Stochastic Processes (2021/22)
Michaelmas term 1

4 Martingales

4.1 Definition and some examples

A martingale is a generalized version of a “fair game”.

A process \((M_n)_{n\ge0}\) is a martingale if

  1. for every \(n\ge0\) the expectation \({\mathsf{E}}M_n\) is finite, equivalently, \({\mathsf{E}}|M_n|<\infty\);

  2. for every \(n\ge0\) and all \(m_n\), \(m_{n-1}\), …, \(m_0\) we have \[\begin{equation} \label{eq:martingale-def} {\mathsf{E}}\bigl(M_{n+1}\mid M_n=m_n,\dots,M_0=m_0\bigr)=m_n\,. \end{equation}\]

For those familiar with the notion of conditioning on a random variable (see next section), \(\eqref{eq:martingale-def}\) can just be written as \({\mathsf{E}}\bigl(M_{n+1}\mid M_n,\dots,M_0\bigr)=M_n\).

We say that \((M_n)_{n\ge0}\) is a supermartingale2 if the equality in \(\eqref{eq:martingale-def}\) holds with \(\le\), ie., \({\mathsf{E}}\bigl(M_{n+1}\mid M_n,\dots,M_0\bigr)\le M_n\); and we say that \((M_n)_{n\ge0}\) is a submartingale, if \(\eqref{eq:martingale-def}\) holds with \(\ge\), ie., \({\mathsf{E}}\bigl(M_{n+1}\mid M_n,\dots,M_0\bigr)\ge M_n\).

Let \((\xi_n)_{n\ge1}\) be independent random variables  3 with \[{\mathsf{E}}\xi_n=0\] for all \(n\ge1\). Then the process \((M_n)_{n\ge0}\) defined via \[M_n\stackrel{{\sf def}}{=}M_0+\xi_1+\dots+\xi_n\] is a martingale as long as the random variable \(M_0\) is independent of \((\xi_n)_{n\ge1}\) and \({\mathsf{E}}|M_0|<\infty\). For example, we will often take \(M_0=0\) or some other deterministic constant.

Indeed, by the triangle inequality, \({\mathsf{E}}|M_n|\le{\mathsf{E}}|M_0|+\sum_{j=1}^n{\mathsf{E}}|\xi_j|<\infty\) for all \(n\ge0\); on the other hand, the independence property implies \[{\mathsf{E}}\bigl(M_{n+1}-M_n\mid M_n,\dots,M_0\bigr)\equiv{\mathsf{E}}\bigl(\xi_{n+1}\mid M_n,\dots,M_0\bigr)={\mathsf{E}}\xi_{n+1}=0\,.\]

Notice that if \({\mathsf{E}}\xi_n\ge0\) for all \(n\ge1\), then \((M_n)_{n\ge0}\) is a submartingale, whereas if \({\mathsf{E}}\xi_n\le0\) for all \(n\ge1\), then \((M_n)_{n\ge0}\) is a supermartingale. More generally, if \((\xi_n)_{n\ge1}\) are independent random variables with \({\mathsf{E}}|\xi_n|<\infty\) for all \(n\ge1\), then the process \(M_n=M_0+(\xi_1-{\mathsf{E}}\xi_1)+\dots+(\xi_n-{\mathsf{E}}\xi_n)\), \(n\ge0\), is a martingale.

Given a sequence \((\xi_k)_{k\ge1}\) of independent Bernoulli variables with common distribution \({\mathsf{P}}(\xi=1)=p\) and \({\mathsf{P}}(\xi=-1)=q=1-p\), define the generated random walk via \(X_n=\sum_{k=1}^n\xi_k\), \(n\ge0\). Show that the process \(M_n=(q/p)^{X_n}\) is a martingale.

Let \((Z_n)_{n\ge0}\) be a branching process with \({\mathsf{E}}Z_1=m<\infty\). We have \({\mathsf{E}}|Z_n|<\infty\) and \({\mathsf{E}}(Z_{n+1}\mid Z_n,\dots,Z_0)=mZ_n\) for all \(n\ge0\). In other words, the process \((Z_n)_{n\ge0}\) is a martingale, a submartingale or a supermartingale depending on whether \(m=1\), \(m>1\) or \(m<1\).
Notice also, that for every \(m\in(0,\infty)\) the process \((Z_n/m^n)_{n\ge0}\) is a martingale.

Let \(\rho\) be the extinction probability for a branching process \((Z_n)_{n\ge0}\); show that \(M_n=\rho^{Z_n}\) is a martingale.

We can also construct many examples of submartingales and supermartingales if we have a martingale and apply certain functions to it. This is a consequence of:

[Conditional Jensen’s inequality] Suppose that \(X\) is an integrable random variable on \((\Omega, \mathcal{F},{\mathsf{P}})\) and \(A\in \mathcal{F}\) with \({\mathsf{P}}(A)>0\).

Suppose that \(\varphi\) is a convex function such that \(\varphi(X)\) is also integrable. Then \[{\mathsf{E}}(\varphi(X)|A)\ge \varphi({\mathsf{E}}(X|A)).\]

This follows from the non-conditional version of Jensen’s inequality, and using the fact that \({\mathsf{P}}(\cdot|A)\) is a valid probability measure.

Let \((X_n)_{n\ge0}\) be a martingale. If, for some convex function \(f(\,\cdot\,)\) we have \({\mathsf{E}}\bigl(|f(X_n)|\bigr)<\infty\) for all \(n\ge0\), then the process \(f(X):=\bigl(f(X_n)\bigr)_{n\ge 0}\) is a submartingale.

Similarly, if for some concave \(f(\,\cdot\,)\) we have \({\mathsf{E}}\bigl(|f(X_n)|\bigr)<\infty\) for all \(n\ge0\), then the process \(f(X)\) is a supermartingale.

Both these claims follow immediately from the Jensen inequalities for conditional expectations.

4.2 A few remarks on conditional expectation

Recall the following basic definition:

Let \((\Omega,{\mathcal{F}},{\mathsf{P}})\) be a probability triple and let \(X\) and \(Y\) be random variables taking values \(x_1\), \(x_2\), …, \(x_m\) and \(y_1\), \(y_2\), …, \(y_n\) respectively. On the event \(\{Y=y_j\}\) one defines the conditional probability \[{\mathsf{P}}\bigl(X=x_i\mid Y=y_j\bigr)\stackrel{{\sf def}}{=}\frac{{\mathsf{P}}(X=x_i,Y=y_j)}{{\mathsf{P}}(Y=y_j)}\] and the conditional expectation: \({\mathsf{E}}\bigl(X\mid Y=y_j\bigr)\equiv\sum_ix_i{\mathsf{P}}\bigl(X=x_i\mid Y=y_j\bigr)\). Then the conditional expectation \(Z={\mathsf{E}}(X\,|\,Y)\) of \(X\) given \(Y\) is defined as follows: \[\text{if } Y(\omega)=y_j, \quad\text{ then }\quad Z(\omega)\stackrel{{\sf def}}{=}z_j\equiv{\mathsf{E}}\bigl(X\mid Y=y_j\bigr)\,.\] Notice that the value \(z_j\) is completely determined by \(y_j\); in other words, \(Z\) is a function of \(Y\), and as such, a random variable. Of course, if \(X\) and \(Y\) are independent, we have \(Z(\omega)\equiv{\mathsf{E}}X\).

