Convergence of Positive Supermartingales
Published:
I really enjoyed reading Chapter two of Discrete Parameter Martingales and I thought why not making a blog post.
The book is quite advanced, but I’ll try my best state everything clearly and simply given my temporal boredom threshold. Haha :)!
Let me fix the probability space first. Our probability space (as usual) is $(\Omega, \mathscr{F}, \mathbb{P})$. HOWEVER, in what follows it is more convenient to define a subset of the above probability space as $(\Omega’, \mathscr{B}, \mathbb{P})$ (because we’re going to work with filterations and naturally they are sub-$\sigma$-algebras that are growing, so we’re always in the subset case). A mapping $X$ only defined on $\Omega’$ is random variable on $\Omega’$ if it is measurable with respect to (w.r.t.) the trace $\sigma$-algebra $\mathscr{F} \cap \Omega’$. So, what is a trace $\sigma$-algebra? I found it here.
Definition. Let $\Omega$ be a set, and let $\mathscr{F}$ be a $\sigma$-algebra on $\Omega$. Let $\Omega’ \subseteq \Omega$ be a subset of $\Omega$. Then, the trace $\sigma$-algebra (of $\Omega’$ in $\mathscr{F}$), $\mathscr{B}$, is defined as:
\[\mathscr{B} := \{\Omega' \cap F : F \in \mathscr{F}\}.\]It is a $\sigma$-algebra on $\Omega’$.
Now we can define what a positive supermartingale is. Let $(\Omega, \mathscr{F}, \mathbb{P})$ be our probability space and $(\mathscr{B}_n, n \in \mathbb{N})$ an increasing sub-$\sigma$-algebrs of $\mathscr{F}$ (a.k.a a filteration).
Definition. An adapted sequence $(X_n, n \in \mathbb{N})$ of positive random variables is called a positive supermartingale if the almost sure (a.s.) inequality
\[X_n \geq \mathbb{E}^{\mathscr{B}_n}(X_{n + 1})\]is satisfied for all $n \in \mathbb{N}$. A supermartingale is by definition a sequence of random variables (r.v.s) which “decrease in conditional mean”. For a sequence positive r.v.s denoting the sequence of values of the fortune of a gambler, the supermartingale condition expresses the property that at each play the game is unfavourable to the player in conditional mean.
We note that the inequality defining a supermartingale implies that
\[X_m \geq \mathbb{E}^{\mathscr{B}_m}(X_p), \; \forall p > m.\]In fact the defintion implies that,
\[\mathbb{E}^{\mathscr{B}_m}(X_n) \geq \mathbb{E}^{\mathscr{B}_m}\mathbb{E}^{\mathscr{B}_n}(X_{n + 1}) \stackrel{\text{tower rule}}{=} \mathbb{E}^{\mathscr{B}_m}(X_{n + 1}),\]if $n \geq m$. Therefore, the sequence ($\mathbb{E}^{\mathscr{B}_m}(X_n)$) is decreasing.
The following proposition is so cool and is so fundamental.
Proposition (Maximal inequality). For every positive supermartingale, the r.v. $\sup_n X_n$ is a.s. finite on the set {$X_0 < \infty$}, and satisfies the following inequality
\[\mathbb{P}^{\mathscr{B}_0}(\sup_n X_n \geq a) \leq \min \left(\frac{X_0}{a}, 1\right)\]for all constants $a > 0$.
Before going into the proof, remember that $\mathbb{P}^{\mathscr{B}}(A) = \mathbb{E}^{\mathscr{B}}(\mathbb{1}_A)$ and
\[\int_B\mathbb{P}^{\mathscr{B}}(A)d\mathbb{P} = \mathbb{P}(A \cap B), \quad B \in \mathscr{B}.\]First, we need an auxiliary lemma which constitutes the switching principle for supermartingales. But before that, I need to talk about the concept of the stopping time.
Stopping time (unsurprisingly) is a concept from gambling. A stopping rule for the player is a rule for leaving the game, based at each time on information it his/her disposal at that time. By this definition, a “dishonest” player who decides to leave the game any time already knowing certain subsequent outcomes of the game is excluded. Also, note that the stopping time could be $\infty$ meaning that the game never ends.
Definition. Let $\mathbb{N}^*$ denote $\mathbb{N} \cup {\infty}$. A mapping $\nu: \Omega \to \mathbb{N}^*$ is called a stopping time if
\[\{ \nu = n \} \in \mathscr{B}_n, \quad \forall n \in \mathbb{N}.\]The $\sigma$-algebra $\mathscr{B}_\nu$ is associated with the stopping time as the subsets of $\Omega$ defined by
\[\mathscr{B}_\nu = \{B: B \in \mathscr{B}_\infty, B \cap \{\nu = n\} \in \mathscr{B}_n, \forall n \in \mathbb{N}\}.\]Note that events in $\mathscr{B}_\nu$ are prior to $\nu$.
