Martingales

Posted: 2016-04-08 , Modified: 2016-04-08

Tags: none

Durrett Ch. 5

See notebook 13.93-95 for exercises.

1 Conditional expectation

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

  1. \(Y\in \mathcal F\).
  2. for all \(A\in \mathcal F\), \(\int_A X\,dP = \int_B Y\,dP\).

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 \(\sgn(Y)\).

  1. Uniqueness (up to a.s.): We have \(0=\int_A Y-Y'\, dP\). Now take \(A=\{Y-Y'\ge \ep>0\}\).
  2. 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 \(X=X^+-X^-\).

Examples

The general strategy for finding the conditional expectation is to “guess and verify”.

  1. If \(X\in \mathcal F\) the \(\E[X|\mathcal F]=X\).
  2. If \(X\) is independent of \(\mathcal F\) then \(\E[X|\mathcal F]=\E X\).
  3. (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 \[\begin{align} \Pj(A|\mathcal G) &= \E[1_A|\mathcal G]\\ \Pj(A|B) &= \fc{P(A\cap B)}{P(B)}\\ \E[X|Y] &= \E[X|\si(Y)]. \end{align}\]
  4. 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.)

Properties

  1. Linearity: \(\E[aX+Y|\mathcal F] = a\E[X|\mathcal F] + \E[Y|\mathcal F]\).
  2. Monotonicity: \(X\le Y\implies \E[X|\mathcal F] \le E[Y|\mathcal F]\).
  3. Monotone convergence: \(X_n\ge 0, X_n\uparrow X\), \(\E X<\iy \implies \E[X_n|\mathcal F] \uparrow \E[X|\mathcal F]\).
  4. Jensen: If \(\ph\) is convex and \(\E|X|, \E|\ph(X)|<\iy\), then \(\ph(\E[X|\mathcal F])\le \E[\ph(X)|\mathcal F]\).
  5. Contraction in \(L^p\)
  6. \(\E[\E[Y|\mathcal F]]=\E Y\).
  7. (Tower property) If \(\mathcal F_1\subeq \mathcal F_2\) then \[\E[\E[X|\mathcal F_1]|\mathcal F_2] = \E[X|\mathcal F_1] = \E[\E[X|\mathcal F_2]|\mathcal F_1].\] (The coarser field wins.)
  8. (Taking \(X\in \mathcal F\) outside) If \(X\in \mathcal F\) and \(\E |Y|, \E|XY|<\iy\) then \(\E[XY|\mathcal F] = X\E[Y|\mathcal F]\).
  9. (Least squares, projection) If \(\E X^2<\iy\), then \(\E[X|\mathcal F]\) is the variable \(Y\in \mathcal F\) that minimizes \(\E(X-Y)^2\).

(Proofs straightforward from definition, omitted.)

2 Martingales, Almost Sure Convergence

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.)

  1. If \(X_n\) is a supermartingale and \(n>m\), then \(\E[X_n|\mathcal F_m]\le X_m\).

    Proof. Induct and use monotonicity, \[\begin{align} \E[X_{m+k}|\mathcal F_m] &= \E[\E[X_{m+k}|\mathcal F_{n+k-1}]|\mathcal F_m]\\ &\le \E[X_{m+k-1}|\mathcal F_m] \end{align}\]
  2. We get a reversed inequality for a submartingale, and equality for a martingale
  3. 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 \(X_n\).
  4. Ex. let \(\ph(x) = |x|^p\).
  5. The same holds if we change the conditions to: \(X_n\) is a submartingale wrt \(\mathcal F_n\), \(\ph\) is increasing convex.
  6. Cor. If \(X_n\) is sub, then \((X_n-a)^+\) is sub.

    If \(X_n\) is super, \(X_n\wedge a\) is super.
  7. 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.
  8. Corollary: If \(N\) is a stopping time and \(X_n\) is super, then \(X_{N\wedge n}\) is super.

    Proof. Let \(H_n = 1_{N\ge n}\).
  9. Theorem (Upcrossing inequality): Let \[\begin{align} N_0&=-1\\ N_{2k-1}&=\inf\set{m>N_{2k-2}}{X_m\le a}\\ N_{2k}&=\inf\set{m>N_{2k-1}}{X_m\ge b}\\ U_n&=\sup\set{k}{N_{2k}\le n}, \end{align}\]

    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 \[ \E(Y_n - Y_0) \ge \E(H\cdot Y)_n \ge (b-a) \E U_n.\]
  10. 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.
  11. If \(X_n\ge 0\) is super then as \(n\to \iy\), \(X_n\to X\) a.s., \(\E X\le \E X_0\).
  12. 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: \(A_n = A_{n-1} + \E[X_n|\mathcal F_{n-1}] - X_{n-1}\).

