This content was uploaded by our users and we assume good faith they have the permission to share this book. If you own the copyright to this book and it is wrongfully on our website, we offer a simple DMCA procedure to remove your content from our site. Start by pressing the button below!
) + < ^ I ^ | M ( f , x"; dy) where ((a'^(r, x^))) is non-negative definite and M(r, x^; •) is a Levy measure. More important, we develop from these considerations the intuitive picture of the process x(*) leaving x(t) like the independent increment process with characteristics a(t, x(t)), b(t, x(r)), and M(r, x(t); •). Throughout this book we will be restricting ourselves to continuous Markov processes. For a continuous process, the Levy measure M must be absent. That is, if x( •) is a continuous Markov process and (0.1) obtains, then 1} ^ C?'(R'') such that (p„-^f(t2, ') uniformly, and set/„(s, •) = r,,,2 (s,-)\\. |a|<« • R'"^^,and A:[0, (x>) x JE^Sj+i by: z{t) = s: (r, x(r)) i ^ } , then F,,((r,x(rAO)eL/)>0 so long as V is an open set in [0, 00) x R*^ for which there is a (/) e C{s, x) with the property that (r, (/)(r)) e V and (w, (/>(w)) G ^ for all w between s and f. Finally, apply this fact to derive Nirenberg's strong maximum principle: if/ e Cl' ^(^) and (dfjdt) + L^f>Q on ^, then/(s, x) = max^, y)g^/(f, y) for some (s, x) e ^ implies that /(r, }^) = / ( s , x) for any (f, y) e ^ with the property that there is a (p e C(s, x) satisfying (w, (/?(w)) G ^, s < w < r, and (p(t) = y. In view of the preceding considerations, the proof is an easy consequence of the relation : 0 and £ > 0, there is a d,t{e) > 0, depending only on d, p, A, A, and \\b\\ as well as h and e, such that for any T > h and (a, i), (s, x) e[0,T-h]x R^ satisfying \(T - s\ v \i - x\ < S^ie): |£Q.^[0]-£:Q-^[0]| <e||a>|| whenever 0:Q -^ C is a bounded Jf^(^ G{x{t)'.t > T))-measurable function. In particular, the family {Qs,x'(s,x) € [0,oo) x R^} is strongly Feller continuous. Proof Let {P^ ^: (s, x) e [0, oo) x R**} be the family of solutions for a and 0. For (5, x) G [0, 00) X R** and s
;) such that l/(n + 1) < t — s < 1/n or n < r — s < n + 1. It is easy to check that p(s,x;t,y) is the desired density. From (1.44) it is immediate that 0, depending only on d, dj (), Ar, and A7, such that for all (s, x) € [0,T)xR'',\h\ < p, and bounded measurable / having support in [0, T] x Q(x*',p) where QCx^.p) = {x:x° - p < Xi < x^ + p,l
U a B < co, and an cc> 0 such that I \F(x\y)-F(x\y)\'^dy\ .,„[T„ > T] \xeR
and for small h> 0, x(t -\- h) — x(t) is like the Gaussian independent increment process having mean b(t, x(t)) and covariance a(r, x(t)). (A slightly different presentation of these ideas is given in the introduction to Ito [1951]. We recommend Ito's discussion to the interested reader.) The structure of this book can now be explained in terms of the ideas introduced in the preceding paragraph. Starting from (0.1), various tacks toward an understanding of the process x(*) suggest themselves. The most analytic of these is the following. Let P(s, x;r, •) denote the transition probability function determined by x(*) (i.e., P{s, x; r, F) = P(x(t) e F |x(5) = x)). From (0.1), we see that — [ P(s, x; r, dy)(p(y) = lim f P(s, x; f, dz) ^^•'
hio -^
X I P(r, z;t-i-K dy)((p(y) - (p(z)) == j P(s, x; t, dy)LMyy Of course, we have used the Chapman-Kohnogorov equation. From this we derive the formal relation: (0.3)
- P ( 5 , x; r, •) = LTP(s. x;t, *),
t > s,
where L,* is the formal adjoint of L,. Equation (0.3) is called ihQ forward equation (in physics and engineering literature it is often referred to as the Fokker-Planck equation). Since it is clear that (0.3')
lim P(s,x;r, ') = S,('% tis
0. Introduction
3
it is reasonable to suppose that one can recapture P(s, x; t, •) from (0.3) and (0.3'). Indeed, this was done with great success by Kolmogorov [1931] and Feller [1936] in their pioneering work on this subject. However, there are severe technical problems with (0.3). In particular, one must tacitly assume that P(s, x; t, •) admits a density p(s, x; t, y) and think of (0.3) as being an equation for p(s, x; r, y) as a function of t and y; and even when such an assumption is justified, there remain inherent difficulties in the interpretation of Lf unless the coefficients are smooth. For this reason, people turned their attention to the backward equation. Namely, starting once again from (0.1) we have
- :^ (^' ^) = - ^ I ^(^' ^; ^' ^yMy) 11= lim - P(s — h, x; s, dy)(u(s, y) — M(S, X)) hio
