Solutions to the Selected Exercises in R. Durrett's Probability: Theory and Examples, 5th edition.

The 2rd part of those solutions can be find in this pdf file.

4.1.2. Prove Chebyshev‘s inequality. If a>0 then

(1)P(|X|a|F)a2E[X2|F]

Proof: Notice that

(2)X21|X|aa2,a.s..

Therefore, form Theorem 4.1.9. (b) we have

(3)E[X2|F]a2E[1|X|a|F].

4.1.4. Use regular conditional probability to get the conditional Holder inequality from the unconditional one, i.e., show that if p,q(1,) with 1/p+1/q=1 then

(4)E[|XY||G]E(|X|p|G)1/pE(|Y|q|G)1/q.

Proof: Note that (R2,B(R2)) is a nice space. Therefore, according to Theorem 4.1.17. there exists a μ which is the regular conditional distribution for (X,Y) given G. In another word,

Now for a.e. w, it follows from the unconditional Holder inequality that

(5)Rd|xy|μ(ω,dx,dy)(R2|x|pμ(ω,dx,dy))1/p(R2|y|qμ(ω,dx,dy))1/q.

Using Theorem 4.1.16., the above inequality implies that

(6)E[|XY||G]E(|X|p|G)1/pE(|Y|q|G)1/q,a.s.

as desired.

4.1.6. Show that if GF and EX2< then

(7)E({XE(X|F)}2)+E({E(X|F)E(X|G)}2)=E({XE(X|G)}2).

Proof: Note that

(8)E(XE(X|F))=E[E(XE(X|F)|F)]=E(E(X|F)2).

Therefore we can verify the following Pythagorean law:

(9)E({XE(X|F)}2)=E(X2)+E(E(X|F)2)2E(XE(X|F))=E(X2)E(E(X|F)2).

Also note that E[E(X|F)|G]=E(X|G). So using this Pythagorean law three times, we get that the desired equality is equivalent to

(10)E[X2]E[E(X|F)2]+E[E(X|F)2]E[E(X|G)2]=E[X2]E[E(X|G)2].

This is trival.

4.1.9. Show that if X and Y are random variables with E(Y|G)=X and EY2=EX2<, then X=Y a.s.

Proof: Using the Pythagorean law (see the proof of exercise 4.1.6.), we have

(11)E({YX}2)=E(Y2)E(X2)=0

So we have X=Y a.s.

4.1.10. If E|Y|< and E(Y|G) has the same distribution as Y, then E(Y|G)=Y a.s.

Proof: First we proof that for each random variable X satisfies the condition of this exercise, we have {E(X|G)0}=a.s.{X0}. In fact, on one hand, Jensen's inequality implies that

(12)|E(X|G)|E(|X||G),a.s..

On the other hand, the condition that E(X|G)=dX says that

(13)E[|E(X|G)|]=E|X|=E(E(|X||G)).

So we must have |E(X|G)|=E(|X||G) as surely.

This leads us to

(14)E[X1E[X|G]0]=E[E[X|G]1E[X|G]0]=E[|E[X|G]|1E[X|G]0]=E[E[|X||G]1E[X|G]0]=E[|X|1E[X|G0]],

which forces that

(15)X1E[X|G]0=|X|1E[X|G]0,a.s..

So we must have {E(X|G)0}{X=|X|}={X0}. Noticing again we have

(16)P(E(X|G)0)=P(X0),

so we must have {E(X|G)0}=a.s.{X0} as required. Now take X=Yc, we get

(17){E(Y|G)c}=a.s.{Yc},cR.

This complete the proof.

4.2.2. Give an example of a submartingale Xn so that Xn2 is a supermartingale.

Proof: Xn=0.

4.2.3. Show that if Xn and Yn are submartingales w.r.t. Fn then XnYn is also.

Proof: Obviously XnYn is adapted to Fn. From XnYn|Xn|+|Yn|, we know XnYn is integrable. Finally, we have

(18)E[Xn+1Yn+1|Fn]E[Xn+1|Fn]E[Yn+1|Fn]XnYn.

4.2.4. Let Xn,n0, be a submartingale with supXn<. Let ξn=XnXn1 and suppose E(supξn+)<. Show that Xn converges a.s..

Proof: For each m0, define stopping time τm:=inf{k:Xk>m}. From supnXnτm+Xτm+=(Xτm1+ξτm)+Xτm1++ξτm+m+supnξn+, we know EsupnXnτm+<. According to Theorem 4.2.11, this says that Xnτm convergence a.s.. (Note that Xnτm is also a submartingale due to Theorem 4.2.9.). Therefore Xn convergence on the event {τm=}. Note that the condition supXn< implies that m=1{τm=}=a.s.Ω. Therefore, Xn converges a.s..

4.2.6. Let (Yk)kN be nonnegative i.i.d. random variables with EYm=1 and P(Ym=1)<1. By example 4.2.3 that Xn=mnYm defines a martingale. (i) Show that Xn0 a.s.. (ii) Use the strong law of large numbers to conclude (1/n)logXnc<0.