Exercises

  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 \(\E[\bullet|\mathcal F_n]\), \[\begin{align} \E[X_{n+1}|\mathcal G_n]&=X_n\\ \E[X_{n+1}|\mathcal F_n] = \E[\E[X_{n+1}|\mathcal G_n]|\mathcal F_n]&=X_n. \end{align}\]
  2. Let \(\mathcal F_n = \si(S_n,\ldots, S_1)\). \[\begin{align} \E[X_{n+1}|\mathcal F_n] &= \E[X_{n+1}|\si(S_n,\ldots, S_1)]\\ &=\E[f(S_n+\xi_n)|S_n] &\text{why?}\\ &=\rc{B(0,r)} \int_{B(S_n,r)} f(y)\le f(S_n)=X_n. \end{align}\]
  3. Let \(X_n = -\rc n\).
  4. Let \(X_n=\sumo in \xi_i\), \(\xi_i=\tcase{-1}{\text{w.p. }1-\rc{10^k}}{10^k}{\text{w.p. }\rc{10^k}}\).
  5. \(A_n = A_{n-1} + \E[1_{B_n}|\mathcal F_{n-1}]\).

3 Examples

  1. Let \(X_n\) be a martingale with \(|X_{n+1}-X_n|\le M<\iy\). Let \[\begin{align} C & = \{\lim X_n \text{ exists and is finite}\}\\ D & = \{\limsup X_n=\iy, \liminf X_n=-\iy\}. \end{align}\]

    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 \(X_{n\wedge \inf\set{n}{X_n\le -K}} + K + M\).
  2. 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 \(X_n = \sumo mn 1_{A_m} - \Pj(A_m|\mathcal F_{m-1})\).
  3. Theorem: If \(\mu_n\ll \nu_n\) for all \(n\), for \(X_n=\dd{\mu_n}{\nu_n}\), \(X=\limsup X_n\), then \[\mu(A) = \int_A X\,d\nu + \mu(A\cap \{X=\iy\}).\]
  4. 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.)
  5. (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)

Branching processes

The Galton-Watson process is defined as follows. Given \(\xi_i^n\) iid nonnegative integer rv’s, define \[\begin{align} Z_0 &=1\\ Z_{n+1} &= \tcase{\sumo i{Z_n} \xi_i^{n+1}}{\text{if }Z_n>0}{0}{\text{if }Z_n=0}. \end{align}\]
  1. Let \(\mathcal F_n = \si(\xi_i^m:i\ge 1, 1\le m\le n)\), \(\mu = \E \xi_i^m\) (same for all \(m\)). Then \(\fc{Z_n}{\mu^n}\) is a martingale wrt \(\mathcal F_n\).
  2. If \(\mu<1\) then \(Z_n=0\) for sufficiently large \(n\).

    Proof. Markov
  3. If \(\mu=1\) 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 \(Z_n=0\).
  4. If \(\mu>1\) then \(\Pj(\forall n, Z_n>0) >0\).

    Proof. Let \(\ph=\sumz k{\iy} p_k s^k\) be the pgf.
    1. Let \(\te_m = \Pj(Z_m=0)\). Then \(\te_m = \sumz k{\iy} p_k \te_{m-1}^k\). (All children families die out.)
    2. \(\ph'(1) =\mu>1 \implies \ph\) has a fixed point.
    3. \(\lim_{m\to \iy} \te_m = \rh\). (Take limits in \(\te_m=\ph(\te_{m-1})\).)
  5. Omitted.