If \({\mathcal{D}}=\{D_1,D_2,\dots,D_m\}\) is a finite partition  4 of \(\Omega\), the collection \({\mathcal{G}}=\sigma({\mathcal{D}})\) of all \(2^m\) possible subsets of \(\Omega\) constructed from blocks \(D_i\) is called the \(\sigma\)-field generated by \({\mathcal{D}}\). A random variable \(Y\) is measurable with respect to \({\mathcal{G}}\) if it is constant on every block \(D_i\) of the initial partition \({\mathcal{D}}\).

In [exmle:conditional-expectation] above we see that \(Z\) is constant on the events \(\{Y=y_j\}\); ie., \(Z\) is measurable w.r.t. the \(\sigma\)-field \(\sigma(Y)\equiv\sigma(\{Y=y_j\}_j)\). Moreover, for every \(G_j\equiv\{Y=y_j\}\), we have \[\begin{aligned} {\mathsf{E}}\bigl(Z\,\mathsf{1}_{G_j})&=z_j{\mathsf{P}}(Y=y_j)=\sum_ix_i{\mathsf{P}}\bigl(X=x_i\mid Y=y_j\bigr){\mathsf{P}}(Y=y_j) \\&=\sum_ix_i{\mathsf{P}}\bigl(X=x_i,Y=y_j\bigr)={\mathsf{E}}\bigl(X\mathsf{1}_{G_j})\,, \end{aligned}\] where the last equality follows from the observation that the random variable \(X\mathsf{1}_{G_j}\) equals \(x_i\) on every event \(G_j\cap\{X=x_i\}=\{Y=y_j\}\cap\{X=x_i\}\) and vanishes identically outside the set \(G_j\).

The last two properties can be used to define conditional expectation in general: Let \((\Omega, {\mathcal{F}},{\mathsf{P}})\) be a probability triple, let \({\mathcal{G}}\subseteq{\mathcal{F}}\) be a \(\sigma\)-field, and let \(X\) be a rv, \(X:\Omega\to{\mathbb{R}}\). The conditional expectation of \(X\) w.r.t. \({\mathcal{G}}\) is the unique random variable \(Z\) such that: \(Z\) is \({\mathcal{G}}\) measurable and for every set \(G\in{\mathcal{G}}\) we have \(\int_GZ{\mathsf{d}}{\mathsf{P}}\equiv{\mathsf{E}}\bigl(Z\mathsf{1}_G)={\mathsf{E}}\bigl(X\mathsf{1}_G)\equiv\int_GX{\mathsf{d}}{\mathsf{P}}\).

Notice that when \(\mathcal{G}=\sigma(D)\), the definition in the remark above coincides with Definition [def:ce_partition].

The following are the most important properties of conditional expectation:

Let \((\Omega, {\mathcal{F}},{\mathsf{P}})\) be a probability triple, and let \({\mathcal{G}}\subset{\mathcal{F}}\) a \(\sigma\)-field. Then for all random variables \(X\), \(X_1\), \(X_2\) and constants \(a_1\), \(a_2\), the following properties hold:

  1. If \(Z={\mathsf{E}}\bigl(X\,|\,{\mathcal{G}}\bigr)\) then \({\mathsf{E}}Z={\mathsf{E}}X\);

  2. If \(X\) is \({\mathcal{G}}\)-measurable, then \({\mathsf{E}}\bigl(X\,|\,{\mathcal{G}}\bigr)=X\);

  3. \({\mathsf{E}}\bigl(a_1X_1+a_2X_2\,|\,{\mathcal{G}}\bigr)=a_1{\mathsf{E}}\bigl(X_1\,|\,{\mathcal{G}}\bigr)+a_2{\mathsf{E}}\bigl(X_2\,|\,{\mathcal{G}}\bigr)\);

  4. If \(X\ge0\), then \({\mathsf{E}}\bigl(X\,|\,{\mathcal{G}}\bigr)\ge0\);

  5. [tower-property-conditional-expectation] If \({\mathcal{H}}\) and \({\mathcal{G}}\) are two \(\sigma\)-fields in \(\Omega\) such that \({\mathcal{H}}\subseteq{\mathcal{G}}\), then \[{\mathsf{E}}\Bigl[{\mathsf{E}}\bigl(X\,|\,{\mathcal{G}}\bigl)\,|\,{\mathcal{H}}\Bigr]={\mathsf{E}}\Bigl[{\mathsf{E}}\bigl(X\,|\,{\mathcal{H}}\bigl)\,|\,{\mathcal{G}}\Bigr]={\mathsf{E}}\bigl(X\,|\,{\mathcal{H}}\bigl)\,;\]

  6. If \(Z\) is \({\mathcal{G}}\)-measurable, then \({\mathsf{E}}\bigl[ZX\,|\,{\mathcal{G}}\bigr]=Z\,{\mathsf{E}}(X\,|\,{\mathcal{G}})\).

  7. If \(Z\) is independent of \({\mathcal{G}}\), then \({\mathsf{E}}\bigl[Z\,|\,{\mathcal{G}}\bigr]={\mathsf{E}}[Z]\).

If \((X_k)_{k\ge1}\) is a sequence of random variables, we can define the generated \(\sigma\)-fields \({\mathcal{F}}_1^X\), \({\mathcal{F}}_2^X\), …, via \[\begin{equation} \label{eq:Xn-sigma-field-def} {\mathcal{F}}_k^X\stackrel{{\sf def}}{=}\sigma(X_1,X_2,\dots,X_k)\,; \end{equation}\] here, the \(\sigma\)-field \({\mathcal{F}}_k^X\) stores all information about the process \((X_n)_{n\ge1}\) up to time \(k\). Observe that these \(\sigma\)-fields form a filtration \(\bigl({\mathcal{F}}_n^X\bigr)_{n\ge1}\) in the sense that \[\begin{equation} \label{eq:FXn-filtration} {\mathcal{F}}_1^X\subseteq{\mathcal{F}}_2^X\subseteq\ldots\subseteq{\mathcal{F}}_k^X\subseteq\ldots\,. \end{equation}\] The notion of filtration is very useful when working with martingales. Indeed, a process \((M_n)_{n\ge0}\) is a martingale if for all \(n\ge0\), \({\mathsf{E}}|M_n|<\infty\) and \({\mathsf{E}}\bigl(M_{n+1}\mid{\mathcal{F}}^M_n)=M_n\). We can also generalise the definition of a martingale:

\((M_n)_{n\ge 0}\) is a martingale with respect to a filtration \((\mathcal{F}_n)_{n\ge 0}\) if for all \(n\ge 0\), \({\mathsf{E}}(|M_n|)<\infty\) and \[{\mathsf{E}}(M_{n+1}|\mathcal{F}_n)=M_n.\]

The original definition of martingale is sometimes referred to as “being a martingale with respect to the natural filtration”.

We say that \(M\) is a martingale with respect to a sequence \((X_n)_{n\ge 0}\) if it is a martingale with respect to the filtration \((\mathcal{F}_n^X)_{n\ge 0}\).
We also notice that by repeatedly applying the tower property in claim [tower-property-conditional-expectation]) of [lem:conditional-expectation-properties] above to the sequence \(\eqref{eq:FXn-filtration}\), we get the following result:

If \((M_n)_{n\ge0}\) is a submartingale w.r.t. \((X_n)_{n\ge0}\), then for all integer \(n\ge k\ge0\), we have \({\mathsf{E}}M_n\ge{\mathsf{E}}M_k\).