Back to the considered lemma:
Lemma (master). Given tow positive supermartingales $(X^{(i)}_n)(i =1, 2)$ and a stopping time $\nu$ such that $X_\nu^{(1)} \geq X_\nu^{(2)}$ on {$\nu < \infty$}, the formula
\[X_n(\omega) = \begin{cases} X_\nu^{(1)} & \mathrm{if}\; n < \nu(\omega)\\ X_\nu^{(2)} & \mathrm{if}\; n \geq \nu(\omega) \end{cases} \quad (n \in \mathbb{N}),\]defines a new positive supermartingale.
Proof. Indeed, the defining formula of the $X_n$ can also be written
\[X_n = \mathbb{1}_{\{v > n\}}X_n^{(1)} + \mathbb{1}_{\{v \leq n\}}X_n^{(2)},\]then it is clear that $X_n$ is $\mathscr{B}_n$ measurable. The supermartingale property of $X^{(i)}_n$ allows us to write
\[\begin{align*} X_n &= \mathbb{1}_{\{v > n\}}X_n^{(1)} + \mathbb{1}_{\{v \leq n\}}X_n^{(2)} \\ &\geq \mathbb{1}_{\{v > n\}} \mathbb{E}^{\mathscr{B}_n}\left[X_{n+1}^{(1)}\right] + \mathbb{1}_{\{v \leq n\}} \mathbb{E}^{\mathscr{B}_n}\left[X_{n+1}^{(2)}\right] \\ & = \mathbb{E}^{\mathscr{B}_n}\left[\mathbb{1}_{\{v > n\}} X_{n+1}^{(1)} + \mathbb{1}_{\{v \leq n\}} X_{n+1}^{(2)}\right]. \tag{since $\nu$ is $\mathscr{B}_n$-measurable} \end{align*}\]The assumption $X_\nu^{(1)} \geq X_\nu^{(2)}$ implies that $X_{n + 1}^{(1)} \geq X_{n + 1}^{(2)}$ on {$\nu = n + 1$}, and
\[\mathbb{1}_{\{v > n\}} X_{n+1}^{(1)} - \mathbb{1}_{\{v > n + 1\}} X_{n+1}^{(1)} + \mathbb{1}_{\{v \leq n\}} X_{n+1}^{(2)} - \mathbb{1}_{\{v \leq n + 1\}} X_{n+1}^{(2)} = X_{n+1}^{(1)} - X_{n+1}^{(2)}.\]Hence,
\[\mathbb{1}_{\{v > n\}} X_{n+1}^{(1)} + \mathbb{1}_{\{v \leq n\}} X_{n+1}^{(2)} \geq \mathbb{1}_{\{v > n + 1\}} X_{n+1}^{(1)} + \mathbb{1}_{\{v \leq n + 1\}} X_{n+1}^{(2)} = X_{n + 1}. \quad\square\]Using the above lemma, let us associate with the positive supermartingale of the proposition the stopping time defined by
\[\nu_a = \begin{cases} \min(n:X_n > a) = 0 & \mathrm{if}\; \sup_n X_n > a\\ \infty & \mathrm{if}\; \sup_n X_n \leq a \end{cases}\; .\]Since $X_{\nu_a} > a$ on {$\nu_a < \infty$} and since the constant $a$ can be considered a supermartingale, the formula,
\(Y_n = \begin{cases} X_n & n < \nu_a\\ a & n \geq \nu_a \end{cases}\) defines a new supermartingale by our previous lemma. Hence, $Y_0 \geq \mathbb{E}^{\mathscr{B}_0}[Y_n]$. Since $Y_0$ is equal to $X_0$ or $a$ according to the relation between $X_0$ and $a$, and since $Y_n \geq a \mathbb{1}_{{\nu_a \leq n}}$, we have:
\[a \mathbb{P}^{\mathscr{B}_0}(\nu_a \leq n) \leq \min(X_0, a).\]Letting $n$ tend to infinity, we obtain
\[\mathbb{P}^{\mathscr{B}_0}(\sup_n X_n > a) \leq \min\left(\frac{X_0}{a}, 1\right),\]Since {$\nu_a < \infty$} = {$\sup_n X_n > a$}. It suffices to replace $a$ by $a\left(1 - k^{-1}\right)$ in the inequality above and let $k$ tend to infinity to obtain the same inequality with $\geq$ instead of $>$ on the left-hand side. Let us integrate both sides over the event {$X_0 < \infty$}, which belongs to $\mathscr{B}_0$. Let $A := {\sup_n X_n > a}$ we find that