Proof. (i) Since Xn is a non-negative martingale, it convergence a.s.ly to a limit, say X. Fix a u>0 such that P(|Y1|u)>0. Then for each ϵ>0, since Xn is independent of Yn+1, we have

(19)P(|Xn+1Xn|ϵu)P(|Xn|ϵ)P(|Yn+11|u).

The left hand side converges to 0 as n, so we must have P(|Xn|>ϵ) converges to 0 as well. Therfore, X=0 a.s..

(ii) We can assume that P(Ym=0)=0, since if P(Ym=0)>0, it is easy to see that

(20)P(n>0 s.t. Xn=0)=1,

which implies that (1/n)logXn.

Now, assuming P(Ym=0)=0, we can write

(21)logXn=m=1nlogY1(,).

According to strong law of large numbers (Theorem 2.4.1. and 2.4.5.), we only have to show that ElogY1[,0).

Define Y1(n)=Y1n1<Y<n, then both Y1(n) and logY1(n) are integrable. From Jensen's inequality, we have

(22)ElogY1(n)logEY1(n),n1.

By monotonicity, taking n, we have

(23)ElogY1logEY1=0.

Now, we only have to show that ElogY10. In fact, if ElogY1=0, we have logY1 is integrable. So from ElogY1=logEY1 and Exercise 1.6.1. we have Y1=1 a.s., which contradicts to the condition P(Y1=1)<1.

4.2.8. Let Xn and Yn be positive integrable and adapted to Fn. Suppose

(24)E(Xn+1|Fn)(1+Yn)Xn

with Yn< a.s.. Prove that Xn converges a.s. to a finite limit.

Proof: Let

(25)Wn=Xnm=1n1(1+Ym),nN,

which is positive, integrable and adapted to Fn. From

(26)E(Wn+1|Fn)=1m=1n(1+Ym)E(Xn+1|Fn)Wn,

we know (Wn) is a supersomartingale. Theorem 4.1.12. says that (Wn) converges a.s. to a finite limit, say W. Since Ym are positive, we have

(27)logm=1n(1+Ym)=m=1nlog(1+Ym)m=1nYm.

From the condition Yn< a.s. and the fact that the left hand side of above is non-decreasing, we have m=1n(1+Ym) converges a.s. to a finite limit. Therefore Xn also converges a.s. to a finite limit.

4.2.9. Suppose Xn1 and Xn2 are supermartingales w.r.t. Fn, and N is a stopping time so that XN1XN2. Then

(28)Yn=Xn11N>n+Xn21Nn

and

(29)Zn=Xn11Nn+Xn21N<n

are supermartinales.

Proof: Clearely, Yn and Zn are integrable and adapted to Fn. Note that

(30)Yn+1=Xn+111N>n+1+Xn+121N=n+1+Xn+121N<n+1Xn+111N>n+1+Xn+111N=n+1+Xn+121N<n+1=Zn+1=Xn+111N>n+Xn+121Nn.

Therefore,

(31)E[Yn+1|Fn]E[Zn+1|Fn]=E[Xn+11|Fn]1N>n+E[Xn+12|Fn]1NnYnZn.

4.3.1. Give an example of a martingale Xn with supn|Xn|< and P(Xn=a i.o.)=1 for a=1,0,1.

Proof: Suppose that (Uk)kN are i.i.d. r.v. with uniform distribution in (0,1). Let X0=0. For each n1, if Xn=0, let Xn+1=1Un+11/21Un+1<1/2; if Xn0, let Xn+1=n2Xn1Un+1n2. Then (Xn) is a martingale since

(32)E[Xn+1|Fn]=1Xn=0E[Xn+1|Fn]+1Xn0n2XnE[Un+1n2|Fn]=Xn.

Note that

(33)P(Xn>1)=P(Xn10,Unn2)1n2.

So B.C. lemma says that P(Xn1 for n large engough)=1, which says that supn|Xn|< a.s..

It is elementary to see that

(34)n=1P(Xn+1=0|Fn)=n=1P(Xn+1=0|Fn)1Xn0=a.s.n=1(11n2)1Xn0.

Notice that we always have n21Xn0<, so the above identity says that

(35){n=1P(Xn+1=0|Fn)<}=a.s.{n=11Xn0<}{Xn=0 i.o.}.

Now, using Theorem 4.3.4. and above we have

(36){Xn=0 i.o.}c{Xn=0 i.o.}

in the sense of a.s.. This can only happen if P(Xn=0 i.o.)=1.

We can also verify that

(37)n=1P(Xn+1=1|Fn)n=1P(Xn+1=1|Fn)1Xn=0=12n=11Xn=0

So from what we have proved, we know that a.s.ly

(38)n=1P(Xn+1=1|Fn)=.

