Martingales
Posted: 2016-04-08 , Modified: 2016-04-08
Tags: none
Posted: 2016-04-08 , Modified: 2016-04-08
Tags: none
Durrett Ch. 5
See notebook 13.93-95 for exercises.
Define conditional expectation and show it is well defined.
Let \(X\) be a random variable on \((\Om, \mathcal F_0, P)\) with \(\E|X|<\iy\). The conditional expectation of \(X\) given \(\mathcal F\) is any (the unique) random variable \(Y\) such that
We show uniqueness and existence. First note that any \(Y\) satisfying the above is integrable: show \(\E |Y|\le \E |X|\) by partitioning based on \(\sgn(Y)\).
Existence is based on the Radon-Nikodym Theorem. Let \(\mu,\nu\) be \(\si\)-finite measures on \((\Om, \mathcal F)\). If \(\nu\ll \mu\) (\(\nu\) is absolutely continuous with respect to \(\mu\), i.e., \(\mu(A)=0\implies \nu(A)=0\)), then there is a function \(f\in \mathcal F\) such that for all \(A\in \mathcal F\), \[\int_A f\,d\mu = \nu(A).\] Write \(f=\dd{v}{\mu}\), and call it the Radon-Nikodym derivative.
For \(X\ge 0\), we have \[ \E(X|\mathcal F) = \dd{\nu}{\mu}\] where \(\mu=P\) and \(\nu(A) = \int_A X\,dP\). For general \(X\), write \(X=X^+-X^-\).
The general strategy for finding the conditional expectation is to “guess and verify”.
(Relate to conditional probability) If \(\mathcal F = \si(\Om_1,\ldots)\) and each \(\Om_i\) has positive probability, then \(\E[X|\mathcal F] = \fc{\E[X;\Om_i]}{\Pj(\Om_i)}\) on \(\Om_i\).
Generalize conditional probability by letting \[\begin{align} \Pj(A|\mathcal G) &= \E[1_A|\mathcal G]\\ \Pj(A|B) &= \fc{P(A\cap B)}{P(B)}\\ \E[X|Y] &= \E[X|\si(Y)]. \end{align}\]If the probability density of \((X,Y)\) is given by \(f(x,y)\), then \[\E[g(X)|Y] = h(Y)\] where \[h(y) = \fc{\int g(x) f(x,y)\dx}{\int f(x,y)\dx}.\] (5 and 6 omitted.)
(Proofs straightforward from definition, omitted.)
A filtration \(\mathcal F\) is an increasing sequence of \(\si\)-fields.
\(X_n\) is adapted to \(\mathcal F_n\) if for all \(n\), \(X_n\in \mathcal F_n\).
\(X_n\) is a martingale if 1. \(\E|X_n|<\iy\) 2. \(X_n\) is adapted to \(\mathcal F_n\) 3. \(\E[X_{n+1}|\mathcal F_n]=X_n\) for all \(n\).
If 3 holds with \(\le,\ge\), \(X_n\) is a supermartingale, submartingale. (Note a supermartingale goes down and a submartingle goes up.)
If \(X_n\) is a supermartingale and \(n>m\), then \(\E[X_n|\mathcal F_m]\le X_m\).
Proof. Induct and use monotonicity, \[\begin{align} \E[X_{m+k}|\mathcal F_m] &= \E[\E[X_{m+k}|\mathcal F_{n+k-1}]|\mathcal F_m]\\ &\le \E[X_{m+k-1}|\mathcal F_m] \end{align}\]If \(X_n\) is a martingale, \(\ph\) is convex, and \(\E[\ph(X_n)]<\iy\), then \(\ph(X_n)\) is a submartingale.
Proof. Use Jensen to bring the \(\ph\) outside in \(\E[\ph(X_{n+1})|\mathcal F_n]\). We need \(X_n\) to be a martingale so the inside becomes \(X_n\).Cor. If \(X_n\) is sub, then \((X_n-a)^+\) is sub.
If \(X_n\) is super, \(X_n\wedge a\) is super.Define \((H\cdot X)_n = \sum_{m=1}^n H_m (X_m-X_{m-1})\). \(H_n\) is predictable if \(H_n\in \mathcal F_{n-1}\) (predictable from information at time \(n-1\)).