Changing signs, we get the inequality \({\mathsf{E}}M_n\le{\mathsf{E}}M_k\) for supermartingales, and therefore the equality \({\mathsf{E}}M_n={\mathsf{E}}M_k\) for martingales.

4.3 Martingales and stopping times

Martingales are extremely useful in studying stopping times:

Let \((X_k)_{k\ge1}\) be a sequence of i.i.d. Bernoulli random variables with common distribution \({\mathsf{P}}(X_1=1)=p\), \({\mathsf{P}}(X_1=-1)=q=1-p\), where \(p\in(0,1)\), \(p\neq1/2\). For fixed \(k\in\{0,1,\dots,N\}\), define the random walk \((S_n)_{n\ge0}\) defined via \(S_n=k+\sum_{j=1}^nX_j\), \(n\ge0\), and consider the process \((Y_n)_{n\ge0}\), where \(Y_n=(q/p)^{S_n}\). Clearly, \(Y_n\) is an \(({\mathcal{F}}_n^X)_{n\ge0}\)-martingale, so that \({\mathsf{E}}(Y_n)={\mathsf{E}}(Y_0)=(q/p)^k\) for all \(n\ge0\), recall [lem:martingale-property].

Let \(T\) be the first time \(S_n\) hits \(0\) or \(N\). If an analogue of the above equality, \({\mathsf{E}}(Y_T)={\mathsf{E}}(Y_0)=(q/p)^k\) were true for (random) time \(T\), we could find the exit probability \({\mathsf{p}}_k={\mathsf{P}}(\, S \text{ hits $0$ before $N$ }|S_0=k)\) from the expression \[{\mathsf{E}}\bigl(Y_T\bigr)=(q/p)^0{\sf p}_k+(q/p)^N\bigl(1-{\sf p}_k\bigr)\,,\] thus obtaining \(\displaystyle{\sf p}_k=\bigl((q/p)^k-(q/p)^N\bigr)/\bigl(1-(q/p)^N\bigr)\).

The method used in the previous example relies on the assumption \({\mathsf{E}}(Y_T)={\mathsf{E}}(Y_0)\) and on the formula for \({\mathsf{E}}(Y_T)\) for a certain random variable \(T\). An important part of the theory of martingales is to study random variables \(T\) for which the above statements are true.  5

Let \(S_n=\sum_{k=1}^nX_k\) be the simple symmetric random walk in \(\{-K,\dots,K\}\), generated by a sequence of i.i.d. symmetric Bernoulli r.v. \(X_k\), where \({\mathsf{P}}(X_1=\pm1)=1/2\). Similarly to [exmle:gamblers-ruin] one can study the hitting time \(T\) of the boundary, \(T=\inf\bigl\{n\ge0:|S_n|=K\bigr\}\): namely, since  6 \[{\mathsf{E}}\bigl((S_T)^2\bigr)=K^2{\mathsf{P}}(S_T=K)+K^2{\mathsf{P}}(S_T=-K)=K^2\,,\] the same heuristics applied to the martingale \((Y_n)_{n\ge0}\), \(Y_n\stackrel{{\sf def}}{=}\bigl(S_n\bigr)^2-n\), leads to \(0={\mathsf{E}}(Y_0)={\mathsf{E}}(Y_T)={\mathsf{E}}\bigl((S_T)^2\bigr)-{\mathsf{E}}(T)\); i.e., it suggests that \({\mathsf{E}}(T)=K^2\).

One of our aims is to discuss general results that justify the above heuristics. To this end, we need to carefully define what we mean by a “stopping time”.

A variable \(T\) is a stopping time for a process \((X_n)_{n\ge0}\), if the occurrence/non-occurrence of the event \(\{T=n\}=\)“we stop at time \(n\) can be determined by looking at the values \(X_0\), \(X_1\), …, \(X_n\) of the process up to time \(n\). Equivalently, if we have \(\{T\le n\}\in{\mathcal{F}}^X_n\equiv\sigma(X_0,\dots,X_n)\) for every \(n\ge 0\).

Let \((X_n)_{n\ge0}\) be a stochastic process with values in \(S\), and let \(T\) be the hitting time of a set \(A\subset S\), namely, \(T=\min\bigl\{n\ge0:X_n\in A\bigr\}\). Then \(T\) is a stopping time for \((X_n)_{n\ge0}\).

Indeed, for every fixed \(n\ge0\), we have \(\{T>n\}\equiv\{X_0\notin A,X_1\notin A,\dots,X_n\notin A\}\); therefore, the event \(\{T>n\}\) and its complement \(\{T\le n\}\) both belong to \({\mathcal{F}}^X_n\).

By contrast, the last time that \(X\) visits \(A\), \(\tilde{T}=\max\{n\ge 0: X_n\in A\}\) is not generally a stopping time. Because we generally cannot tell just by looking at \(X_0,\dots, X_n\), whether the process will visit \(A\) after time \(n\) or not.

Let \(k\in{\mathbb{N}}\) be fixed, and let \(S\) and \(T\) be stopping times for a process \((X_n)_{n\ge0}\). Show that the following are stopping times:

  1. \(T\equiv k\),

  2. \(S\land T\equiv\min(S,T)\),

  3. \(S\lor T\equiv\max(S,T)\).

Let \(A\) be a set of states, and let \(T=T_A\) be the moment of the first visit to \(A\), ie., \(T=\min\{n\ge0:X_n\in A\}\). Consider \(S=S_A=\min\{n>T_A:X_n\in A\}\), the moment of the second visit to \(A\). Show that \(S_A\) is a stopping time for \((X_n)_{n\ge0}\). Is the variable \(L=\max\bigl\{n\ge0:X_n\in A\bigr\}\) a stopping time for \((X_n)_{n\ge0}\)?

Let \((M_n)_{n\ge0}\) be a submartingale w.r.t. \((X_n)_{n\ge0}\), and let \(T\) be a stopping time for \((X_n)_{n\ge0}\). Show that the process \((L_n^T)_{n\ge0}\), \[L_n^T\stackrel{{\sf def}}{=}M_{n\land T}\equiv\begin{cases}M_n\,,&\quad n\le T\,,\\ M_T\,,&\quad n>T\,,\end{cases}\] is a submartingale w.r.t. \((X_n)_{n\ge0}\). Deduce that if \((M_n)_{n\ge0}\) is a martingale w.r.t. \((X_n)_{n\ge0}\), then the stopped process \((L_n^T)_{n\ge0}\) is also a martingale w.r.t. \((X_n)_{n\ge0}\).

4.4 Optional stopping theorem

The optional stopping (or sampling) theorem (OST) tells us that, under quite general assumptions, whenever \(X_n\) is a martingale, then \(X_{T\land n}\) is a martingale for a stopping time \(T\). Such results are very useful in deriving inequalities and probabilities of various events associated with such stochastic processes.