Using Theorem 4.3.4., we have that P(Xn=1 i.o.)=1. Similarly, we have P(Xn=1, i.o.)=1.

4.3.3. Let Xn and Yn be positive integrable and adpted to Fn. Suppose E(Xn+1|Fn)Xn+Yn, with Yn< a.s.. Prove that Xn converges a.s. to a finite limit.

Proof: Define Mn=Xnk=1n1Yk. Then (Mn) is a supermartingale, since

(39)E[Mn+1|Fn]Xn+Ynk=1nYk=Mn.

Define stopping times

(40)Nm:=inf{n:k=1nYk>m},mN,

Then it is easy to see that, for each mN,

(41)MnNm+m=XnNmk=1nNm1Yk+m,nN,

is a non-negative supermartingale. Therefore, Mn convergences a.s.ly on event {Nm=}. Finally, notice that event

(42)m=1{Nm=}={mN s.t. Yn<m}

is with probability 1.

4.3.5. Show n=2P(An|m=1n1Amc)= implies P(m=1Amc)=0.

Proof: Note that, there is a partition {A~n:nN} for the event m=1Am satisfying that

(43)m=1nAm=m=1nA~m,nN.

Define a filtration (Fn) such that

(44)Fn=σ(Am;m=1,,n)=σ(A~m;m=1,,n),nN.

Notice also that {A~1,,A~n,m=1nAmc} is a partition for the underlying probablity space Ω. Therefore, according to Example 4.1.5., we have

(45)P(An+1|m=1nAmc)=P(An+1|Fn),on m=1nAmc.

From the condition of this exercise, we have

(46)m=1P(Am+1|Fm)=,on m=1Amc.

Now, using Theorem 4.3.4. we get that

(47)m=1Amc{Am i.o.}(m=1Amc)c

in the sense of almost sure. This can only happen if P(m=1Amc)=0.

For the next two exercises, in the context of Kakutani dichotomy for infinite product measures on page 235, suppose Fn, Gn are concentrated on {0,1} and have Fn(0)=1αn, Gn(0)=1βn.

4.3.9. Show that if αn< and βn= then μν.

Proof: Let A:={ξn0 i.o.}. According to B.C. lemma, condition αn< says that μ(A)=0; condition βn= says that ν(A)=1. So we must have μν.

4.3.10. Suppose 0<αn,βn<1. Show that |αnβn|< is sufficient for μν in general.

Proof: Let (Uk)kN be i.i.d. r.v. uniform distribution on [0,1] w.r.t. probability space (Ω,F,P). Define {0,1}N×{0,1}N random element (ξ1,ξ2) by

(48)ξk1:=1Ukαk,ξk2:=1Ukβk,kN.

Then we have ξ1 has distribution μ and ξ2 has distribution ν. Note that

(49)k=1P(ξk1ξk2)=k=1|αkβk|<,

therefore, according to B.C. lemma, we have

(50)P(K=1k>K{ξk1=ξk2})=1.

This says that there exists KN such that P(k>K{ξk1=ξk2})>0. On the other hand, it is obvious that

(51)P(k=1K{ξk1=ξk2})>0,

so from the independency, we have

(52)P(ξ1=ξ2)=P(k=1K{ξk1=ξk2})P(k>K{ξk1=ξk2})>0.

Now, suppose that μν is not true, then according to Kakutani dichotomy, we have μν. This says that, there exists a subset ARN, wuch that μ(A)=ν(Ac)=1. In this case, we have

(53)P(ξ1A)=μ(A)=1=ν(Ac)=P(ξ2Ac),

which says that

(54)P(ξ1ξ2)P(ξ1A,ξ2Ac)=1.

This is a contradiction.

4.3.13. Galton and Watson who invented the process that bears their names were interested in the survival of family names. Suppose each family has exactly 3 children but coin flips determine their sex. In the 1800s, only male children kept the family name so following the male offspring leads to a branching process with p0=1/8, p1=3/8, p2=3/8, p3=1/8. Compute the probability ρ that the family name will die out when Z0=1.

Proof: According to Theorem 4.3.12. we know that ρ is the only solution of

(55)φ(ρ)=ρ

in [0,1), where

(56)φ(ρ)=18+38ρ+38ρ2+18ρ3.

Solving this gives that ρ=52. 4.4.3. Suppose MN are stopping times. If AFM then L=M1A+N1Ac is a stopping time.

Proof: According to Theorem 7.3.6. we have AcFMFN. Therefore, for each t0, we have

(57)A{Mt}Ft;Ac{Nt}Ft.

From the above, we have {Lt}=(A{Mt})(Ac{Nt})Ft.

4.4.5. Prove the following variant of the conditional variance formula. If FG then

(58)E(E[Y|G]E[Y|F])2=E(E[Y|G])2E(E[Y|F]2).

Proof: Note that E(E[Y|G]|F)=E[Y|F]. So according to the Pythagorean law (see the Slution to Excise 4.1.6.) we get the desired result.