\[\begin{align*} \int_{\{X_0 < \infty\}}\mathbb{P}(A \mid \mathscr{B}_0) d\mathbb{P} & \leq \int_{\{X_0 < \infty\}} \min\left(\frac{X_0}{a}, 1\right)d\mathbb{P} \Rightarrow \\ \int_{\mathscr{B}_0}\mathbb{E}(\mathbb{1}_A \mid \mathscr{B}_0) d\mathbb{P} &\leq \int_{\{X_0 < \infty\}} \min\left(\frac{X_0}{a}, 1\right)d\mathbb{P} \Rightarrow \\ \mathbb{E}[\mathbb{1}_{\mathscr{B}_0} \cdot \mathbb{E}(\mathbb{1}_A \mid \mathscr{B}_0)] &\leq \int_{\{X_0 < \infty\}} \min\left(\frac{X_0}{a}, 1\right)d\mathbb{P} \Rightarrow \\ \mathbb{E}[\mathbb{E}(\mathbb{1}_{\mathscr{B}_0} \cdot\mathbb{1}_A \mid \mathscr{B}_0)] &\leq \int_{\{X_0 < \infty\}} \min\left(\frac{X_0}{a}, 1\right)d\mathbb{P} \Rightarrow \\ \mathbb{E}(\mathbb{1}_{\mathscr{B}_0} \cdot\mathbb{1}_A) &\leq \int_{\{X_0 < \infty\}} \min\left(\frac{X_0}{a}, 1\right)d\mathbb{P} \Rightarrow \\ \mathbb{P}(\mathscr{B}_0 \cap A) &\leq \int_{\{X_0 < \infty\}} \min\left(\frac{X_0}{a}, 1\right)d\mathbb{P} \Rightarrow \\ \mathbb{P}(X_0 < \infty, \sup_n X_n > a) &\leq \int_{\{X_0 < \infty\}} \min\left(\frac{X_0}{a}, 1\right)d\mathbb{P}. \end{align*}\]As $a$ tends to infinity, the R.H.S. tends to zero by the dominated convergence theorem and we have
\[\mathbb{P}(X_0 < \infty, \sup_n X_n = \infty) = 0. \; \square\]Box. (Analysis II) A box $B$ in $\mathbb{R}^n$ is any set of the form
\[B = \Pi_{i = 1}^n (a_i, b_i) := \{(x_1, \dots, x_n) \in \mathbb{R}^n: x_i \in (a_i, b_i)\: \text{for all}\: 1 \leq i \leq n\},\]where $b_i \geq a_i$ are real numbers. The volume of this box is defined as
\[\mathrm{vol}(B) := \Pi_{i=1}^n (b_i - a_i).\]Outer Measure. (Analysis II) If $\Omega$ is a set, we define the outer measure $m^*(\Omega)$ of $\Omega$ to be quantity
\[m^*(\Omega) := \inf \left\{\sum_{j \in J}\mathrm{vol}(B_j): (B_j)_{j \in J}\: \mathrm{covers}\: \Omega;\: J\: \text{at most countable}\right\}.\]Measurable Sets. (Analysis II) Let $E$ be a subset of $\R^n$. We say $E$ is Lebesgue measurable, or measurable for short, iff we have the identity
\(m^*(A) = m^*(A \cap E) + m^*(A-E)\) for every subset $A$ of $\R^n$. If $E$ is measurable, we define the Lebesgue measure of $E$ to be $m(E)= m^*(E)$; if $E$ is not measurable, we leave $m(E)$ undefined. In words, $E$ being measurable means that if we use the set $E$ to divide up an arbitrary set $A$ into two parts, we keep the additivity property.
Measurable Functions. (Analysis II) Let be a measurable subset of $\mathbb{R}^n$ and let $f: \Omega \to \R^m$ be a function. A function $f$ is measurable iff $f^{-1}(V)$ is measurable for every open set $V \subseteq \R^m$. Another characterization of measurable functions is given by: Let $\Omega$ be a measurable subset of $\R^n$. A function $f: \Omega \to \R^*$ is said to be measurable iff $f^{-1}((a, \infty]))$ is measurable for every real number $a$.
Absolutely Integrable. (Analysis II) Let $\Omega$ be a measurable subset of . A measurable function $f: \Omega \to \mathbb{R}^* := \mathbb{R} \cup {-\infty, +\infty}$ is said to be absolutely integrable if the integral $\int_\Omega f$ (w.r.t. the Lebesgue measure) is finite.