(You can’t beat an unfavorable game.) If \(X_n\) is super, \(H_n\ge 0\) is predictable, and each \(H_n\) is bounded, then \((H\cdot X)_n\) is super.
Proof. Expand \(\E[(H\cdot X)_{n+1}|\mathcal F_{n}]\) and take the \(H_{n+1}\) out.Corollary: If \(N\) is a stopping time and \(X_n\) is super, then \(X_{N\wedge n}\) is super.
Proof. Let \(H_n = 1_{N\ge n}\).number of upcrossings by time \(n\).
If \(X_m\) is sub, then \[(b-a) \E U_n\le E(X_n-a)^+ - \E (X_0-a)^+.\]
Motivation. To prove the martingale convergence theorem (10). A martingale doesn’t converge if it has infinitely many upcrossings of some \([a,b]\). Show this happens with probability 0 and union-bound.
Proof. Consider a betting system trying to take advantage of upcrossings, \[H_m=\begin{cases} 1, &\text{if }N_{2k-1}<m\le N_{2k}\text{for some }k\\ 0,&\text{otherwise}\end{cases}.\] Let \(Y_n = a+(X_n-a)^+\). We have \[ \E(Y_n - Y_0) \ge \E(H\cdot Y)_n \ge (b-a) \E U_n.\]Martingale convergence theorem: If \(X_n\) is sub, \(\sup \E X_n^+<\iy\), then as \(n\to \iy\), \(X_n\) converges a.s. to a limit \(X\) with \(\E|X|<\iy\).
Proof. The expected number of upcrossings is finite. So \(\limsup X_n=\liminf X_n\) a.s.Doob’s decomposition: Any submartingale \(X_n\) can be written uniquely as \(X_n=M_n+A_n\), \(M_n\) a martingale and \(A_n\) predictable with \(A_0=0\).
Proof. Solve for what \(A_n\) must be: \(A_n = A_{n-1} + \E[X_n|\mathcal F_{n-1}] - X_{n-1}\).
For \(i\le j\), \(\si(X_i)\subeq \mathcal F_i \subeq \mathcal F_j\) so \(\si(X_i)\subeq \mathcal F_j\). Hence \(\mathcal G_j\subeq \mathcal F_j\).
Using definition and then taking \(\E[\bullet|\mathcal F_n]\), \[\begin{align} \E[X_{n+1}|\mathcal G_n]&=X_n\\ \E[X_{n+1}|\mathcal F_n] = \E[\E[X_{n+1}|\mathcal G_n]|\mathcal F_n]&=X_n. \end{align}\]\(A_n = A_{n-1} + \E[1_{B_n}|\mathcal F_{n-1}]\).
Then \(\Pj(C\cup D)=1\).
Proof. Show that when \(\liminf X_n>-\iy\) then \(\lim X_n\) exists and is finite. By symmetry, also \(\limsup X_n<\iy \implies \lim X_n\) exists. To show existence, apply 2.11 to \(X_{n\wedge \inf\set{n}{X_n\le -K}} + K + M\).Martingale Borel-Cantelli: Let \(\mathcal F_n\) be a filtration with \(\mathcal F_0=\{\phi,\Om\}\) and \(A_n\in \mathcal F_n\). Then \[\{A_n\text{ i.o.}\} = \bc{\sumo n{\iy} \Pj(A_n|\mathcal F_{n-1}) = \iy}.\]
Proof. Apply 3.1 to \(X_n = \sumo mn 1_{A_m} - \Pj(A_m|\mathcal F_{m-1})\).Lemma: \(X_n\) in 3 is a martingale wrt \(\mathcal F_n\).
Proof of Theorem. The technical difficulty is that we can’t work with the R-N derivative when one of \(\mu,\nu\) is singular wrt the other. So interpolate between them by setting \(\rh = \fc{\mu + \nu}2\). We have \[\dd{\mu_n}{\nu_n} = \left. \dd{\mu_n}{\pf{\mu_n+\nu_n}2}\right/ \dd{\nu_n}{\pf{\mu_n+ \nu_n}2}.\] Take \(n\to \iy\). (Some technical work here.)(skipped)
Polya urn scheme: An urn starts with \(r,g\) red and green balls. At each step draw a ball, replace it, and add \(c\) more balls of the same color. The fraction of green balls \(X_n\) after the \(n\)th draw is a martingale. (p. 241)
If \(\mu<1\) then \(Z_n=0\) for sufficiently large \(n\).