4.4.7. Let Xn be a martingale with X0=0 and EXn2<. Show that

(59)P(max1mnXmλ)EXn2EXn2+λ2.

Proof: According to Theorem 4.2.6. we have (Xn+c)2 is a submartingale where c is an arbitrary real number. Therefore, for each cR, according to Doob's inequality

(60)P(max1mnXmλ)P(max1mn(Xn+c)2(λ+c)2)E(Xn+c)2(λ+c)2=EXn2+c2(λ+c)2.

Now, taking c=EXn2λ we have

(61)P(max1mnXmλ)EXn2EXn2+λ2.

4.4.8. Let Xn be a submartingale and log+x=max(logx,0). Prove

(62)EX¯n(1e1)1{1+E(Xn+log+(Xn+))},

where X¯n=maxk=1nXk.

Proof: Fix an M>1. Note that

(63)E[X¯nM]=0P(X¯nMλ)dλ1+1MP(X¯nMλ)dλ1+1MP(X¯nλ)dλ

Doob's inequality then says that

(64)E[X¯nM]1+1M1λE[Xn+;X¯nλ]dλ1+E[Xn+1M1λ1X¯nλdλ]=1+E[Xn+log(X¯nM)].

Now, use the calculus fact that alogbalog+a+b/e, we have

(65)E[X¯nM]1+E[Xn+log+Xn++X¯nMe].

This says that

(66)E[X¯nM](11e)1{1+E[Xn+logXn+]}.

Finally, taking M, using monotone convergence theorem, we get the desired result.

4.4.9. Let Xn and Yn be martingales with EXn2< and EYn2<. Show that

(67)E[XnYn]E[X0Y0]=m=1nE[(XmXm1)(YmYm1)].

Proof: Since E[Xn+1Xn|Fn]=0 and YnFn we have by Theorem 4.4.7. that E[(Xn+1Xn)Yn]=0. Similarly we have E[(Yn+1Yn)Xn]=0. Now it is easy to calculate that

(68)E(Xn+1Xn)(Yn+1Yn)=E(Xn+1Xn)Yn+1=E[Xn+1Yn+1XnYn+Xn(YnYn+1)]=E[Xn+1Yn+1XnYn].

From this to the desired result is trival.

4.4.10. Let Xn,n0, be a martingale and let ξn=XnXn1 for n1. If EX02,m=1Eξm2< then XnX a.s. and in L2.

Proof: Using the result in Excise 4.4.9, we have

(69)EXn2=EX02+m=1nEξm2.

Therefore supnEXn2=EX02+m=1Eξm2<. According to Theorem 4.4.6. we get the desired result.

4.6.4. Let Xn be r.v.'s taking values in [0,). Let D={Xn=0 for some n1} and assume

(70)P(D|X1,,Xn)δ(x)>0a.s. on {Xnx}.

Use Theorem 4.6.9 to conclude that P(D{limnXn=})=1.

Proof: Let Fn=σ(X1,,Xn),n1 and F=σ(nFn). According to DF, we have by Levy's 0-1 law that E[D|Fn]1D a.s.. For each x>0, and each element ω{Xnx i.o.}, there exists a sequence of integers (ni)iN such that for each iN, we have Xni(ω)x. Therefore, for this ω,

(71)1D(ω)=limnE[D|Fn](ω)=limiE[D|Fni](ω)δ(x)>0.

So we must have 1D=1 on this event {Xnx i.o.}. This says that {Xnx i.o.}D for each x>0. Therefore, xN{Xmx i.o.}D. Finally, noticing that xN{Xmx i.o.}={limnXn=}c, we must have the desired result.

4.6.5. Let Zn be a branching process with offspring distribution pk. Use the last result to show that if p0>0 then P(limnZn=0 or )=1.

Proof: Let D:={limnZn=0}={Zn=0,nN} be the event of extinction. Let (ξin)i,nN be i.i.d. r.v. used in (4.3.4.). Let Fn=σ(Z1,,Zn). Now for each x>0, on event {0<Znx} we have

(72)P(D|Fn)P(Zn+1=0|Fn)=P(ξin+1=0,i=1Zn|Fn)=p0Znp0x>0.

On event {Zn=0}, we have P(D|Fn)=1p0x>0. Now, using Exercise 4.6.4. we have

(73)P(D{limnZn=})=1

as desired.

4.6.7. Show that if FmF and YnY in L1 then E(Yn|Fn)E(Y|F) in L1.

Proof: According to Theorem 4.6.8. we have E[Y|Fn]E[Y|F] in L1. So we only have to show that

(74)E(Yn|Fn)E(Y|Fn)L10.

In fact,

(75)E[|E(Yn|Fn)E(Y|Fn)|]E[E[|YnY||Fn]]=E|YnY|0.

4.7.3. Prove directly from the definition that if (Xl)lN{0,1} are exchangeable

(76)P(X1=X2=Xk=1|Sn=m)=(nknm)/(nm).