Pointwise Convergence. (Analysis II) The most obvious notion of convergence of functions is pointwise convergence, or convergence at each point of the domain. Let $\left(f^{(n)} \right)_{n = 1}^\infty$ be a sequence of functions from one metric space $(X, d_X$) to another $(Y, d_Y)$, and let $f: X \to Y$ be another function. We say that $\left(f^{(n)} \right)_{n = 1}^\infty$ converges pointwise to $f$ if we have
\[\lim_{n \to \infty}f^{(n)}(x) = f(x), \quad \forall x \in X,\]i.e.,
\[\lim_{n \to \infty}d_Y\left(f^{(n)}(x), f(x)\right) = 0.\]We call the function f the pointwise limit of the functions $f^{(n)}$.
Limits of measurable functions are measurable (Analysis II) Let $\Omega$ be a measurable subset of $\mathbb{R}^n$, for each positive integer $n$, let $f_n: \Omega \to \mathbb{R}^*$ be a measurable function. Then, the functions $\sup_n f_n, \inf_n f_n, \limsup_{n \to \infty} f_n, \mathrm{and}\, \liminf_{n \to \infty} f_n$ are also measurable. In particular, if $f_n$ converge pointwise to another function $f: \Omega \to \mathbb{R}^*$, then $f$ is also measurable.
Proof. We first prove the claim about $\sup_n f_n$. Call this function $g$. We have to prove that $g^{-1}((a, \infty]))$ is measurable for every $a$. First we show that
\[g^{-1}((a, \infty])) = \cup_{n \geq 1}f_n^{-1}((a, \infty])),\]then the claim follows since the countable union of measurable sets is measurable. To show the above display consdier two cases:
- $\subseteq$
If $x \in g^{-1}((a, \infty]))$, then $g(x) = sup_n f_n(x) > a$. If every $f_n(x) \leq a$, then $\sup_n f_n(x) \leq a$ which is a contradiction. Hence, there exists some $n$ with $f_n(x) > a$. Thus, $x \in f^{-1}((a, \infty]))$ for that $n$, so $x$ belongs to the union.
- $\supseteq$
If $x \in \cup_{n \geq 1}f_n^{-1}((a, \infty]))$, then for some $n$ we have $f_n(x) > a$. Therefore, $g(x) = \sup_n f_n(x) \geq f_n(x) > a$ and $x \in g^{-1}((a, \infty]))$.
Since both inclusions hold, the sets are equal.
One important note that we know {$g > a$} $= \cup_n$ {$f_n > a$}. But, {$g \geq a$} $\neq \cup_n$ {$f_n \geq a$}. Because if the supremum is at least $a$ does not mean that at least one of the $f_n$s will match it. We can fix this by using an approximation from below:
\[\{g \geq a\} = \cap_{m = 1}^\infty \cup_{n = 1}^\infty \{f_n \geq a - \frac 1m\}.\]Back to the main proposition, a similar argument works for $\inf_n f_n$. The claim for $\limsup$ and $\liminf$ then follow from the identities
\[\limsup_{n \to \infty} f_n = \inf_{N \geq 1} \sup_{n \geq N} f_n, \quad \liminf_{n \to \infty} f_n = \sup_{N \geq 1} \inf_{n \geq N} f_n.\]Lebesgue Monotone Convergence Theorem. (Analysis II) Let $\Omega$ be a measurable subset of $\R^n$, and let $f_n$ be a sequence of non-negative functions from $\Omega$ to $[0, \infty]$, which are increasing in the sense that
\[0 \leq f_1(x) \leq f_2(x) \leq \dots \quad \forall x \in \Omega,\]Note we are assuming that $f_n(x)$ is increasing with respect to $n$; this is a different notion from $f_n(x)$ increasing with respect to $x$. We have
\[0 \leq \int_\Omega f_1 \leq \int_\Omega f_2 \leq \dots\]and
\[\int_\Omega \sup_n f_n = \sup_n \int_\Omega f_n.\]Proof. For the first conclusion, we should prove that
If $0 \leq f(x) \leq g(x)$ for all $x \in \Omega$ for measurable non-negative functions $f, g$, then we have $\int_\Omega f \leq \int_\Omega g$. Let $h := g - f \geq 0$. We proceed by showing that $0 \leq \int_\Omega h \leq \infty$. Since $f, g$ are measurable, $h$ is measurable and since $h$ is non-negative as well, its Lebesgue integral is equal to
\[\int_\Omega h = \sup \left\{\int_\Omega s: s \text{ is simple, non-negative and dominated by } h \right\}.\]Fix an $s$. Since $s$ is simple, it can be written as $s = \sum_{j = 1}^{N} c_j m(E_j)$, where $c_j > 0, \cup_j E_j = \Omega$. The sum of non-negative numbers is between zero and $\infty$, hence their $\sup$ is between zero and infinity.