Proof. MarkovIf \(\mu=1\) and \(\Pj(\xi_i^m=1)<1\) then \(Z_n=0\) for sufficiently large \(n\).
Proof. \(Z_n\) is a martingale. \(Z_n=Z_\iy\) for large enough \(n\); the only way this can happen when \(\Pj<1\) is \(Z_n=0\).If \(\mu>1\) then \(\Pj(\forall n, Z_n>0) >0\).
Proof. Let \(\ph=\sumz k{\iy} p_k s^k\) be the pgf.Omitted.
Doob’s inequality: Let \(X_m\) be a submartingale, \(\ol X_n = \max_{0\le m\le n} X_m^+\), \(\la>0\). Then \[ \la \Pj(A) \le \E X_n 1_A \le \E X_n^+.\]
Proof. Interpret the event \(A\) as early stopping by defining the stopping time \(N=\inf \set{m}{X_m\ge \la \text{ or }m=n}\). Now apply 1.
Corollary: Kolmogorov’s inequalityBranching processes: Letting \(X_n=Z_n/\mu^2\), use 7 to get an expression for \(\E[X_n^2|\cal F_{n-1}]\) and \(\E X_n^2\). This shows \(\sup \E X_n^2<\iy\), so \(X_n\to X\) in \(L^2\).
A good counterexample is the 1-D simple random walk starting at 1 with absorbing barrier at 0.
\(X_i\) are uniformly integrable if \[\lim_{M\to \iy}\pa{\sup_{i\in I} \E(|X_i|;|X_i|>M)}=0.\] Note this implies \(\sup_{i\in I}\E|X_i|<\iy\).
Given \((\Om,\mathcal F_0, P)\) and \(X\in L^1\), \(\set{\E(X|\mathcal F)}{\mathcal F\text{ is a $\si$-field $\subeq \mathcal F_0$}}\) is uniformly integrable.
Proof. \(\le \E(|X|;\E(|X||\mathcal F)>M)\) and use Chebyshev and the fact that \(P(A_n)\to 0\implies \E(|X|;A_n)\to 0\).Proof. (\(n\implies n+1\))
If martingale \(X_n\to X\) in \(L^1\) then \(X_n = \E(X|\mathcal F_n)\).
Proof. Use 4 with \(A\in \mathcal F_n\).If \(\mathcal F_n \uparrow \mathcal F_\iy\) (i.e., \(\mathcal F_\iy = \si\pa{\bigcup_n \mathcal F_n}\)), then as \(n\to \iy\), \(\E(X|\mathcal F_n) \to \E(X|\mathcal F_\iy)\) a.s. and in \(L^1\).
Proof. \(Y_n=\E(X|\mathcal F_n)\) is a martingale. Use 6 and 5. Use the \(\pi\)-\(\la\) theorem to get that \(\int_A X\,dP = \int_A Y_\iy\,dP\) for all \(A\in \mathcal F_\iy\).Theorem (Levy’s 0-1 law): If \(\mathcal F_n\uparrow \mathcal F_{\iy}\) and \(A\in \mathcal F_{\iy}\) then \(\E(1_A|\mathcal F_n)\to 1_A\) a.s.
Note this proves Kolmogorov’s 0-1 law: take \(A\) to be in the tail \(\si\)-field.Theorem (DCT for conditional expectations): Suppose \(Y_n\to Y\) a.s., \(|Y_n|\le Z\) for all \(n\) where \(\E Z<\iy\), and \(\mathcal F_n\uparrow \mathcal F_\iy\) then \[\E(Y_n|\mathcal F_n)\to \E(Y|\mathcal F_\iy)\text{ a.s.}\]. (If instead \(Y_n\to Y\) is \(L^1\), then convergence is \(L^1\).)