4 Doob’s inequality, \(L^p\) convergence

  1. If \(X_n\) is a submartingale and \(N\) is a stopping time with \(\Pj(N\le k)=1\), then \[ \E X_0\le \E X_N \le \E X_k.\] (Note the first inequality is not true in general if \(N\) is unbounded.) Proof.
    1. Show that \(X_n - X_{N\wedge n}\) is a submartingale. To see this, write it as a dot product with a predictable sequence, \(X_n-X_{N\wedge n} = 1_{\{n-1 \ge N\}}\cdot X\).1
    2. (Alternative) Condition on \(N\) and use the submartingale property. See exercise 1.
  2. 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 inequality
  3. \(L^p\) maximum inequality: If \(X_n\) is a submartingale then for \(p>1\), \[ \E (\ol X_n^p) \le \pf{p}{p-1}^p \E(X_n^+)^p.\] Proof. Our “only tool” is to use Doob’s inequality to bound \(\Pj(\ol X_n\ge \la)\) (actually, \(\ol X_n\wedge M\)). To make this term appear, bound \(\E[(\ol X_n\wedge M)^p]\) by \(\int_0^{\iy} p\la^{p-1}\Pj(\ol X_n\wedge M \ge \la)\). We get one less power of \(\ol X_n \wedge M\) on the RHS. Use Holder and bootstrap.
  4. (\(L^1\) version)
  5. \(L^p\) convergence theorem: If \(X_n\) is a matringale with \(\sup \E|X_n|^p<\iy\) when \(p>1\), then \(X_n\to X\) a.s. and in \(L^p\). Proof. The hypothesis implies \((\E X_n^+)\) is bounded so a.s. follows from martingale convergence. To show \(X_n\to X\) in \(L^p\), it suffices to show that \(|X_n-X|^p\le Y\) with \(\E|Y|<\iy\) and \(X_n\to X\) a.s. (Dominated convergence theorem.) Bound \(|X_n-X|^p\le (2\sup |X_n|^p)\) and use \(L^p\) maximal inequality to bound the RHS (use MCT on \(\sup_{0\le m\le n}X_n\)).
  6. Orthogonality of martingale increments: Let \(X_n\) be a martingale with \(\E X_n^2<\iy\) for all \(n\). If \(m\le n, Y\in \cal F_m, \E Y^2<\iy\), then \[ \E[(X_n - X_m)Y]=0. \] Proof. Cauchy Schwarz gives this is \(L^1\). Now just use the martingale property.
  7. Conditional variance formula: If \(X_n\) is a \(L^2\) martingale, \(\E[(X_n-X_m)^2|\cal F_m]=\E [X_n^2|\cal F_m]-X_m^2\).
  8. Branching 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 \(L^2\).

Exercises

  1. Use submartingale property with the fact that \(N=j\) is in \(\cal F_j\). Now sum over \(N\).
  2. ?

A good counterexample is the 1-D simple random walk starting at 1 with absorbing barrier at 0.

  1. We have \(\E X_0=1>\E S_N\) where \(N=\inf\set{n}{S_n=0}\).
  2. We have \(\E(\max_m X_m) = \iy\), so no \(L^1\) maximal inequality holds.

5 Uniform integrability; convergence in \(L^1\)

\(X_i\) are uniformly integrable if \[\lim_{M\to \iy}\pa{\sup_{i\in I} \E(|X_i|;|X_i|>M)}=0.\] Note this implies \(\sup_{i\in I}\E|X_i|<\iy\).

  1. 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 \(P(A_n)\to 0\implies \E(|X|;A_n)\to 0\).
  2. (Relationship between uniform integrability and convergence in \(L^1\): If \(X_n\to X\) in probability, TFAE:
    1. \(\set{X_n}{n\ge 0}\) is uniformly integrable.
    2. \(X_n\to X\) in \(L^1\).
    3. \(\E|X_n|\to \E|X|<\iy\).

    Proof. (\(n\implies n+1\))

    1. Let \(\ph_M=\text{clamp}(-M,M)\). By the triangle inequality, \[\E|X_n-X| \le \E|\ph_M(X_n) - \ph_M(X)| + \E(|X_n|; |X_n|>M) + \E(|X|;|X|>M).\] The first term goes to 0 (\(\lim_{n\to \iy}\)) by bounded convergence, the second term goes to 0 (\(\lim_M\to \iy}\lim_{n\to \iy}\)) by uniform integrability; the last term goes to 0 by Fatou.
    2. Jensen.
    3. Write \(\E(|X_n|; |X_n|>M) \le\E|X_n| - \E \psi_M(|X_n|)\) where \(\psi_M(x)=x\) on \([0,M-1]\) and 0 on \([M,\iy)\). Use DCT.
  3. Theorem (Submartingale L1 convergence): For a submartingale, TFAE:
    1. Uniformly integrable
    2. Converges a.s. and in \(L^1\)
    3. Converges in \(L^1\) Proof.
    4. Uniform integrability \(\implies \sup \E|X_n|<\iy \implies X_n\to X\) a.s. by martingale convergence theorem. + Use 2.
    5. Clear.
    6. \(X_n\to X\) in \(L^1 \implies\) \(X_n\to X\) i.p. + Use 2.
  4. If \(X_n\to X\) in \(L^1\) then \(\E(X_n;A)\to \E(X;A)\).
  5. If martingale \(X_n\to X\) in \(L^1\) then \(X_n = \E(X|\mathcal F_n)\).

    Proof. Use 4 with \(A\in \mathcal F_n\).
  6. Theorem (Martingale L1 convergence): For a martingle, TFAE:
    1. Uniformly integrable
    2. Converges a.s. and in \(L^1\)
    3. Converges in \(L^1\)
    4. There is an integrable random variable \(X\) so that \(X_n=\E (X|\mathcal F_n)\)
  7. 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 \(A\in \mathcal F_\iy\).
  8. 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.
  9. 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.