Now the second conclusion. Following a similar argument of above,
\[\int_\Omega \sup_m f_m \geq \int_\Omega f_n\]for every $n$ including its superimum, i.e.,
\[\int_\Omega \sup_m f_m \geq \sup_n \int_\Omega f_n.\]If we show that $\int_\Omega \sup_m f_m \leq \sup_n \int_\Omega f_n$, then the proof will be completed. To proceed, by the definition of $\int_\Omega \sup_m f_m$, it is sufficient to show that for all non-negative measurable function $s$ that it is dominated by $\sup_m f_m$ the following holds
\(\int_\Omega s \leq \sup_n \int_\Omega f_n.\) If we show that
\(\int_\Omega s \leq \sup_n \int_\Omega f_n + \epsilon\int_\Omega s\) for every $0 < \epsilon < 1$ (note that our functions are all non-negative hence the range of the epsilon); the claim follows by taking limits as $\epsilon \to 0$. $(\star)$
By construction,
\[s(x) \leq \sup_n f_n (x), \quad \forall x \in \Omega.\]Hence, for all $x \in \Omega$, there exists an $N$ such that
\(f_N(x) \geq (1 - \epsilon)s(x).\) Since $f_n$ are increasing, this implies taht $f_n(x) \geq (1 - \epsilon)s(x)$ for all $n \geq N$.
Define the sets
\(E_n := \{x \in \Omega: f_n(x) \geq (1 - \epsilon) s(x)\}.\) Since $f_n$ is measurable by assumption, $s$ is measurable because it’s a simple function, the function $f_n(x) - s(x) + \epsilon f(x)$ is measurable, hence is each $E_n$. We also have that $E_1 \subseteq E_2 \subseteq \dots$, and $\cup_{n = 1}^\infty E_n = \Omega$. We have that
\[(1 - \epsilon) \int_{E_n}s = \int_{E_n} (1 - \epsilon)s \leq \int_{E_n} f_n \stackrel{(\star\star)}{\leq} \int_{\Omega} f_n,\]if we show that $(\star\star)$ holds. Then by taking limits as $\epsilon \to 0$ and taking suprema, it suffices to show that
\[\sup_n \int_{E_n}s = \int_{\Omega}s,\]which would prove $(\star)$.
Since $s$ is a simple function, $\int_\Omega s = \sum_{j = 1}^N c_j \cdot m(F_j)$, where $F_1, F_2, \dots$ are disjoint and in $\Omega$. Similarly, $\int_{E_n} s = \sum_{j = 1}^N c_j \cdot m(F_j \cap E_n)$. Since the sums are finite and the summands are positive, it suffices to show that $\sup_n m(F_j \cap E_n) = m(F_j)$. $(\star\star\star)$
So to summarize, we need to prove the following that prove $(\star)$.
- $\int_{E_n} f_n \leq \int_{\Omega} f_n$. $(\star\star)$
- $\sup_n m(F_j \cap E_n) = m(F_j)$. $(\star\star\star)$
The following proposition proves $(\star\star)$:
Let $\Omega$ be a measurable set, and $f: \Omega \to [0, \infty]$ be a non-negative measurable function. If $\Omega’ \subseteq \Omega$ is measurable, then $\int_{\Omega’} f = \int_{\Omega} f \mathbb{1}_{{\Omega’}} \leq \int_\Omega f$.
Proof. By the definition of the Lebesgue integral
\[\begin{align*} \int_\Omega f &= \sup \left\{\int_\Omega s: s \text{ is simple, non-negative and dominated by } f \right\}, \\ \int_{\Omega'} f &= \sup \left\{\int_{\Omega'} s: s \text{ is simple, non-negative and dominated by } f \right\}. \end{align*}\]Fix $s$. We have
\[\begin{align*} \int_\Omega s &= \sum_{j = 1}^N c_j \cdot m(F_j), \\ \int_{\Omega'} s &= \sum_{j = 1}^N c_j \cdot m(F_j \cap \Omega'). \end{align*}\]But, $F_j \cap \Omega’ \subseteq F_j$. Therefore, $m\left(F_j \cap \Omega’\right) \leq m(F_j)$. By multiplying both sides by positive constants $c_j$ and summing over $j$ we get that $\int_{\Omega’} s \leq \int_{\Omega} s$. Taking suprema from both sides competes the proof.