Proof. Show \(|\E(Y_n|\mathcal F_n) - \E(Y|\mathcal F_n)| \to 0\) a.s. as \(n\to \iy\), and then use 5.7. To show the inequality, bound by a variable that doesn’t depend on \(N\), \[ \limsup_{n\to \iy} \E(|Y_n-Y||\mathcal F_n)\le \lim_{n\to \iy} \E(W_N|\mathcal F_n)\] where \(W_N\) is defined as the maximum Cauchy difference.
A backwards martingale is a martingale indexed by \(n\le 0\) adapted to an increasing sequence \(\mathcal F_n\): \[ \E(X_{n+1}|\mathcal F_n) = X_n,\quad n\le -1. \]
\(X_{-\iy}=\lim_{n\to -\iy} X_n\) exists a.s. and in \(L^1\). (If \(X_0\in L^p\), then convergence is in \(L^p\).)
Proof. Use the upcrossing inequality to show the limit exists (the proof is same as before except that \(n\to -\iy\) instead of \(\iy\)). \[(b-a)\E U_n \le \E(X_0-a)^+.\] By 5.1, \(X_n=\E(X_0|\mathcal F_n)\) is uniformly integrable; 5.2 shows convergence in \(L^1\).If \(X_{-\iy}= \lim_{n\to -\iy} X_n\) and \(\mathcal F_{-\iy} = \bigcap_n \mathcal F_n\), then \(X_{-\iy}= \E(X_0|\mathcal F_{-\iy})\).
Proof. Check that \(\int_A X_{-\iy} \,dP = \int_A X_0\,dP\) for all \(A\in \mathcal F_{-\iy}\).If \(\mathcal F_n\downarrow \mathcal F_{-\iy}\) as \(n\to -\iy\) (\(\mathcal F_{-\iy} = \bigcap_n\mathcal F_n\)), then \[\E(Y|\mathcal F_n)\to \E(Y|\mathcal F_{-\iy})\text{ a.s., in }L^1.\]
Ballot Theorem: Let \(\xi_j\in \N\) (ex. \(\{0,2\}\)) rv’s, \(S_k=\sumo ik \xi_i\), \(G=\set{S_j<j\text{ for }1\le j\le n}\). Then \(\Pj(G|S_n)\ge \pa{1-\fc{S_n}{n}}^+\) with equality if \(\Pj(\xi_j\le 2)=1\).
Proof: Let \(T=\inf\set{k\ge -n}{X_k}\). (Take averages over fewer elements; stop when it’s \(\ge 1\).) (Let \(T=-1\) if it’s \(\phi\).) \(X_T\ge 1\) on \(G^c\) and \(0\) on \(G\). \[\Pj(G^c|S_n) \le \E(X_T|\mathcal F_{-n}) = \fc{S_n}n.\]Hewitt-Savage 0-1 law (alternate proof): Lemma: Let \(A_n(\ph) = \rc{n\fp k} \sum_i \ph(X_{i_1},\ldots, X_{i_k})\). If \(X_i\) are iid and \(\ph\) is bounded, \(A_n(\ph) \to \E\ph(X_1,\ldots, X_k)\) a.s.
Proof. \[\begin{align} A_n(\ph) &= \E(\ph(X_1,\ldots, X_k)|\mathcal E_n) & \text{symmetry}\\ &\to \E (\ph(X_1,\ldots, X_k)|\mathcal E)\\ &=\E (\ph(X_1,\ldots, X_k)) \end{align}\]The last is because most permutations move \(X_1,\ldots, X_k\) away from positions in \([1,k]\). Formally, we first have \(\E[\ph(X_1,\ldots, X_k)|\mathcal E] \in \si(X_{k+1},\ldots)\), and then show lemma: if \(\E X^2<\iy\), \(\E(X|\mathcal G)\in \mathcal F\), and \(X\) is independent of \(\mathcal F\), then \(\E(X|\mathcal G) = \E X\). (Proof: Show \(\Var(Y)=0\).)
Then \(\mathcal E\) is independent of \(\mathcal G_k = \si(X_1,\ldots, X_k)\), so \(\mathcal E\) is independent of \(\si\pa{\bigcup_k \mathcal G_k}\).Theorem (de Finetti): If \(X_1,\ldots\) are exchangeable, then conditional on \(\mathcal E\), \(X_1,\ldots\) are iid.2
(\(X_1,\ldots\) is exchangeable if for each \(n\), \(\pi\in S_n\), \((X_1,\ldots, X_n)\) and \((X_{\pi(1)},\ldots, X_{\pi(n)})\) have the same distribution.)