[Optional Stopping Theorem][thm:OST] Let \((M_n)_{n\ge0}\) be a martingale w.r.t. \((X_n)_{n\ge0}\), and let \(T\) be a stopping time for \((X_n)_{n\ge0}\). Then the equality \[\begin{equation} \label{eq:OST} {\mathsf{E}}\bigl[M_T\bigr]={\mathsf{E}}[M_0] \end{equation}\] holds whenever one of the following conditions holds:

(OST-1)

the stopping time \(T\) is bounded, i.e., \({\mathsf{P}}(T\le N)=1\) for some \(N<\infty\);

(OST-2)

\({\mathsf{E}}T<\infty\), and the martingale \((M_n)_{n\ge0}\) has bounded increments, i.e., \(|M_{n+1}-M_n|\le K\) for all \(n\) and some constant \(K\);

(OST-3)

\({\mathsf{P}}(T<\infty)=1\), and the martingale \((M_n)_{n\ge0}\) is bounded, i.e., \(|M_n|\le K\) for all \(n\) and some constant \(K\).

If \(M_n\) records gambler’s fortune, by (OST-3), one cannot make money from a fair game, unless an unlimited amount of credit is available.

One can generalize (OST-3) by replacing the condition of boundedness, \(|M_n|\le K\), by that of uniform integrability for the martingale \((M_n)_{n\ge0}\): a sequence of random variables \((Y_n)_{n\ge0}\) is uniformly integrable if \[\begin{equation} \label{eq:UI-def} \lim_{K\to\infty}\sup_{n\ge0}{\mathsf{E}}\bigl(|Y_n|\mathsf{1}_{\{|Y_n|>K\}}\bigr)=0\,. \end{equation}\]

Let the SSRW \((S_n)_{n\ge0}\) be generated as in [exmle:SSRW-expected-exit-time-via-martingales]. Put \[H\stackrel{{\sf def}}{=}\inf\{n\ge0:S_n=1\}\,.\] Since this RW is recurrent,  7 we deduce that \({\mathsf{P}}(H<\infty)=1\). However, the (OST) does not apply, as \(0={\mathsf{E}}(S_0)\neq{\mathsf{E}}(S_H)\equiv1\). It is an instructive Exercise to check which conditions in each of the above (OST) are violated.

We now give the proof of [thm:OST].

Let us first assume that \(T\) satisfies (OST-1); that is, there is some \(N<\infty\) with \(0\le T\le N\). Then the decomposition \[M_T=\sum_{n=0}^NM_T\mathsf{1}_{\{T=n\}}=\sum_{n=0}^NM_n\mathsf{1}_{\{T=n\}}%=\sum_{n=0}^NM_n\bl(\ind(T>n-1)-\ind(T>n)\br) =M_0+\sum_{n=0}^{N-1}\bigl(M_{n+1}-M_n\bigr)\mathsf{1}_{\{T>n\}}\] holds. Now for each \(0\le n\le N-1\), (a) of Lemma [lem:conditional-expectation-properties] allows us to rewrite \({\mathsf{E}}\bigl( (M_{n+1}-M_n)\mathsf{1}_{\{T>n\}}\bigr)\) as \[{\mathsf{E}}\bigl({\mathsf{E}}\bigl((M_{n+1}-M_n)\mathsf{1}_{\{T>n\}}|\,{\mathcal{F}}^X_n\bigr)\bigr)={\mathsf{E}}\bigl(\mathsf{1}_{\{T>n\}}{\mathsf{E}}\bigl(M_{n+1}-M_n|\,{\mathcal{F}}^X_n\bigr)\bigr)\] where we have used that \(\{T>n\}\) is \(\mathcal{F}_n^X\)-measurable (by definition of a stopping time). Since \(M\) is a martingale, \({\mathsf{E}}\bigl(M_{n+1}-M_n|\,{\mathcal{F}}^X_n\bigr)=0\) and hence \({\mathsf{E}}\bigl( (M_{n+1}-M_n)\mathsf{1}_{\{T>n\}}\bigr)=0\) for each \(n\). Linearity of expectation then implies that \({\mathsf{E}}(M_T)={\mathsf{E}}(M_0)\) as required.

Now let us assume that \(T\) satisfies (OST-2). For fixed \(n\), \(\min(T,n)=T\wedge n\) is clearly a bounded stopping time (it is less than \(n\) with probability one), so by (OST-1) we have that \({\mathsf{E}}M_{T\land n}={\mathsf{E}}M_0\) for all \(n\ge0\). We now want to take a limit as \(n\to \infty\). For this we write \[|M_{T}-M_{T\land n}|=\bigl|\sum_{ k>n}(M_k-M_n)\mathsf{1}_{\{T=k\}}|=\bigl|\sum_{k>n}\bigl(M_k-M_{k-1}\bigr)\mathsf{1}_{\{T\ge k\}}\bigr|\] (in particular, noting that this gives \(|M_T|\le KT\) when \(n=0\) and so \(M_T\) is integrable). Taking expectations then gives that \({\mathsf{E}}(|M_T-M_{T\land n}|)\le K\sum_{k>n} {\mathsf{P}}(T\ge k)\) which tends to \(0\) as \(n\to \infty\). Hence \({\mathsf{E}}(M_T)=\lim_{n\to \infty}{\mathsf{E}}(M_{T\land n})=\lim_{n\to \infty} {\mathsf{E}}(M_0)={\mathsf{E}}(M_0)\).

Finally, we can assume (OST-3) and deduce the result in a similar way. The strategy is the same, but instead of writing \(M_T-M_{T\wedge n}\) as a telescoping sum, we note that \(|M_{T}-M_{T\wedge n}|\le 2K\mathsf{1}_{\{T>n\}}\) so that \[\bigl|{\mathsf{E}}M_T-{\mathsf{E}}M_{T\land n}\bigr|\le 2K{\mathsf{P}}(T>n)\to0\qquad\text{ as $n\to\infty$\,.}\]

For those who are familiar with the dominated convergence theorem. In (OST-3) it is sufficient to assume that \(\bigl|M_{n\land T}\bigr|\le K\) for all \(n\ge0\) and \({\mathsf{P}}(T<\infty)=1\). Indeed, by the dominated convergence theorem we deduce that then \(\bigl|M_T\bigr|\le K\), so that the argument in the proof above applies.

With suitable martingales (OST) gives very powerful results.

The following instructive example is due to D. Williams:

(ABRACADABRA) A monkey types symbols at random, one per unit time, on a typewriter having \(26\) keys. How long on average will it take him to produce the sequence ‘ABRACADABRA’?

Solution To compute the expected time, imagine a sequence of gamblers, each initially having £1, playing at a fair gambling casino. Gambler arriving just before time \(n\) (\(n\ge1\)) bets £1 on the event \(\{n\text{th letter will be {\sf A}}\}\). If he loses, he leaves. If he wins, he receives £26 all of which he bets on the event \(\{n+1\text{\,st letter will be {\sf B}}\}\). If he loses, he leaves. If he wins, he receives £\(26^2\) all of which he bets on the event \(\{n+2\text{\,nd letter will be {\sf R}}\}\) and so on through the whole ABRACADABRA sequence.