The following proposition helps proving $(\star\star\star)$:
If $A_1 \subseteq A_2 \subseteq \dots$ is an increasing sequence of measurable sets, then
\[m\left(\cup_{j = 1}^\infty A_j \right) = \lim_{j \to \infty} m(A_j).\]Proof.
Define
\[B_1 := A_1, \quad B_j = A_j \setminus A_{j - 1}, \quad j \geq 1.\]Then, $A := \cup_{j = 1}^\infty A_j = \cup_{j = 1}^\infty B_j$, and $A_N := \cup_{j = 1}^N A_j = \cup_{j = 1}^N B_j$. Hence,
\[m(A) = m\left(\cup_{j = 1}^\infty B_j\right) = \lim_{N \to \infty} m\left(\cup_{j = 1}^N B_j\right) = \lim_{N \to \infty} m(A_N).\]So in $(\star\star\star)$, we have
\[\sup_n m(F_j \cap E_n) \stackrel{\text{(monotonicity)}}{=} \lim_{n \to \infty} m(F_j \cap E_n) = m\left(F_j \cap \left(\cup_{n = 1}^\infty E_n\right) \right) = m\left(F_j \cap \Omega \right) = m(F_j).\]Fatou’s Lemma. (Analysis II) Let $\Omega$ be a measurable subset of $\mathbb{R}^n$, and let $f_1, f_2, \dots$ be a sequence of non-negative functions from $\Omega$ to $[0, \infty]$. Then
\[\int_\Omega \liminf_{n \to \infty} f_n \leq \liminf_{n \to \infty}\int_\Omega f_n.\]Proof.
\[\begin{align*} \int_\Omega \liminf_{n \to \infty} f_n & = \int_\Omega \sup_{m \geq 1} \inf_{n \geq m} f_n \\ & = \sup_{m \geq 1} \int_\Omega \inf_{n \geq m} f_n \tag{monotone convergence}. \end{align*}\]$\inf_{n \geq m} f_n \leq f_j$ for all $j \geq m$. Hence, $\int_\Omega \inf_{n \geq m} f_n \leq \int_\Omega f_j$. By taking infima w.r.t. $j$, we have
\[\int_\Omega \inf_{n \geq m} f_n \leq \inf_{j \geq m}\int_\Omega f_j.\]Therefore, we obtained the following that completes the proof
\[\int_\Omega \liminf_{n \to \infty} f_n \leq \sup_{m \geq 1}\int_\Omega \inf_{n \geq m} f_n \leq \sup_{m \geq 1}\inf_{j \geq m}\int_\Omega f_j = \liminf_{n \to \infty} \int_{\Omega} f_n.\]Lebesgue Dominated Convergence Theorem. (Analysis II) Let $\Omega$ be a measurable subset of $\mathbb{R}^n$, and let $f_1, f_2, \dots$ be a sequence of measurable functions from $\Omega$ to $\mathbb{R} \cup {-\infty, +\infty}$ which converge pointwise. Suppose also that there is an absolutely integrable function $F: \Omega \to [0, \infty]$ such that $|f_n(x)| < F(x)$ for all $x \in \Omega$ and all $n = 1, 2, 3, \dots$. Then,
\[\int_{\Omega} \lim_{n \to \infty} f_n = \lim_{n \to \infty}\int_{\Omega} f_n.\]Proof. If $F$ was infinite on a set of non-zero meaure, then $F$ would not be absolutely integrable thus the set where $F$ is infinite has zero measure. We may delete this set from (this does not affect any of the integrals) and thus assume without loss of generality that $F(x)$ is finite for every $x \in \Omega$, which implies the same assertion for the $f_n(x)$.
Let $f:\Omega \to \mathbb{R} \cup$ {$-\infty, +\infty$} be the function $f(x) := \lim_{n \to \infty}f_n(x)$ which exists by assumption. Since $f$ is the limit of measurable functions, it’s measurable. Also, since $|f_n(x)| \leq F(x)$ for all $n$ and all $x \in \Omega$ , we see that each $f_n$ is absolutely integrable, and by taking limits we obtain $| f(x)| \leq F(x)$ for all $x \in \Omega$ , so $f$ is also absolutely integrable. Our task is to show that $\lim_{n \to \infty}\int_{\Omega} f_n = \int_\Omega f$. The functions $F + f_n$ are non-negative and converge pointwise to $F + f$. So by Fatou’s lemma
\[\int_\Omega F + f \leq \int_\Omega F + \liminf_{n \to \infty} \int f_n \Rightarrow \int_\Omega f \leq \liminf_{n \to \infty} \int f_n.\]But also, $F - f_n$ are non-negative and converge pointwise to $F + f$. So, again by Fatou’s lemma
\[\int_\Omega F - f \leq \int_\Omega F + \limsup_{n \to \infty} \int f_n \Rightarrow \int_\Omega f \geq \limsup_{n \to \infty} \int f_n.\]Hence, we showed that
\[\int_\Omega f \leq \liminf_{n \to \infty} \int f_n \leq \limsup_{n \to \infty} \int f_n \leq \int_\Omega f.\]Hence, the $\liminf$ and $\limsup$ of $\int_\Omega f_n$ are equal as we wanted.