Proof. Again by symmetry, \(A_n(\ph)\to \E[\ph(X_1,\ldots, X_k)|\mathcal E]\). For \(f,g\) on \(\R^k, \R\), calculate \(A_n(f)A_n(g)\) in terms of \(A_n(\ph), A_n(\ph_j)\) where \(\ph_j = (f\circ \pi_{-j})(g\circ \pi_j)\) to get \[ A_n(\ph) = \fc{n}{n-k+1} A_n(f) A_n(g) - \rc{n-k+1} \sumo j{k-1} A_n(\ph_j).\] Take limits, note that \(\E(\ph|\mathcal E) = \E(\ph_j | \mathcal E)\), and induct.
If \(X_n\) is a uniformly integrable submartingale then for any stopping time \(N\), \(X_{N\wedge n}\) is uniformly integrable.
Proof. From 4.1, \(\E X_{N\wedge n}^+ \le \E X_n^+\). Use the martingale convergence theorem to ge \(X_{N\wedge n} \to X_N\) a.s. and \(\E |X_N|<\iy\). Now split up \(|X_{N\wedge n}|\) based on \(N\wedge n \stackrel?n\) to show it’s uniformly integrable.If \(X_n\) is a uniformly integrable submartingale, then for any stopping time \(N\le \iy\), \(\E X_0\le \E X_n\le \E X_\iy\) where \(X_\iy = \lim X_n\).
Proof. Use 4.1 and 5.3, which shows that a uniformly integrable submartingale converges in \(L^1\).Proof. Use 7.3 with \(X_n=Y_{M\wedge n}, N=L\).
Let \(N=\begin{cases} L,&\text{on }A\\ M,&\text{on }A^c\end{cases}\). Use the first part on \(N\le M\) to get \(\E Y_N\le \E Y_M\). Note that \(N=M\) on \(A^c\) to get \[A\in \mathcal F_L\implies \E(Y_L;A) \le \E(Y_M;A) = \E[\E[Y_M|\mathcal F_L];A].\] (Define \(N\) so we can just localize the inequality to \(A\).)Theorem (Generalization of Wald’s equation): If \(X_n\) is a submartingale, \(\E[|X_{n+1}-X_n||\mathcal F_n]\le B\) a.s., and \(\E N<\iy\), then \(X_{N\wedge n}\) is uniformly integrable and \(\E X_N\ge \E X_0\).3 (For a martingale, we have \(=\).)
Proof. \[|X_{N \wedge n}| \le |X_0| + \sum_{m=0}^{\iy} |X_{m+1} - X_m| 1_{(N>m)} \le B\sum_{m=0}^{\iy} \Pj(N>m) = B \E N.\]
Let \(\xi_i=1,-1\) with probability \(p,1-p\). Suppose \(p>\rc 2\), \(a<0<b\).
(Absorption probabilities) Let \(T_x=\inf\set{n}{S_n=x}\). Then \[\Pj(T_a<T_b) = \fc{\ph(b)-\ph(0)}{\ph(b)-\ph(a)}.\]
Check that \(T_a\wedge T_b<\iy\) a.s. and apply 7.3. (Check directly, or use 5.6 on \(\ph(S_{N\wedge n})\).)\(\E T_b = \fc{b}{2p-1}\).
Proof. \(S_n-(p-q)n\) is a martingale. Either show \(\E T_b<\iy\), or apply 4.1 to \(T_b\wedge n\) and let \(n\to iy\).
Recall this means dot product with the differences.↩
ex. Chinese restaurant process.↩
To get Wald’s equation, let \(S_n\) be a random walk and \(X_n = S_n-n\E(S_1-S_0)\). 6. (Stopping theorem not requiring uniform integrability) If \(X_n\ge 0\) is a supermartingale and \(N\le \iy\) is a stopping time, then \(\E X_0\ge \E X_N\).
Proof. \(\E X_0\ge \E X_{N\wedge n}\). Now use MCT and Fatou on \(\E(X_N; N<\iy) + \E(X_N; N=\iy)\).↩