rangle}$
  2. Linearity (in the first argument): $\langle a \mathbf{u} + b \mathbf{v}, \mathbf{w} \rangle = a \langle \mathbf{u}, \mathbf{w} \rangle + b \langle \mathbf{v}, \mathbf{w} \rangle$
  3. Positive-Definiteness: $\langle \mathbf{u}, \mathbf{u} \rangle > 0$ for any $\mathbf{u} \neq \mathbf{0}$

Something that will be very useful is a map from a vector space to the real numbers that can be thought of as assigning a “size” to vectors in the space. We call this a norm, and if we equip a vector space with a norm, then we have a normed vector space.

Definition (Norm).
Let $V$ be a vector space over a scalar field $K$. A norm, $\rvert \rvert \cdot \rvert \rvert: V \rightarrow \mathbb{R}$, is a map satisfying:
  1. Non-negativity: $\rvert \rvert x \rvert \rvert \geq 0$ for all $x \in V$
  2. Positive definiteness: $\rvert \rvert x \rvert \rvert = 0$ if and only if $x$ is the zero vector, for all $x \in V$
  3. Absolute homogeneity: $\rvert \rvert \lambda x \rvert \rvert = \rvert \lambda \rvert \rvert \rvert x \rvert \rvert$ for all $\lambda \in K$ and $x \in V$
  4. Triangle inequality: $\rvert \rvert x + y \rvert \rvert \leq \rvert \rvert x \rvert \rvert + \rvert \rvert y \rvert \rvert$ for all $x, y \in V$

We can define the canonical norm of an inner product space as $\rvert \rvert x \rvert \rvert \sqrt{\langle x, x \rangle}$. Thus, any inner product space is a normed vector space. A special type of normed vector space is the Banach space.

Definition (Banach Space).
Let $X$ be a vector space over a scalar field (perhaps $\mathbb{R}$ or $\mathbb{C}$), and let $\rvert \rvert \cdot \rvert \rvert: X \rightarrow \mathcal{R}$ be a norm. Together, $(X, \rvert \rvert \cdot \rvert \rvert)$ form a normed space. If this space is also complete, then $(X, \rvert \rvert \cdot \rvert \rvert)$ is a Banach space.
Note: Any finite-dimensional normed vector space is a Banach space. This includes finite-dimensional Euclidean spaces.

By “complete”, we mean that the space does not have any “holes” in it. Formally put, any Cauchy sequence taking values in $X$ converges to a point in $X$ as well.

Norms can also induce what we call a distance metric or function (or just metric for short) which assigns a value to represent how “far apart” two vectors in our space are.

Definition (Distance Metric).
Let $X$ be a set. A distance metric is any function $d: X \times X \rightarrow \mathbb{R}$ satisfying the following for all $x, y, z, \in X$:
  1. $d(x, x) = 0$
  2. Positivity: $x \neq y \implies d(x,y) > 0$
  3. Symmetry: $d(x,y) = d(y,x)$
  4. Triangle inequality: $d(x,z) \leq d(x,y) + d(y, z)$

The induced metric (i.e. the distance metric induced by the norm of a vector space) is the function $d: V \times V \rightarrow \mathbb{R}$ satisfying $d(x, y) = \rvert \rvert x - y \rvert \rvert$ for all $x,y \in V$. If we combine a metric with a set, then we get a metric space, which is just a set on which we have a particular sense of distance between its elements.

Definition (Metric Space).
Let $X$ be a set, and let $d$ be a distance metric. A metric space is the ordered pair $(X, d)$. A metric space is called complete if every Cauchy sequence in $X$ converges to a point in $X$.

Using our definitions of inner product and complete metric spaces, we can define what is known as a Hilbert space.

Definition (Hilbert Space).
A Hilbert space is a real (or complex) inner product space that is also a complete metric space where the distance metric is that induced by its inner product.