Remark. The preceding proof is valid without any change if we change the constant $a$ with any $\sB_0$-measurable r.v. Say $A$ is such an r.v. Then, it means if $\mathbb{P}^{\sB_0}(A \leq \sup_n X_n) = 1$, then $A \leq X_0$, i.e., $X_0$ is largest $\sB_0$-measurable r.v. dominated by $\sup_n X_n$. More generally, for any $\sB_p$-measurable r.v. $A$ that satisfies $A \leq \sup_{n \leq p} X_n$ can be seen as applying the preceding results to the supermartigale $(\sup_{n \leq p} X_n, X_{p+1}, X_{p+2}, \dots)$ adapted to the $(\sB_p, \sB_{p + 1}, \dots)$.
Theorem. Every positive supermartingale $(X_n)$ almost surely converges. Furthermore, the limit $X_\infty = \lim_{n \to \infty} X_n$ a.s. satisfies the following inequality
\[\Ev{X_\infty}{\sB_n} \leq X_n, \quad n \ \in \mathbb{N}.\]Proof.
First, we’ll discuss what it means that a sequence of real numbers converges.
Given a sequence of real number $(x_n)$ in $\R^*$ and a pair of real number $a, b$ where $a<b$, let us define $\nu_k (k \geq 1)$ as some special moments in time, where
\[\begin{align*} \nu_1 &= \min\{n: n \geq 1, x_n \leq a\} \\ \nu_2 &= \min\{n: n \geq \nu_1, x_n \geq b\} \\ \nu_3 &= \min\{n: n \geq \nu_2, x_n \leq a\} \\ \nu_4 &= \min\{n: n \geq \nu_3, x_n \geq b\} \\ & \vdots . \end{align*}\]If some $\nu_k$ is not defined, we put it equal to $\infty$ and all subsequent indices. Let $\beta_{a, b}$ be the largest value of $p$ where $\nu_{2p}$ is finite. If all $\nu_k$ are finite, then $\beta_{a, b} = \infty$. $\beta_{a, b}$ denote the number of upcrossings of the sequence $(x_n$) on $[a, b]$. We can see that
\[\liminf_{n \to \infty} x_n < a < b < \limsup_{n \to \infty} x_n \Rightarrow \beta_{a, b} = \infty \Rightarrow \liminf_{n \to \infty} x_n \leq a < b \leq \limsup_{n \to \infty} x_n.\]Therefore, we can deduce that a sequence in $\R^*$ is convergent iff $\beta_{a, b} <\infty$ for every $a < b \in R$. Now, let’s considering the case of the sequence of r.v.s $(X_n)$. Note that the r.v.s $\nu_k(\omega$) are $\sB_n$-measurable. This is due to
\(\{\nu_{2p} = n\} = \cup_{m < n} \{\nu_{2p - 1} = m\: \mathrm{and}\: X_{m + 1} < b, \dots, X_{n - 1} < 1, X_n \geq b\},\) and a similar argument for the odd indices. The event {$\beta_{a, b} \geq p$} $=$ {$\nu_{2p} < \infty$} shows that $\beta_{a, b}$ is also an r.v. Hence, the convergence criterion is
\[\{X_n \to \cdot\} = \cap_{a < b \in \R} \{\beta_{a, b} < \infty\} \stackrel{\Q\, \text{is dense}}{=} \cap_{a < b \in \Q} \{\beta_{a, b} < \infty\}.\]Rationals are dense. (Analysis 1) If $x$ and $y$ are two rationals that $x < y$, then there exists a third rational $z$ such that $x < z < y$.
Proof.
Set $z := \frac{x + y}{2}$. Since $x <y$, then $\frac{x}{2} < \frac{y}{2}$. If we add $\frac{y}{2}$ to the both sides, we get $z < y$. A symmetrical argument shows $x < z$, hence $x < z < y$. $\square$
A homeomorphism. (Topology) Let $X$ and $Y$ be topological spaces (like open intervals); let $f: X \to Y$ be a bijection. If both the function $f$ and the inverse function $f^{-1}: Y \to X$ are continuous, then $f$ is called homeomorphism. You may have studied in modern algebra the notion of an isomorphism between algebraic objects such as groups or rings. An isomorphism is a bijective correspondence that preserves the algebraic structure involved. The analogous concept in topology i s that of homeomorphism; i t is a bijective correspondence that preserves the topological structure involved.