f^^
= L,u(s, x\ where u(s, x) = J P(s, x; r, dy)(p{y\ 0 < s < t. (Notice that the preceding computation is not fully justified since we do not know that u(s, •) is in our test-function space. Nonetheless, the argument is correct in spirit.) Hence we arrive at (0.4)
— P(s, x\u •) + ^sP{s. x; r, •) = 0,
(0.4')
limP(5, x;r, •) = M-)-
0 < s < r,
(It should be clear why (0.3) is the forward equation and (0.4) is the backward equation: (0.3) involves the forward (i.e., future) variables whereas (0.4) involves the backward (i.e., past) variables.) Again one might suspect that (0.4) and (0.4') determine F(s, x; r, *) and now there are no problems about mterpretation. The study of diffusion theory via the backward equation has been one of the more powerful and successful approaches to the subject and we have included a sketch of this procedure in Chapters 2 and 3. The major objection to the study of diffusion theory by the method just described is that the hard machinery used comes from the theory of partial differential equations and the probabilistic input is relatively small. A more probabilistically satisfactory approach was suggested by Levy and carried out by Ito [1951]. The idea here is to return to the intuitive picture of x(f -\- h) — x(t\ for small /i > 0, looking like the Gaussian independent increment process with drift b(u ^(0) ^^^ covariance a(r, x(t)). In differential form, this intuitive picture means that (0.5)
dx(t) = G(U x(t)) dp(t) + b(u x(t)) dt
where j?( •) is a ^/-dimensional Brownian motion and a is a square root of a. Indeed, G(U x{t))(P(t -\- h) - P(t)) + b(t, x(t)) will be just such a Gaussian process;
4
0. Introduction
and if {x(s), 0 < s < r} is {P(s): 0 < s < r}-measurable, then cr(r, x{t)) x (p(t + h)- P(t)) + b(t, x(t)) will be conditionally independent of {x(5): 0 < s < r} given x(t). There are two problems of considerable technical magnitude raised by (0,5). First and foremost is the question of interpretation. Since a Brownian path is nowhere differentiable it is by no means obvious that sense can be made out of a differential equation like (0.5). Secondly, even if one knows what (0.5) means, one still has to learn how to solve such an equation before it can be considered to be useful. Both these problems were masterfully handled by Ito, a measure of the success of his solution is the extent to which it is still used. We develop Ito's theory of stochastic integration in Chapter 4 and apply it to equations like (0.5) in Chapter 5. With Chapter 6 we begin the study of diffusion theory along the lines initiated by us in Stroock and Varadhan [1969]. In order to understand this approach, we return once again to (0.1). From (0.1), it is easy to deduce that: E[(p(x(t2))\x(sls
= lim -E[E[(p(x(t2 + h)) - (p(x(t2))\x(s), s < t2]\x(sl s < rj hiO
^
Thus .'2
E[(p(x(t2)) - (p(x(ti)) -
LM^t)) dt\x(sl s < r j = 0;
•'fi
or in other words (0.6)
X^(t)^(pix{t))-
(LMx(s))ds
is a martingale for all test functions (p. (Notice that the line of reasoning leading from (0.1) to (0.6) is essentially the same as that from (0.1) to the forward equation.) One can now ask if the property that X^(*) is a martingale for all test functions (p uniquely characterizes the process x(*) apart from specifying x(Q). To be more precise, given L^, consider the following problems: (i) Is there for each x e R*^ a. probability measure P on C([0, oo), R'^) such that P(x(0) = x) = 1 and X^(') is a martingale for all test functions (p? and (a) Is there at most one such P for each x? Problems (i) and (ii) constitute what we call the martingale problem for L,. Of course problems (i) and (n) are interesting only if one can also answer (in) If (i) and (ii) have affirmative answers, what conclusions can be drawn? To convince oneself that these are reasonable questions, one should recall that in the case when d = I and L, = jd^/dx^, Levy (cf. Doob [1953] or Exercise 4.6.6) characterized Wiener measure as the unique probability measure P on
0. Introduction
5
C([0, oo), i?^) such that P(x(0) = 0) = 1 and x(t) and x^(t) - t are martingales. That is, he showed that in this case one only needs the functions (p(x) = x and il/(x) = x^. (Actually [cf. Exercise 4.6.6], this is a general phenomenon, since under general conditions one can show that X^( •) is a martingale for all test functions (p if X^^(-) and X^.^.(*) are martingales for (pj(x) = Xj and \l/ij(x) = x,Xj, 1
(0.7)
mf