Helpful Definitions

Definition (Open Cover).
An open cover of subset $A \subseteq \mathbb{R}$ is any collection $\mathcal{C}$ of open subsets of $\mathbb{R}$ such that $A \subseteq \bigcup_{C \in \mathcal{C}} C$. A finite subcover of an open cover $\mathcal{C}$ of $A$ is any finite subset of sets in $\mathcal{C}$.
Definition (Almost Every).
Let $(X, \mathcal{S}, \mu)$ be a measure space, and let $A \in \mathcal{S}$. We say that $A$ contains $\mu$-almost every element of $X$ if $\mu(X \setminus A) = 0$ (in words, if $A$ contains all of $X$ except a subset of measure $0$).
Definition (Almost Everywhere).
Let $\mu$ be a $\sigma$-finite measure on $(\Omega, \mathcal{F})$, and let $\phi$ and $\psi$ be functions on $(\Omega, \mathcal{F}, \mu)$. We say that $\phi \geq \psi$ $\mu$-almost everywhere if $\mu(\{ \omega: \phi(\omega) < \psi(\omega) \}) = 0$.
Definition (Almost Surely).
Let $(\Omega, \mathcal{F}, P)$ be a probability space. An event $A$ happens almost surely if $P(A) = 1$.
Definition (Partial Order).
A partial order is a binary relation, $\leq$, between a set, $X$, and itself satisfying the following for any $a, b, c \in X$:
  1. Reflexivity: $a \leq a$
  2. Antisymmetry: $a \leq b$ and $b \leq a \implies a = b$
  3. Transitivity: $a \leq b$ and $b \leq c \implies a \leq c$
Partial orders as defined above are sometimes called reflexive, weak, or non-strict. A strict partial order is a binary relation, $<$, between a set, $X$, and itself satisfying the following for all $a, b, c \in X$:
  1. Irreflexivity: $\neg(a < a)$
  2. Asymmetry: $a < b \implies \neg (b < a)$
  3. Transitivity: $a < b$ and $b < c \implies a < c$
Definition (Total Order)6.
A total order, also called a linear order, is a partial order satisfying the additional property for all $a, b, c \in X$:
  1. Totality: $a \leq b$ or $b \leq a$
Total orders as defined above are sometimes called non-strict. A strict total order is a strict partial order that satisfies the following additional proerpty for all $a, b \in X$:
  1. Connectivity: $a \neq b \implies a < b$ or $b < a$

Assorted Results

Doob's (First) Convergence Theorems7.
Let $(\Omega, \mathcal{F}, P)$ be a probability space, and let $\mathbb{F} = (\mathcal{F}_t)_{t \geq 0}$ be a filtration such that $\mathcal{F}_t$ is a sub-$\sigma$-field of $\mathcal{F}$ for all $t$. (That is, $(\Omega, \mathcal{F}, \mathbb{F}, P)$ is a filtered probability space). Suppose we also have $X; [0, \infty) \times \Omega \rightarrow \mathbb{R}$, a right-continuous supermartingale with respect to $\mathbb{F}$.
For $t \geq 0$, define $X^-_t = \max\{-X_t, 0 \}$. Assume $\underset{t > 0}{\sup} \mathbb{E}[X_t^-] < +\infty$. Then (the point-wise limit) $X(\omega) = \underset{t \rightarrow + \infty}{\lim} X_t(\omega)$ exists and is finite for all $\omega \in \Omega$ except a $P$-null set.

References

  1. Durrett, R. (n.d.). Probability: Theory and Examples.  2 3 4

  2. Axler, S. (2019). Measure, Integration & Real Analysis. Germany: Springer International Publishing. 

  3. Definition: Restricted Measure. ProofWiki. (2022, May 13). https://proofwiki.org/wiki/Definition:Restricted_Measure. 

  4. Wikipedia contributors. (2025, April 30). Radon–Nikodym theorem. In Wikipedia, The Free Encyclopedia. Retrieved 20:38, May 22, 2025, from https://en.wikipedia.org/w/index.php?title=Radon%E2%80%93Nikodym_theorem&oldid=1288156682 

  5. Wikipedia contributors. (2025, April 25). Fatou’s lemma. In Wikipedia, The Free Encyclopedia. Retrieved 20:34, May 22, 2025, from https://en.wikipedia.org/w/index.php?title=Fatou%27s_lemma&oldid=1287282802. 

  6. Wikipedia contributors. (2025, May 11). Total order. In Wikipedia, The Free Encyclopedia. Retrieved 20:34, May 22, 2025, from https://en.wikipedia.org/w/index.php?title=Total_order&oldid=1289905165. 

  7. Wikipedia contributors. (2025, April 14). Doob’s martingale convergence theorems. In Wikipedia, The Free Encyclopedia. Retrieved 20:35, May 22, 2025, from https://en.wikipedia.org/w/index.php?title=Doob%27s_martingale_convergence_theorems&oldid=1285525796. 

This project is maintained by aerosengart