Hence, to prove our proposition, we need to prove $\beta_{a, b} < \infty$ a.s. for every $0 < a < b \in \R$. We only considered positive numbers since $(X_n)$ is positive and we have a homeomorphism between $\R^*$ and $[0, \infty]$ using $f(x) = x, x \geq 0$.
To this end, we need the Dubin’s inequalities.
Dubin’s inequalities. For every positive super martingale $(X_n)$ the upcrossing numbers are r.v.s satisfying
\[\Pr{\beta_{a, b} \geq k}{\sB_0} \leq \left(\frac{a}{b}\right)^k \min\left(\frac{X_0}{a}, 1\right),\]for every integer $k \geq 1$ and real numbers $a < b$. Therefore, r.v.s $\beta_{a, b}$ are a.s. finite.
Proof. First note that the set of stopping times {$\omega: \nu_k(\omega) =n$} belong to $\sB_n$ so $\nu_k$ are $\sB_n$-measurable. Now extending the master lemma, we can define the following supermartingale:
\[\begin{align*} Y_n & =1 \quad \mathrm{if}\: 0 \leq n < \nu_1, \\ &= \frac{X_n}{a} \quad \mathrm{if}\: \nu_1 \leq n < \nu_2, \\ &= \frac{b}{a} \cdot 1 \quad \mathrm{if}\: \nu_2 \leq n < \nu_3, \\ & \vdots \\ & = \left(\frac{b}{a}\right)^{k -1} \cdot \frac{X_n}{a} \quad \mathrm{if}\: \nu_{2k - 1} \leq n < \nu_{2k}, \\ & = \left(\frac{b}{a}\right)^{k} \quad \mathrm{if}\: n \geq \nu_{2k}. \end{align*}\]In fact we have that:
\[\begin{align*} 1 &\geq \frac{X_{\nu_1}}{a}, \\ \frac{X_{\nu_2}}{a} &\geq \frac{b}{a}, \\ &\vdots \\ \left(\frac{b}{a}\right)^{k - 1} \cdot \frac{X_{\nu_{2k}}}{a} &\geq \left(\frac{b}{a}\right)^k. \end{align*}\]By construction we have that $Y_0 = \min\left(1, \frac{X_0}{a}\right)$. Since {$Y_n$} is a supermartingale we have that $Y_0 \geq \Ev{Y_n}{\sB_0}$. Also on the set {$n \geq \nu_{2k}$}, $Y_n \geq \left(\frac{b}{a}\right)^{k}$, or equivalently $Y_n \geq \left(\frac{b}{a}\right)^{k} \cdot \I${$n \geq \nu_{2k}$}. Using all these facts, we have:
\[\begin{align*} \left(\frac{b}{a}\right)^k \Pr{\nu_{2k} \leq n}{\sB_0} \leq \min\left(1, \frac{X_0}{a}\right). \end{align*}\]All that remains is to let $n \to \infty$ and noting that {$\nu_{2k} < \infty$} $=$ {$\beta_{a,b} \geq k$}. $\square$ ___
The final part of the proof involves showing that the limit $X_\infty = \lim_{n \to \infty} X_n$ a.s. satisfies the following inequality
\(\Ev{X_\infty}{\sB_n} \leq X_n, \quad n \ \in \mathbb{N}.\) The inequality
\[\Ev{\inf_{m \geq n} X_m}{\sB_p} \leq \Ev{X_n}{\sB_p} \leq X_p\]is valid if $n > p$. Let $n$ tend to infinity, then
\[\Ev{X_\infty}{\sB_p} = \lim_{n \to \infty} \Ev{\inf_{m \geq n} X_m}{\sB_p} \leq X_p, \quad p \in \mathbb{N}.\]where we used the Lebesgue dominated convergence theorem to exchange the expectation and the limit because $\left(\inf_{m \geq n} X_m, \, n \in \mathbb{N}\right)$ is an increasing bounded convergent sequence in $n$. The proof is completed.
We had already showed that $\sup_n X_n < \infty$ on {$X_0 < \infty$}, which implies that $X_\infty < \infty$ on {$X_0 < \infty$}; applying this result to the supermartingale $(X_n, n \geq p)$ adapted to the sequence $(\sB_n, n \geq p)$, we find that $X_\infty$ a.s. on {$X_p < \infty$} for all $p \in \mathbb{N}$ which is another way of completing the last part of our proof.
The end! :)