Proof: Define

(77)N={w{0,1}n:wl=1,1lk;l=1nwl=m};M={w{0,1}n:l=1nwl=m}.

Note that, for each wN, there exists a purmutation Γω on {1,,n} such that

(78)wΓω(l)=11lm,l{1,,n}.

Now, writting X=(Xl)lN, we have

(79)P(Xl=1,1lk;Sn=m)=wNP(Xl=wl,1ln)=wNP(XΓω(l)=11lm,1ln)=wNP(Xl=11lm,1ln)=#NP(Xl=11lm,1ln).

Similarly we have P(Sn=m)=#MP(Xl=11lm,1ln).

Therefore, we have

(80)P(X1=X2=Xk=1|Sn=m)=#N#M=(nknm)/(nm).

4.7.4. If (Xk)kNR are exchangeable with EXi2< then E(X1X2)0.

Proof: Note that

(81)0(n2)1E(X1++Xn)2=(n2)1nEX12+EX1X2nEX1X2.

4.7.5. If (Xk)kN are i.i.d. with EXi=μ and var(Xi)=σ2< then

(82)(n2)11i<jn(XiXj)22σ2.

Proof: Note that

(83)(n2)11i<jn(XiXj)2=:AnEn.

Therefore, since En+1En, we know Fk:=Ek,k=1,2, is a filtration with index N. Therefore we have

(84)An=E[An|En]=1(n)21i,jnE[(XiXj)2|En]=E((X1X2)2|En)=E((X1X2)2|Fn)nThm 4.7.3.E((X1X2)2|F),a.s.

According to Hewitt-Savage 0-1 law, we have F=E is trival. So

(85)E((X1X2)2|F)=E((X1X2)2)=2σ2.

4.8.3. Let Sn=ξ1++ξn where the ξi are independent with Eξi=0 and var(ξi)=σ2. Sn2nσ2 is a martingale. Let T=min{n:|Sn|>a}. Then we have ETa2/σ2.

Proof: Without loss of generality, we assume ET<. (Otherwise, the desired result is trival.) According to Wald's second identity (Excise 4.8.4. below), we have

(86)σ2ET=EST2a2.

4.8.4. Let Sn=ξ1++ξn where the ξi are independent with Eξi=0 and var(ξi)=σ2. Show that if T is a stopping time with ET< then EST2=σ2ET.

Proof: Since SnT2(nT)σ2,n1 is a martingale, we have

(87)E[SnT2σ2(nT)]=0,nN.

Therefore, we have

(88)supnE[SnT2]=σ2supnE(nT)σ2ET<.

This tells us that SnT,n1 is a L2-martingale. Therefore SnTL2ST and

(89)EST2=limnESnT2=limnσ2E(nT)=MCTσ2E(T).

4.8.5. Let (ξk)kN be independent with P(ξi=1)=p and P(ξi=1)=1p where p<1/2. Let Sn=S0+ξ1++ξn and let V0=min{n0:Sn=0}. Theorem 4.8.9 tells us that ExV0=x/(12p). Let Yi=ξi(pq) and note that EYi=0 and

(90)var(Yi)=var(Xi)=EXi2(EXi)2

then it follows that (Sn(pq)n)2n(1(pq)2) is a martingale. (a) Use this to conclude that when S0=x the variance of V0 is

(91)x1(pq)2(qp)3.

(b) Why must the answer in (a) be of the form cx?

Proof. (a). Since V0 is a stopping time with finite expectation. Using Wald's second identity (Excise 4.8.7.), we have

(92)Ex[(SV0(pq)V0)2V0(1(pq)2)]=x2.

From the fact that SV0=0 and ExV0=x/(12p), we can calculate the desired result.

(b) Define Vy=min{n0,Sn=y}. Then, according to S0=x>0, we have Vx=0. From the fact that |Sn+1Sn|=1, and the fact that EV0<, we know that

(93)0=VxVx1V0<.

Moreover, it can be verified that {Ty=Vy1Vy:y=x,x1,} are i.i.d. random variables. (Ty are the time process (Sn) spend from first hitting position y to first hitting position y1.) So

(94)EV0=k=1xETk=xc.

4.8.7. Let Sn be a symmetric simple random walk starting at 0, and let T=inf{n:Sn(a,a)} where a is an integer. Find constants b and c so that Yn=Sn46nSn2+bn2+cn is a martingale and use this to compute ET2.

Proof: First, since Sn2n is a martingale, we have SnT2nT is a martinale. Therefore

(95)E[SnT2]=E[nT].

Note that SnT2 is bounded by a2; nT is monotonic in n. Therefore, using bounded/monotonic convergence theorem, we get

(96)a2=E[T].()

It is elementary to verify that

(97)E[Yn+1|Fn]Yn=(2b6)n+b+c5.

Therefore, Yn is a martingale iff b=3 and c=2. Now, set b=3,c=2. since (YTn)nN is also a martingale, we have

(98)E[SnT4+3(nT)2+2(nT)]=E[6(nT)SnT2].