Let \(X_n\) denote the total winnings of the casino after the \(n\)th day. Since all bets are fair the process \((X_n)_{n\ge0}\) is a martingale with mean zero. Let \(N\) denote the time until the sequence ABRACADABRA appears. At time \(N\), gambler \(N-10\) would have won £\(26^{11}-1\), gambler \(N-3\) would have won £\(26^4-1\), gambler \(N\) would have won £\(26-1\) and all other \(N-3\) gamblers would have lost their initial fortune. Therefore, \[X_N=N-3-(26^{11}-1)-(26^4-1)-(26-1)=N-26^{11}-26^4-26\] and since (OST-2) can be applied (check this!), we deduce the \({\mathsf{E}}(X_N)={\mathsf{E}}(X_0)=0\), that is \[{\mathsf{E}}(N)=26+26^4+26^{11}\,.\]

Notice that the answer could also be obtained by considering a finite state Markov chain \(X_n\) on the state space of \(12\) strings representing the longest possible intersection of the tail of the typed text with the target word ABRACADABRA, ie., \(\bigl\{{\sf ABRACADABRA, ABRACADABR, \dots, ABRA, ABR, AB, A},\varnothing \bigr\}\), as there \(N\) is just the hitting time of the state ABRACADABRA from the initial condition \(X_0=\varnothing\).

Use an appropriate (OST) to carefully derive the probability \({\sf p}_k\) in [exmle:gamblers-ruin].

Use an appropriate (OST) to carefully derive the expectation \({\mathsf{E}}(T)\) in [exmle:SSRW-expected-exit-time-via-martingales].

Consider the simple symmetric random walk \((S_n)_{n\ge0}\), generated by a sequence of i.i.d. Bernoulli r.v. \(X_k\) with \({\mathsf{P}}(X_1=\pm1)=1/2\), ie., \(S_n=\sum_{k=1}^nX_k\). For integer \(a<0<b\), let \(T\) be the stopping time \(T=\inf\bigl\{n\ge0:S_n\notin(a,b)\bigr\}\equiv\inf\bigl\{n\ge0:S_n\in\{a,b\}\bigr\}\).

  1. Show that \((S_n)_{n\ge0}\) is a martingale and use an appropriate (OST) to find \({\mathsf{P}}(S_T=a)\) and \({\mathsf{P}}(S_T=b)\).

  2. Show that \((M_n)_{n\ge0}\) defined via \(M_n=(S_n)^2-n\) is a martingale w.r.t. the process \((S_n)_{n\ge0}\).

  3. For fixed integer \(K>0\), carefully apply an appropriate (OST) to \(M_n\) and prove that \({\mathsf{E}}(T\land K)={\mathsf{E}}\bigl[(S_{T\land K})^2\bigr]\).

  4. Deduce that \({\mathsf{E}}(T)=-ab\).

A coin showing heads with probability \(p\) is tossed repeatedly. Let \(w\) be a fixed sequence of outcomes such as ‘HTH’, and let \(N\) denote the number of (independent) tosses until the word \(w\) is observed. Using an appropriate martingale, find the expectation \({\mathsf{E}}N\) for each of the following sequences: ‘HH’, ‘HTH’, ‘HHTTHH’.

Let \((Y_n)_{n\ge0}\) be a supermartingale w.r.t. a sequence \((X_n)_{n\ge0}\) and let \(H_n\in{\mathcal{F}}_{n-1}^X=\sigma(X_0,\dots,X_{n-1})\) satisfy \(0\le H_n\le c_n\), where the constant \(c_n\) might depend on \(n\). Then the process \(W_n=W_0+\sum_{m=1}^nH_m\bigl(Y_m-Y_{m-1}\bigr)\), \(n\ge0\), is a supermartingale w.r.t. \((X_n)_{n\ge0}\).

Following the proof of the optional stopping theorem, we observe that since \((Y_n)_{n\ge0}\) is a supermartingale w.r.t. \((X_n)_{n\ge0}\), \[{\mathsf{E}}\bigl(H_m(Y_m-Y_{m-1})\bigr)={\mathsf{E}}\bigl[H_m\,{\mathsf{E}}(Y_m-Y_{m-1}\,|\,{\mathcal{F}}_{m-1})\bigr]\le0\,.\]

If \((Y_n)_{n\ge0}\) describes the stock price process, and \(H_m\) is the number of stocks held during the time \((m-1,m]\) (decided when the price \(Y_{m-1}\) is known), then \(W_n\) describes the fortune of an investor at time \(n\). As \((W_n)_{n\ge0}\) is a supermartingale w.r.t. \((X_n)_{n\ge0}\), we have \({\mathsf{E}}W_n\le{\mathsf{E}}W_0\) for all \(n\ge0\).

The famous “doubling martingale” corresponds to doubling the bet size until one wins, ie., to taking \(H_m=2^{m-1}\mathsf{1}_{\{T>m\}}\), where \(T\) is the first moment when the price goes up, ie., \(T=\min\{m>0:Y_m-Y_{m-1}=1\}\). Since the stopped process \((W_{n\land T})_{n\ge0}\) is a supermartingale, for all \(n\ge0\) we have \({\mathsf{E}}(W_{n\land T})\le{\mathsf{E}}(W_0)\), ie., on average, the doubling strategy does not produce money if one bets against a (super)martingale.

Let \((S_n)_{n\ge0}\) be a random walk generated by a sequence \((X_n)_{n\ge0}\) of i.i.d. steps with \({\mathsf{E}}|X|<\infty\) and \({\mathsf{E}}(X)=m\). If \(T\) is a stopping time for \((X_n)_{n\ge0}\) with \({\mathsf{E}}(T)<\infty\), then the optional stopping theorem implies that \(S_T\) is integrable and \[{\mathsf{E}}(S_T-S_0)=m\,{\mathsf{E}}(T).\]

To show this, first notice that \(S_n-n{\mathsf{E}}X=S_n-mn\) is a martingale and for every \(n\ge0\) the variable \(T\land n\) is a bounded stopping time. By (OST-1), we have \[\begin{equation} \label{eq:Wald-stopped-time} {\mathsf{E}}(S_0)={\mathsf{E}}\bigl(S_{n\land T}-m(n\land T)\bigr). \end{equation}\] This rearranges to \[{\mathsf{E}}\bigl(S_{n\land T}-S_0\bigr)=m{\mathsf{E}}(n\land T)\] for every \(n\). Now, the RHS converges to \({\mathsf{E}}(T)\) as \(n\to \infty\) since \(|{\mathsf{E}}(T)-{\mathsf{E}}(n\land T)|\le {\mathsf{E}}(T\mathsf{1}_{\{T>n\}})=\sum_{k>n} k{\mathsf{P}}(T=k)\), where the tail sums on the right go to zero as \(n\to \infty\) by the assumption that \(T\) is integrable. Next, by writing \(|S_T|=\sum_{k\ge 0} |X_k|\mathsf{1}_{\{T\ge k\}}\) as a telescoping sum, where \({\mathsf{E}}(|X_k|\mathsf{1}_{\{T\ge k\}})={\mathsf{P}}(T\ge k){\mathsf{E}}(|X_1|)\) and \(\sum_{k} {\mathsf{P}}(T\ge k)<\infty\), we see that \(|S_T|\) is integrable. 8 Similarly, we can bound \({\mathsf{E}}(|S_T-S_{T\wedge n}|)\le {\mathsf{E}}(|X_1|)\sum_{k>n} {\mathsf{P}}(T\ge k)\) which tends to \(0\) as \(n\to \infty\). This implies that \({\mathsf{E}}(S_{T\land n})\to {\mathsf{E}}(S_T)\) as \(n\to \infty\), and combining the above completes the argument.