Exercises

  1. We have \(\fc{\E(|X_i|;|X_i|\ge M)}{\E(\ph(|X_i|);|X_i|\ge M)}\to 0\), and the denominator is \(\le C\).
  2. By 5.7, \(\E(\te|Y_{[1,n]})\to \E(\te|Y_{[1,\iy)})\). By the Law of Large Numbers, the average of \(Y_{[1,n]}\) converges, say to \(\mu\). We have by calculation that \(\fc{\E(\mu|Y_{[1,n]})}{\E(\mu'|Y_{[1,n]})}\to \iy\). It converges i.p. to \(\te\). 5.6/7 gives that it converges a.s.
  3. ?
  4. ?
  5. By \(L^1\) convergence, \[\begin{align} \int_{X_m\le x\text{ i.o.},\neg D} \E(1_D|X_1,\ldots, X_n) &\to \int_{X_m\le x\text{ i.o.},\neg D} 1_D = 0\\ \pat{LHS} &\ge \int_{X_m\le x\text{ i.o.},\neg D, X_n\le x} \E(1_D|X_1,\ldots, X_n)\\ &\ge \Pj(X_n\le x) \Pj(D^c \wedge X_m\le x\text{ i.o.}|X_n\le x). \end{align}\] We may assume that if \(X_n=0\), then \(X_{n+1}=0\) (redefine the random variable so this is true). \(\Pj(X_n\le x)\) is bounded away from 0, so the second term goes to 0. Now \[\Pj(D^c\wedge X_m\le x\text{ i.o.}) = \Pj(D^c\wedge X_m\le x\text{ i.o.}|X_n\le x) + \cdots.\] Choosing \(n\) large enough, we make the RHS as small as we want.
  6. The condition in 5 holds.

6 Backwards Martingales

A backwards martingale is a martingale indexed by \(n\le 0\) adapted to an increasing sequence \(\mathcal F_n\): \[ \E(X_{n+1}|\mathcal F_n) = X_n,\quad n\le -1. \]

  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 \(L^1\).
  2. 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 \(A\in \mathcal F_{-\iy}\).
  3. If \(\mathcal F_n\downarrow \mathcal F_{-\iy}\) as \(n\to -\iy\) (\(\mathcal F_{-\iy} = \bigcap_n\mathcal F_n\)), then \[\E(Y|\mathcal F_n)\to \E(Y|\mathcal F_{-\iy})\text{ a.s., in }L^1.\]

Examples

  1. Strong law of large numbers: Let \(\xi_i\) be iid with \(\E|\xi_i|<\iy\). Let \(S_n= \sumo in \xi_i\) and \(X_{-n} = \fc{S_n}{n}\). Calculate by symmetry that \(\E(\xi_{n+1}|\mathcal F_{-(n+1)}\) and show \(X_{-n}\) is a backwards martingale. So \(\limn \fc{S_n}n = \E(X_{-1}|\mathcal F_{-\iy})\). Now use the Hewitt-Savage 0-1 law, \(\mathcal E\supeq \mathcal F_{-\iy}\) is trivial.
  2. 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 \(G\). \[\Pj(G^c|S_n) \le \E(X_T|\mathcal F_{-n}) = \fc{S_n}n.\]
  3. 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 \(\si\pa{\bigcup_k \mathcal G_k}\).
  4. 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.

7 Optional stopping theorems

  1. 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.
  2. If \(\E|X_N|<\iy\) and \(X_n1_{N>n}\) is uniofrmly integrable, then \(X_{N\wedge n}\) is uniformly integrable.
  3. 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 \(L^1\).
  4. Theorem (Optional stopping): If \(L\le M\) are stopping times and \(Y_{M\wedge n}\) is a uniformly integrable submartingale then \[\begin{align} \E Y_L &\le \E Y_M\\ \E X_N &\le \mathcal F_L. \end{align}\]

    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 \(A\).)
  5. 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. \[|X_{N \wedge n}| \le |X_0| + \sum_{m=0}^{\iy} |X_{m+1} - X_m| 1_{(N>m)} \le B\sum_{m=0}^{\iy} \Pj(N>m) = B \E N.\]

Example

Let \(\xi_i=1,-1\) with probability \(p,1-p\). Suppose \(p>\rc 2\), \(a<0<b\).

  1. Let \(\ph(x) = \pf{1-p}{p}^x\). \(\ph(S_n)\) is a martingale.
  2. (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 \(\ph(S_{N\wedge n})\).)
  3. \[\begin{align} \Pj(T_a<\iy) &= \pf{1-p}{p}^{-a}\\ \Pj(T_b<\iy) &= 1. \end{align}\] Proof. Let \(b\to \iy\); \(a\to \iy\), respectively in 2.
  4. \(\E T_b = \fc{b}{2p-1}\).

    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 \(n\to iy\).


  1. Recall this means dot product with the differences.

  2. ex. Chinese restaurant process.

  3. 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 \(\E(X_N; N<\iy) + \E(X_N; N=\iy)\).