Note that (SnT4)nN is bounded by a4; (nT)nN is monotonic in n; ((nT)SnT2)nN is dominated by Ta2. Therefore, using bounded/monotonic/dominated convergence theorem, we get

(99)E[a4+3T2+2T]=E[6Ta2].

From (), we have ET2=(5a42a2)/3.

4.8.10. Consider a favorable game in which the payoff ξk are 1,1 or 2 with probability 1/3 each. Use the results of the previous problem to compute the probability we ever go broke (i.e. our winings Wn reach 0) when we start with i.

Proof: It is elementary to verify that, if θ0=ln(21)<0, then

(100)E[exp(θ0ξk)]=13(eθ0+eθ0+e2θ0)=1.

It is well known that Xn:=exp(θ0Wn) is a martingale (the so-called exponential martingale). Note that it is non-negative, so it must have almost sure limit X. In fact, since Wn almost surely, we must have X=0.

Now, consider the martingale XnT where T is the broken time (hitting time at 0). Note that WnT0, so XnT[0,1] is a bounded martingale. Therefore, we have

(101)XnTnL1XT1T<+X1T==1T<.

This implies that

(102)P(T<)=E[1T<]=E[X0]=exp(θ0i)=(21)i.

6.1.1. Show that the class of invariant events I is a σ-field, and XI if and only if X is invariant, i.e., Xφ=X a.s.

Proof: I is a sigma-field since (1) if AI, then φ1Ac=(φ1A)c=a.s.Ac, which says that AcI. (2) I since φ1()=. (3) if (Ak)kN is a sequence in I, then

(103)φ1nNAn=nNφ1An=a.s.n1An,

which says that nNAnI.

Also note that

(104)XI{XB}I, Borel B{ω:X(ω)B}=a.s.φ1{ω:X(ω)B}={ω:Xφ(ω)B}, Borel BX=a.s.Xφ.

6.1.2. Call A almost invariance if P(AΔφ1(A))=0 and call C invariant in the strict sense if C=φ1(C). (i) Let A be any set, let B=n=0φn(A). Show φ1(B)B. (ii) Let B be any set with φ1(B)B and let C=n=0φn(B). Show that φ1(C)=C. (iii) Show that A is almost invariant if and only if there is a C invariant in the strict sense with P(AΔC)=0.

Proof: (i) φ1B=n=0φ1φnA=n=1φnAB. (ii) Since φ1BB, we have that φ1(C)=nNφ1φnB=n=1φnB=n=0φnB=C.

(iii) Define B and C as above. Since A is invariance, we have A=a.s.φ1A. It can be verified that if two measurable subsets Ω1,Ω2 of Ω satisfies Ω1=a.s.Ω2, then φ1Ω1=a.s.φ1Ω2. In fact,

(105)P(φ1(Ω1)Δφ1(Ω2))=P(φ1(Ω1ΔΩ2))=P(Ω1ΔΩ2)=0.

Using this fact multiple times we have A=a.s.φk(A) for any kN. Therefore B=a.s.A. And we also have B=a.s.φk(B) for any kN. This tells us that A=a.s.C. Yet (ii) already shows that C is strictly invariance.

6.1.3. (i) Show that if θ is irrational, xn=nθ mod 1 is dense in [0,1). (ii) Use Theorem A.2.1. to show that if A is a Borel subset of [0,1) with |A|>0, then for any δ>0 there is an interval J=[a,b) so that |AJ|>(1δ)|J|. (iii) Let θ be irrational. Combine this with (i) to conclude if A is an a subset of [0,1) which is invariant under the operator

(106)φ:yy+θ mod 1

and |A|>0, then |A|=1.

Proof: (i) Consider a 1-1 map x[0,1)e2πxiS1:={zC:|z|=1}. For any α,βS1, there is a natural distance

(107)d(α,β):=length of the shorter arc connecting α and β on S1.

We only need to prove that {αn=e2πnθi:nN} is dense on S1. More precisely, fixing an arbitrary β on S1 and a large N, we only have to prove that there exists a n such that d(αn,β)2πN.

In fact, it is easy to verify that

  1. d(αn,αm)=d(0,αnm) for all n,mN;

  2. all αn are distinct, so d(αn,αm)2πN for some m<nN. Fix this m and n.

  3. S1=k=0{α:α lies on the shorter arc connecting αk(nm) and α(k+1)(nm)}

Now, for that fixed β, we know from 3. that there exists a k0 such that β lies on the shorter arc connecting αk(nm) and α(k+1)(nm). Therefore, for this k,

(108)d(αk(nm),β)d(αk(nm),α(k+1)(nm))=d(0,αnm)2πN,

as desired.