4.5 Martingale convergence theorem

This subsection is optional and will not be examined.
The following example has a number of important applications.

[Pólya’s urn][exmle:Polya-urn] An urns contains one green and one red ball. At every step a ball is selected at random, and then replaced together with another ball of the same colour. Let \(X_n\) be the number of green balls after \(n\)th draw, \(X_0=1\). Then the fraction \(M_n=X_n/(n+2)\) of green balls is a martingale w.r.t. the filtration \(({\mathcal{F}}^X_n)_{n\ge0}\).

Indeed, as \(|M_n|\le1\) we have \({\mathsf{E}}|M_n|\le1\) for all \(n\ge0\), and since \[{\mathsf{P}}(X_{n+1}=k+1\mid X_n=k)=\frac{k}{n+2}\,,\qquad{\mathsf{P}}(X_{n+1}=k\mid X_n=k)=1-\frac{k}{n+2}\,,\] we get \({\mathsf{E}}\bigl(X_{n+1}\mid{\mathcal{F}}_n^X\bigr)=\frac{n+3}{n+2}\,X_n\), equivalently, \({\mathsf{E}}(M_{n+1}\mid{\mathcal{F}}^X_n)=M_n\).

Show that \({\mathsf{P}}(M_n=\frac{k}{n+2})=\frac{1}{n+1}\) for \(1\le k\le n+1\), ie., \(M_n\) is uniformly distributed in \(\bigl\{\tfrac1{n+2},\tfrac2{n+2},\ldots,\tfrac{n+1}{n+2}\bigr\}\).

[exse:Polya-urn-one] suggests that in the limit \(n\to\infty\) the distribution of \(M_n\) becomes uniform in \((0,1)\):

Show that \(\lim\limits_{n\to\infty}{\mathsf{P}}\bigl(M_n<x\bigr)=x\) for every \(x\in(0,1)\).

In view of [exse:Polya-urn-one], a natural question is: does the proportion \(M_n\) of green balls fluctuate between \(0\) and \(1\) infinitely often or does it eventually settle down to a particular value? The following example shows that the latter is true. Our argument is based upon the following observation: if a real sequence \(y_n\) does not converge, for some real \(a\), \(b\) with \(-\infty<a<b<\infty\) the sequence \(y_n\) must go from the region below \(a\) to the region above \(b\) (and back) infinitely often.

For fixed \(n\ge0\) let \(M_n<a\in(0,1)\), and let \(N=\min\{k>n: M_n>b\}\) for some \(b\in(a,1)\). Since \(N_m=N\land m\) is a bounded stopping time, by (OST-1) we have \({\mathsf{E}}{M_{N_m}}={\mathsf{E}}{M_n}<a\) if only \(m>n\). On the other hand, \[{\mathsf{E}}{M_{N_m}}\ge{\mathsf{E}}\bigl(M_{N_m}\mathsf{1}_{N\le m}\bigr)\equiv{\mathsf{E}}\bigl(M_{N}\mathsf{1}_{N\le m}\bigr)>b\,{\mathsf{P}}(N\le m)\,.\] In other words, \({\mathsf{P}}(N\le m)<\frac{a}{b}\) and consequently \({\mathsf{P}}(N<\infty)\le\frac{a}{b}<1\), ie., the fraction \(M_n\) of green balls ever gets above level \(b\) with probability at most \(\frac{a}{b}\). Suppose that at certain moment \(N\in(n,\infty)\) the fraction of green balls became bigger than \(b\). Then a similar argument shows that with probability at most \((1-b)/(1-a)\) the value \(M_n\) becomes smaller than \(a\) at a later moment.

Put \(S_0=\min\{n\ge0:M_n<a\}\), and then, inductively, for \(k\ge0\), \[\begin{equation} \label{eq:Polya-urn-ST-times} T_k=\min\bigl\{n>S_k:M_n>b\bigr\}\,,\qquad S_{k+1}=\min\bigl\{n>T_k:M_n<a\}\,. \end{equation}\] The argument above implies that \[\begin{equation} \label{eq:probability-of-upcrossings} {\mathsf{P}}(S_k<\infty)\le\prod_{j=1}^k\Bigl({\mathsf{P}}(T_{j-1}<\infty\mid S_{j-1}<\infty){\mathsf{P}}(S_j<\infty\mid T_{j-1}<\infty)\Bigr) \end{equation}\] with the RHS bounded above by \(\bigl(\frac{a}{b}\bigr)^k\bigl(\frac{1-b}{1-a}\bigr)^k\to0\) as \(k\to\infty\). As a result, the probability of infinitely many crossing (ie., \(S_k<\infty\) for all \(k\ge0\)) vanishes.

Clearly, the argument above applies to all strips \((a,b)\subset(0,1)\) with rational endpoints. Thus, with probability one, 9 trajectories of \(M_n\) eventually converge to a particular value. 10

Write \(U_{(a,b)}\) for the total number of upcrossings of the strip \((a,b)\) by the process \((M_n)_{n\ge0}\). By using the approach of [exmle:Polya-urn-convergence] and noticing that \(\bigl\{U_{(a,b)}\ge m\bigr\}\subset\bigl\{S_{m-1}<\infty\bigr\}\) or otherwise, show that \({\mathsf{E}}U_{(a,b)}<\infty\).

The argument in [exmle:Polya-urn-convergence] also works in general. Let \((M_n)_{n\ge0}\) be a martingale w.r.t. filtration \(({\mathcal{F}}^X_n)_{n\ge0}\). For real \(a\), \(b\) with \(-\infty<a<b<\infty\) let \(U_{(a,b)}\) be the total number of upcrossings of the strip \((a,b)\). The following result (or some of its variants) is often referred to as Doob’s Upcrossing Lemma:

Let the martingale \((M_n)_{n\ge0}\) have uniformly bounded expectations, ie., for some constant \(K\) and all \(n\ge0\), \({\mathsf{E}}|M_n|<K<\infty\). If \(U_{(a,b)}\) is the number of upcrossings of a strip \((a,b)\), then \({\mathsf{E}}U_{(a,b)}<\infty\).

With stopping times as in \(\eqref{eq:Polya-urn-ST-times}\), put \(H_n=1\) if \(S_m<n\le T_{m}\) and put \(H_n=0\) otherwise. Then the process \(W_n=\sum_{k=1}^nH_k(M_k-M_{k-1})\) is a martingale w.r.t. \((M_n)_{n\ge0}\), cf. [lem:betting-martingale-transform]. It is easy to check that \(W_n\ge(b-a)\,U_n-|M_n-a|\) (draw the picture!), where \(U_n=\max\{m\ge0:T_m\le n\}\) is the number of upcrossings of the strip \((a,b)\) up to time \(n\). As a result  \[0={\mathsf{E}}W_0={\mathsf{E}}W_n\ge(b-a)\,{\mathsf{E}}U_n-{\mathsf{E}}|M_n-a|\ge(b-a)\,{\mathsf{E}}U_n-(K+|a|)\,,\] so that \({\mathsf{E}}U_n\le(K+|a|)/(b-a)\) for all \(n\ge0\), and thus \({\mathsf{E}}U_{(a,b)}<\infty\).