0 < s < 7 8eR''\{0}
rjr-2
>0
|^|
\x\
and (0.8)
lim
sup
5iO
o<s
||a(s, x^) - a(s, x^)\\ = 0,
then the martingale problem for L, is well-posed (i.e., existence and uniqueness hold). As a dividend of our proof, we show that L, determines a strong Markov, strongly Feller continuous process. The contents of Chapter 8 are somewhat tangential to the main thrust of our development. What we do there is expand on the theme initiated in Watanabe and Yamada [1971] in their investigation of the relationship between Ito's approach and the martingale problem. In Chapter 9 we return to L,'s having coefficients of the sort studied in Chapter 7. Here we take advantage of certain analytic relations and estimates upon which our proof of uniqueness in Chapter 7 turns. In brief, the results of these considerations are various L^-estimates for the transition probability function of the process determined by L,. Chapter 10 extends the martingale problem approach to unbounded coefficients. The point made here is that this extension is elementary, provided
6
0. Introduction
one can show that the diffusion process does not "explode." We give some standard conditions that can be used to test for explosion. Again in Chapter 11 we deal with L/s of the sort studied in Chapters 7 and 10. This time we are interested in stability results for the associated processes. These results can be naturally divided into two categories: convergence of Markov chains to diffusions (i.e., invariance principles of the sort initiated by Erdos and Kac and perfected by Donsker) and convergence of diffusions to other diffusions. Both categories are surprisingly easy to handle given the results of Chapters 7 and 9. The final chapter, Chapter 12, takes up the question of what can be done in those circumstances when existence of solutions to a martingale problem can be proved but uniqueness cannot. The idea here, is to make a careful " selection " of solutions so that they fit together into a Markov family. The procedure that we use goes back to Krylov [1973]. We also show in Chapter 12 that every solution to a given martingale problem can, in some sense, be built out of those solutions which are part of a Markov family. The only parts of the book which we have not yet discussed are the beginning and the end. Chapter 1 provides an introduction to those parts of measure and probability theories which we consider most important for an understanding of this book. Although the material here is not new, much of it has been reworked. In particular, our criteria for compactness in Section 1.4 strikes us as a useful variation on the ideas of Prohorov. Finally, in spite of our attempt to make it look as if it were, the appendix is not probabihty theory. Instead, it is that part of the theory of singular integrals on which we rely in Chapters 7 and 9. At the present time, one has to depend on these results from outside probabiHty theory and we have provided a proof in the Appendix in order to make the book self-contained. It is now time for us to thank the many people and organizations to whom we are deeply indebted. The original work out of which this book grew was performed while both of us were at the Courant Institute of Mathematical Sciences. During that period we were encouraged and stimulated by many people, particularly: M. Kac, H. P. McKean Jr., S. Sawyer, M. D. Donsker and L. Nirenberg; and we were supported by grants from the Air Force, the Sloan and Ford foundations as well as general C.I.M.S. funds. Whether this book would ever see the light of day was cast into considerable doubt by the departure from C.I.M.S. of one of us to the Rock'y Mountains in 1972. At that time not a sentence of it had been written. However, in 1976 we had the good fortune to visit Paris together under the auspices of Professors Neveu and Revuz; and it was at that time (much to the dismay of an accompanying wife) that we actually began to write this book. Progress from that point on has been slow but steady. During the interim we have incurred a considerable debt of gratitude to several people: wives Lucy and Vasu; secretaries Janice Morgenstern, Gloria Lee, Susan Parris and Helen Samoraj; students Marty Day and Pedro Echeverria; colleagues Richard Holley, G. Papanicolaou, M. D. Donsker, and E. Fabes; gadfly J. Doob, and pubHsher Springer Verlag. To all these we extend our heart felt thanks along with the promise that they do not necessarily have to read what we have written.