(ii) Let ϵ=δ1δ|A|. Using Theorem A.2.1. there exists countable disjoint intervals Jk:=[ak,bk),k=1, such that Ak=1Jk and k=1|Jk|<|A|+ϵ. Suppose that non of those intervals Jk satisfies the desired property that |AJ|>(1δ)|J|, then

(109)k=1|Jk|k=1|JkA|1δ=|A|1δ.

Therefore ϵ>|A|1δ|A|=ϵ. This is a contradiction.

(iii) Fix an arbitrary 1>δ>0. Note that if J=[a,b) is an interval satisfies the condition |AJ|(1δ)|J| then either interval J=[a,a+b2) or interval J=[a+b2,b) also satisfy the same condition. This and (ii) implies that for any small ϵ>0, there exists an interval J=[a,b) satisfies |AJ|(1δ)|J| and |J|<ϵ. Fix this ϵ>0 and interval J. Let NN be the unique integer such that 1N+1|J|<1N. Thanks to (i), for each k=0,,N1, there exists an integer nk such that

(110)kNφnka<φnkb<k+1N.

Note that Jk=[φnka,φnkb),k=0,N1 are disjoint intervals, A is φ-invariant i.e. φ1A=A, and φ is measure preserving. Therefore

(111)|A||k=0N1(AJk)|=k=0N1|AJk|=k=0N1|φnk(AJk)|=k=0N1|AJ|N(1δ)|J|(1ϵ)(1δ).

Since ϵ>0 and δ>0 are arbitrary, so |A|=1.

6.1.4. For any stationary sequence {Xn,n0}, there is a two-sided stationary sequence {Yn:nZ} such that (Xn)n0=d(Yn)n0.

Proof: Give a stationary process (Xn)nN. According to Kolmogrove's extension theorem, there is a stochastic process (Yn)nZ such that for any n1<n2<<nk, we have

(112)(Yni)i=1k=d(Xnin1)i=1k.

(It is elementary to verify that hose finite dimensional distribution if consistent.) So, (Xn)nN=d(Yn)nN.

We also need to verify that {Yn} is stationary. This is elementary from its definition.

6.1.5. If (Xk)kN is a stationary sequence and g:RNR is measurable then Yk=g(Xk,Xk+1,) is a stationary sequence. If Xn is ergodic then so is Yn.

Proof: The shift operator is defined as usual

(113)θw=(w2,w3,),wRN.

Define another operator G:RNRN with

(114)Gw=(g(w),g(θw),,g(θkw)),wRN,

then Y=G(X). It can also be verified that Gθ=θG.

Therefore for each measurable subset ARN, we have

(115)YθkAG(X)θkA(θkG)(X)A(Gθk)(X)AθkXG1AXθkG1A.

Therefore, if X is stationary, we have

(116)μY(θ1A)=μX(θ1G1A)=μX(G1A)=μY(A),

which says that Y is also stationary.

Note that if A is invariant, i.e. θ1A=A, then so is G1A, since θ1G1A=G1θ1A=G1A. Therefore, if X is egodic, then for any invariant subset ARN, we have

(117)μY(A)=μX(G1A){0,1},

which says that Y is also ergodic.

6.1.6. Let (Xk)kN be a stationary sequence. Let n< and let (Yk)kN be a sequence so that (Ynk+1,,Yn(k+1)),k0 are i.i.d. and (Y1,,Yn)=(X1,,Xn). Finally, let ν be uniformly distributed on {1,2,,n}, independent of Y, and let Zm=Yν+m for m1. Show that Z is stationary and ergodic.

Proof: The shift operator is defined as usual

(118)θw=(w2,w3,),wRN.

It is easy to see that for each measurable ARN, we have

(119)P(YθnA)=P(YA).

Therefore

(120)P(Zθ1A)=P(Yθν1A)=k=1n1nP(Yθk1A)=k=1n1nP(YθkA)=P(ZA).

This says that Z is stationary.

Now assume that A is shift invariant i.e. θ1A=A . Note that

(121){ZA}=k=1n{ZA,ν=k}=k=1n{YθkA,ν=k}={YA}

Since Y is ergodic wrt operator θn and A is shift invariant wrt operator θn, so we have P(YA){0,1}. This says that P(ZA){0,1}. So Z is ergodic.

6.1.7. Let φ(x)=1/x[1/x] for x(0,1) and A(x)=[1/x], where [1/x]= the largest integer 1/x. Then an=A(φnx),n=0,1,2, gives the continued fraction representation of x, i.e.

(122)x=1/(a0+1/(a1+1/)).

Show that φ preserves μ(A)=1log2Adx1+x for A(0,1).

Proof: It can be verified that for each 0<ab<1, we have

(123)φ1[a,b)=nN(1n+b,1n+a].

Therefore, we can calculate that

(124)μφ1[a,b)=nN1log21n+b1n+adx1+x=1log2abdx1+x=μ[a,b).

Using π-λ theorem, we can verifyopen φ preserves μ.

Exercise 6.2.1. Show tha if XLp with p>1 then the convergence in Theorem 6.2.1 occurs in Lp.