Let \((M_n)_{n\ge0}\) be a martingale as in [lem:Doob-upcrossings]. Then there exists a random variable \(M_\infty\) such that \(M_n\to M_\infty\) with probability one.

If \(M_n\) does not converge, for some rational \(a\), \(b\) with \(-\infty<a<b<\infty\) we must have \(U_{(a,b)}=\infty\). However, by [lem:Doob-upcrossings], \({\mathsf{E}}U_{(a,b)}<\infty\) implying that \({\mathsf{P}}(U_{(a,b)}=\infty)=0\). As the number of such pairs \((a,b)\) is countable, the result follows.

Let \((X_k)_{k\ge1}\) be independent variables with \[{\mathsf{P}}\Bigl(X=\frac32\Bigr)={\mathsf{P}}\Bigl(X=\frac12\Bigr)=\frac12\,.\] Put \(M_n=X_1\cdot\ldots\cdot X_n\) with \(M_0=1\). Show that \(M_n\) is an \(({\mathcal{F}}^X_n)_{n\ge0}\) martingale. Deduce that \(M_n\to M_\infty\) with probability one. Can you compute \({\mathsf{E}}(M_\infty)\)?

4.6 Additional problems

Let \((\eta_n)_{n\ge1}\) be independent positive random variables with \({\mathsf{E}}\eta_n=1\) for all \(n\ge1\). If a random variable \(M_0>0\) is independent of \((\eta_n)_{n\ge0}\) and \({\mathsf{E}}M_0<\infty\), then the process \((M_n)_{n\ge0}\) defined via \(M_n=M_0\prod_{j=1}^n\eta_j\) is a martingale w.r.t. \((\eta_n)_{n\ge1}\).

1ex Interpreting \(\eta_n-1\) as the (fractional) change in the value of a stock during the \(n\)th time interval, the martingale \((M_n)_{n\ge0}\) can be used to model stock prices. Two often used examples are:
Discrete Black-Sholes model: take \(\eta_j=e^{\zeta_j}\), where \(\zeta_j\) is Gaussian, \(\zeta_j\sim{\mathcal{N}}(\mu,\sigma^2)\);
Binomial model: take \(\eta_j=(1+a)e^{-r}\) and \(\eta_j=(1+a)^{-1}e^{-r}\) with probabilities \(p\) and \(1-p\) respectively.

Let \((S_n)_{n\ge0}\) be the random walk from [exmle:SSRW-expected-exit-time-via-martingales]. Find constants \(a\), \(b\), \(c\) such that the process \((S_n)^4+an(S_n)^2+bn^2+cn\) is an \((X_n)_{n\ge0}\)-martingale. Use the heuristc in [exmle:SSRW-expected-exit-time-via-martingales] to predict the value of the second moment \({\mathsf{E}}(T^2)\) of the exit time \(T\).

A standard symmetric dice is tossed repeatedly. Let \(N\) be the number of (independent) tosses until a fixed pattern is observed. Using an appropriate martingale, find \({\mathsf{E}}N\) for the sequences ‘123456’ and ‘123321’.

Suppose that the process in [exmle:Polya-urn] is modified as follows: for a fixed integer \(c>1\), every time a random ball is selected, it is replaced together with other \(c\) balls of the same colour. If, as before, \(X_n\) denotes the total number of green balls after \(n\) draws, show that the the fraction \(M_n=\frac{X_n}{2+nc}\) of green balls forms a martingale w.r.t. \(({\mathcal{F}}^X_n)_{n\ge0}\).

Find the large-\(n\) limit of the distribution of the martingale \((M_n)_{n\ge0}\) from [exse:Polya-urn-general].

Let \((X_n)_{n\ge0}\) be a birth-and-death process in \({\mathcal{S}}=\{0,1,\dots\}\), ie., a Markov chain in \({\mathcal{S}}\) with transition probabilities \(p_{00}=r_0\), \(p_{01}=p_0\), and \(p_{m,m-1}=q_m\), \(p_{m,m}=r_m\), \(p_{m,m+1}=p_m\) for \(m>0\), while \(p_{m,k}=0\) for all other pairs \((m,k)\in{\mathcal{S}}^2\). Let \(X_0=x\), and for \(y\ge0\) denote \(T_y\stackrel{{\sf def}}{=}\min\{n\ge0:X_n=y\}\).

  1. Show that the process \(\bigl(\varphi(X_n)\bigr)_{n\ge0}\) with \(\varphi(z)\stackrel{{\sf def}}{=}\sum\limits_{y=1}^z\prod\limits_{x=1}^{y-1}\dfrac{q_x}{p_x}\) is a martingale.

  2. Show that for all \(0\le a<X_0=x<b\) we have \[{\mathsf{P}}\bigl(T_b<T_a\bigr)=\bigl(\varphi(x)-\varphi(a)\bigr)/\bigl(\varphi(b)-\varphi(a)\bigr)\,.\] Deduce that state \(0\) is recurrent iff \(\varphi(b)\to\infty\) as \(b\to\infty\).

  3. Now suppose that \(p_m\equiv p\), \(q_m\equiv q=1-p\), and \(r_m=0\) for \(m>0\), whereas \(p_0=p\) and \(r_0=q\). Show that in this case the result in part b) above becomes \[{\mathsf{P}}\bigl(T_b<T_a\bigr)=\bigl((q/p)^a-(q/p)^x\bigr)/\bigl((q/p)^a-(q/p)^b\bigr)\,.\]

  4. Find \({\mathsf{P}}\bigl(T_b<T_a\bigr)\) if in the setup of part c) one has \(p=q=1/2\).

Let \((\xi_k)_{k\ge1}\) be i.i.d. random variables with \({\mathsf{P}}(\xi=1)=p<\tfrac12\), \({\mathsf{P}}(\xi=-1)=q=1-p\), and \({\mathsf{E}}\xi>0\). Let \((S_n)_{n\ge0}\) be the generated random walk, \(S_n=x+\xi_1+\dots+\xi_n\), and let \(T_0=\min\{n\ge0:S_n=0\}\) be the hitting time of \(0\). Deduce that for all \(x>0\), \({\mathsf{P}}(T_0<\infty)=(q/p)^x\). Compare this to the result of [exmle:SRW-hitting-time].

Let \((Z_n)_{n\ge0}\) be a homogeneous branching process with \(Z_0=1\), \(m={\mathsf{E}}Z_1>0\) and finite variance \(\sigma^2={\mathsf{Var}}(Z_1)\). Show that \(M_n=Z_n/m^n\) is a martingale.

  1. Let \(m>1\). By using [exse:branching-process-two-moments] or otherwise show that \({\mathsf{E}}(M_n)\) is uniformly bounded. Deduce that \(M_n\to M_\infty\) almost surely. What can you say about \({\mathsf{E}}M_\infty\)?

  2. What happens if \(m\le1\)? Compute \({\mathsf{E}}(M_\infty)\).

Hint: Recall [exse:uniform-integrability].

Let \((X_n)_{n\ge0}\) be a sequence of i.i.d. Bernoulli random variables with \({\mathsf{P}}(X=1)=p\) and \({\mathsf{P}}(X=-1)=1-p=q\). Let \((S_n)_{n\ge0}\) be the generated random walk with \(S_0=x>0\), and let \(T=\min\{n\ge0:S_n=0\}\) be the hitting time of the origin. [exmle:SRW-hitting-time] suggests that \({\mathsf{E}}(T)=x/(q-p)<\infty\) for \(q>p\).