Proof: Take an arbitrary M>0. Let XM:=X1|X|M and XM:=X1|X|>M. We claim that

(125)lim supn1nk=0n1XφkE[X|I]p2XMp.

In fact, on one hand we have

(126)1nk=0n1XMφkna.s.&LpE[XM|I],

where the almost sure convergence is due to Ergodic theorem, and the Lp convergence is then followed by bounded convergence theorem. On the other hand, we have

1nk=1n1XMφkE[XM|I]p1nk=1n1XMφkp+E[XM|I]p2XMp,n0.

Now, since M is arbitrary and that XMp0 as M, we get the desierd result.

Exercise 6.2.2 (1) Show that if gn(w)g(w) a.s. and E(supk|gk|)<, then limn1nm=0n1gm(φmw)=E(g|I)a.s.

Proof: We claim that

(127)lim supn1nm=0n1gmφE[g|I]a.s..

In fact, taking an arbitrary M>0, we can define a almost sure finite random variable hM:=supmM|gmg| using the condition E(supk|gk|)<. Then we have by ergodic theorem that

1nm=0n1gmφm1nm=0M1(g+h0)φm+1nm=Mn1(g+hM)φmnE[g+hM|I]a.s..

According to the fact that |hM|g+supk|gk| and hM0 as M, we have E[g+hM|I]E[g|I] a.s. due to the dominated convergence theorem. Therefore, the claim is true. Applying this claim to (gn)n=1,, we get that

(128)lim infn1nm=0n1gmφE[g|I]a.s..

Exercise 6.2.3 Let Xj=Xφj, Sk=X0++Xk1, Ak=Sk/k and Dk=max(A1,,Ak). Show that if α>0 then

(129)P(Dk>α)α1E|X|.

Proof: Define Xj=Xjα, Sk=X0++Xk1, Ak=Sk/k and Dk=max(A1,,Ak). Then, it is easy to see that Sk=Skαk, Ak=Akα and Dk=Dkα. Lemma 6.2.2. says that E[X;Dk>0]0. Therefore E|X|E[X;Dk>α]αP(Dk>α) as desired.

Exercise 6.3.1 Let gn=P(S10,,Sn0) for n1 and g0=1. Show that

(130)ERn=m=1ngm1

Where Sn and Rn is the same as Theorem 6.3.1.

Proof: Note that

(131)Rn=1+1Sn1{Sn}+1Sn2{Sn1,Sn}++1S1{S2,,Sn}.

Therefore,

ERn=1+m=1n1P(Sm{Sm+1,,Sn})=1+m=1n1P(Sm+1Sm0,,SnSm0})=1+m=1n1P(S10,,Snm0})=m=1ngm1.

Exercise 6.3.2 Under the setting of Theorem 6.3.2. Show that if we assume P(Xi>1)=0,EXi>0, and the sequence Xi is ergodic, then P(A)=EXi.

Proof: It is elementary analysis that if sn/nc>0, then we must have

(132)n1max1knskc

and

(133)infk=1,sk>.

Ergodic theorem syas that

(134)SnnEXi>0,a.s.

so we must have

(135)n1max1knSkEXi,a.s.

and

(136)M:=infk=1,Sk>,a.s.

Note, from the condition P(Xi>1)=0, we have

(137)max1knSkRnmax1knSkm,

which now implies that

(138)RnnEXi.

However, from Theorem 6.3.1. we already know that n1RnP(A). Therefore, we must have P(A)=EXi.

Exercise 6.3.3 Show that if P(XnA at least once)=1 and AB= then

(139)E(1mT11XmB|X0A)=P(X0B)P(X0A).

Proof: We can find a two-side stationary process which has the same finite demisional distribution same as (Xn)nN. With some abuse of notations, we denote such two-side stationary process as (Xn)nZ. Now, we can verify that

P(X0A)E[m=1T11XmB|X0A]=E[tNm=1t1XmB,T1=t;X0A]=m=1P(XmB,T1m;X0A)=m=1P(X0A,X1A,,Xm1A,XmB)=m=1P(XmA,Xm+1A,,X1A,X0B)=P(X0B).

Exercise 6.3.4 Consider the special case in which Xn{0,1}, and let P¯=P(|X0=1). Here A=1 and so T1=inf{m>0:Xm=1}. Show P(T1=n)=P¯(T1n)/E¯T1.

Proof: From Theorem 6.3.3. we know that P(X0=1)E¯T1=1. Therefore

(140)P¯(T1n)E¯T1=P(T1n|X0=1)P(X0=1)=P(T1n,X0=1).

On the other hand, with some abuse of natations, assuming that (Xn)nZ is a two-sided stationary sequence, we have

(141)P(T1=n)=m=0P(Xm=1,Xm+1=0,,Xn1=0,Xn=1)(142)=m=0P(X0=1,X1=0,,Xm+n1=0,Xm+n=1)(143)=P(X0=1,T1n).