  1. Use \(\eqref{eq:Wald-stopped-time}\) to deduce that \((q-p){\mathsf{E}}(n\land T)={\mathsf{E}}\bigl(S_0-S_{n\land T}\bigr)\le{\mathsf{E}}(S_0)=x\); then take \(n\to\infty\) to show that \({\mathsf{E}}(T)<\infty\);

  2. Use the Wald equation to deduce that indeed \({\mathsf{E}}(T)=\frac{x}{q-p}\). Can you give a heuristic explanation of this result?

  3. Argue, without using the Wald equation, that \({\mathsf{E}}(T)=cx\) for some constant \(c\).

  4. Use the Wald equation and an argument by contradiction to show that if \(p\ge q\), then \({\mathsf{E}}(T)=\infty\) for all \(x>0\).

Let a variable \(Y\) satisfy \({\mathsf{E}}(Y^2)<\infty\). Show that \({\mathsf{E}}|Y|<\infty\). Hint Notice that \({\mathsf{Var}}(|Y|)\ge0\).

Let \((X_n)_{n\ge0}\) be an irreducible Markov chain in \({\mathcal{S}}=\{0,1,\dots\}\) with bounded jumps, and let a function \(\varphi:{\mathcal{S}}\to{\mathbb{R}}^+\) satisfy \(\varphi(x)\to\infty\) as \(x\to\infty\). Let \(K\ge0\) be such that \[{\mathsf{E}}_x\varphi(X_1)\stackrel{{\sf def}}{=}{\mathsf{E}}\bigl[\varphi(X_1)|X_0=x\bigr]\le\varphi(x)\] for all \(x\ge K\).

  1. If the function \(\varphi(x)\) is monotone, show that the set of states \(\{0,1,\dots,K\}\), and thus the whole space \({\mathcal{S}}\) is recurrent for \((X_n)_{n\ge0}\).
    Hint If \(H_K=\min\bigl\{n\ge0:0\le X_n\le K\}\), show that \(\varphi(X_{n\land H_K})\) is a supermartingale. Deduce that if \(T_M=\min\bigl\{n\ge0:X_n\ge M\}\), then \(\varphi(x)\ge\varphi(M){\mathsf{P}}(T_M<H_K)\).

  2. Argue that the result holds for \(\varphi(x)\ge0\) not necessarily monotone, but only satisfying \(\varphi(x)\to\infty\) as \(x\to\infty\).
    Hint With \(T_M\) as above, show that \(\varphi_M^*\stackrel{{\sf def}}{=}\min\{\varphi(x):x\ge M\}\to\infty\) as \(M\to\infty\).

Let \((X_k)_{k\ge1}\) be independent variables with \({\mathsf{P}}(X=\pm1)=\frac12\). Show that the process \[M_n=\sum_{k=1}^n\frac1k\,X_k\] is a martingale w.r.t. \(({\mathcal{F}}^X_n)_{n\ge0}\) and that \({\mathsf{E}}\bigl[(M_n)^2\bigr]<K<\infty\) for some constant \(K\) and all \(n\ge0\). By using [exse:uniform-integrability] or otherwise, deduce that with probability one, \(M_n\to M_\infty\) for some random variable \(M_\infty\). In other words, the random sign harmonic series converges with probability one.

Thinking of a population of \(N\) haploid individuals who have one copy of each of their chromosomes, consider a fixed population of \(N\) genes that can be of two types \(A\) or \(a\). In the simplest version of this model the population at time \(n+1\) is obtained by sampling with replacement from the population at time \(n\). If we let \(X_n\) to be the number of \(A\) alleles at time \(n\), then \(X_n\) is a Markov chain with transition probability \[p_{ij}=\binom{N}{j}\Bigl(\frac{i}{N}\Bigr)^j\Bigl(1-\frac{i}{N}\Bigr)^{N-j}\,.\] Starting from \(i\) of the \(A\) alleles and \(N-i\) of \(a\) alleles, what is the probability that the population fixates in the all \(A\) state? Hint You can use the heuristics of [exmle:gamblers-ruin] but need to justify your computation!

Let \((X_n)_{n\ge0}\) be a Markov chain with a (countable) state space \({\mathcal{S}}\) and the transition matrix \({\mathbf{P}}\), and let \(h(x,n)\) be a function of the state \(x\) and time \(n\) such that  11 \[h(x,n)=\sum_yp_{xy}h(y,n+1)\,.\] Show that \((M_n)_{n\ge0}\) with \(M_n=h(X_n,n)\) is a martingale w.r.t. \((X_n)_{n\ge0}\).

Let \((X_n)_{n\ge0}\) be a Markov chain with a (countable) state space \({\mathcal{S}}\) and the transition matrix \({\mathbf{P}}\). If \(\boldsymbol{\psi}\) is a right eigenvector of \({\mathbf{P}}\) corresponding to the eigenvalue \(\lambda>0\), ie., \({\mathbf{P}}\boldsymbol{\psi}=\lambda\boldsymbol{\psi}\), show that the process \(M_n=\lambda^{-n}\psi\bigl(X_n\bigr)\) is a martingale w.r.t. \((X_n)_{n\ge0}\).


  1. current version by Ellen Powell (ellen.g.powell@durham.ac.uk); based on previous notes by Nicholas Georgiou and Ostap Hyrniv↩︎

  2. If \(M_n\) traces your fortune, then “there is nothing super about a supermartingale”.↩︎

  3. Notice that we do not assume that all \(\xi_n\) have the same distribution!↩︎

  4. [general-measurability] In the general countable setting, if \({\mathcal{D}}=\{D_1,D_2,\dots\}\) forms a denumerable (ie., infinite countable) partition of \(\Omega\), then the generated \(\sigma\)-field \({\mathcal{G}}=\sigma({\mathcal{D}})\) consists of all possible subsets of \(\Omega\) which are obtained by taking countable unions of blocks of \({\mathcal{D}}\). Similarly, a variable \(Y\) is measurable w.r.t. \({\mathcal{G}}\), if for every real \(y\) the event \(\{\omega:Y(\omega)\le y\}\) belongs to \({\mathcal{G}}\) (equivalently, can be expressed as a countable union of blocks of \({\mathcal{D}}\).↩︎

  5. Notice that the gambler’s ruin problem can be solved by using the methods of finite Markov chains, so we indeed know that the result above is correct.↩︎

  6. the equality is correct, because \({\mathsf{P}}(T<\infty)=1\) here!↩︎

  7. alternatively, recall the result in [exmle:SRW-hitting-time].↩︎

  8. Here we are really using the monotone convergence theorem: if \(0\le Z_n\le Z\) for every \(n\) and \(Z_n\uparrow Z\) as \(n\to \infty\), then \({\mathsf{E}}(Z_n)\to{\mathsf{E}}(Z)\) as \(n\to \infty\).↩︎

  9. If \(M_n\) does not converge, it must cross at least one of countably many strips \((a,b)\) with rational points infinitely many times.↩︎

  10. which is random and depends on the trajectory↩︎

  11. This result is useful, eg., if \(h(x,n)=x^2-cn\) or \(h(x,n)=\exp\{x-cn\}\) for a suitable \(c\).↩︎