Chapter 1
Preliminary Material: Extension Theorems, Martingales, and Compactness
1.0. Introduction As mentioned in the Introduction, the point of view that we take will involve us in a detailed study of measures on function spaces. There are a few basic tools which are necessary for the construction of such measures. The purpose of this chapter is to develop these tools. In the process, we will introduce some notions (e.g., conditioning and martingales) which will play an important role in what follows. Section 1.1 contains the basic theorem of Prohorov and Varadarajan characterizing weakly compact families of measures on a Pohsh space. Using their results, we prove the existence of conditional probability distributions. The final topics in Section 1.1 are the extension theorems of Tulcea and Kolmogorov. Section 1.2 introduces the notions of progressively measurable functions and martingales. In connection with martingales we prove Doob's inequality, his stopping time theorem and a useful integration by parts formula. Finally we prove a result connecting martingale theory and conditioning. In Section 1.3 we specialize the results of Section 1.1 to the case when our Polish space is C([0, oo); R^) (i.e., the space of K*^ valued continuous functions on [0, 00) with the natural topology induced by uniform convergence on bounded intervals). Section 1.4 contains a useful sufficient condition for compactness of a family of measures on C([0, oo); R*^) in terms of certain martingales associated with them.
1.1. Weak Convergence, Conditional Probability Distributions, and Extension Theorems Throughout this section (X, D) will stand for a Polish space (i.e., a complete separable metric space) and ^ = Mx its Borel a-field. We denote by M(X) the set of all probabihty measures on {X, 0t) and by Q(X) the set of all bounded continuous functions on X. We will view M{X) as a subset of the dual space of C^pi) and give it the inherited weak* topology. It will turn out that this topology makes M(Z) into a metric space. 1.1.1 Theorem. Let ^„ e M{X)for each n> I. Given fi e M(X), the following are equivalent:
8
1. Preliminary Material: Extension Theorems, Martingales, and Compactness
(i) lim„^^ ^ fdii„ = ^ fdfifor every fe Q(A') (a) lim„^^ I fd^„ = ] fdfifor every fe Up(X) where Up(X) is the set of bounded uniformly continuous functions on (X, p) and p is any equivalent metric on X. (Hi) lim sup„_^oo Mn(^) ^ f^(C)for any closed set C in X (iv) lim inf„^oo/i„(G) > fi(G)for any open set G in X (v) lim„^<„ fi„(B) = fi(B)for any B e ^ such that p(dB) = 0. Proof That (/) implies (ii) is obvious, and {Hi) is equivalent to (iv) by complementation. Also the following simple argument shows that (Hi) and (iv) together imply (.): lim sup fi„(B) < lim sup p„(B) < p(B) = p(B^) < lim inf p„(B^) < lim ini fi„(B). It remains to show that (ii) implies (Hi) and (v) implies (i). To prove that (ii) implies (Hi), let C be a closed set and choose
/.w=
1 + P(x. C)
k> 1
where p(x, C) = 'mf{p(x, y): ye C}. Then /^ e Up(X) for each k > 1 and fk(x) I xc(x) for each xe X. Therefore p(C) = lim j fkdfi= lim lim j /^ dp„ > lim sup p„(C). fc-^CC «-»QO
k-*co
Finally, to see that (v) implies (i), take/e Q(X) and, given e > 0, choose N and {ai}^~ ^ so that -l = ao
< £.
Hence, lim sup <2s + lim sup | j | j ; a.ZB, MM„ - fl E a.XB, UA^ < 2£ + X l«il lim sup |/i„(B,) - /i(B,)| 1
n-* 00
<2£.
Since e > 0 was arbitrary, the proof is complete. Q
1.1. Weak Convergence, Conditional Probability Distributions, and Extension Theorems
9
Remark. The equivalence of (i) and (ii) implies that for any X, fi e M(X): if I fdix = j fdX f o r / e Up(X), then ^ fd^ = ^ fdXfov all functions/in Q ( X ) and in fact X = ji on ^. Since X is a separable metric space, by TychonofPs embedding theorem, X is homeomorphic to a subset of a compact metric space. Thus X admits an equivalent metric p with respect to which it is totally bounded. Choose such a p and let X denote the completion of (X, p). Then X is compact and U^{X) is isomorphic to C(X), which is separable. With these remarks we will prove that the weak* topology on M(X) is metrizable. In fact, define for ^, Xe M(X)
where {cp,,: /c > 0} is dense in Up(X). By the remark following Theorem 1.1.1, A is clearly a metric on M(X). 1.1.2 Theorem. If A is defined on M(X) by (1.1), then the topology induced on M(X) by A is the weak* topology. Proof. Obviously the topology induced by A is weaker than the weak* topology. We now show that for any p e M(X) and any weak* neighborhood N of yu, there is a ^ > 0 such that
{X:A(lp)<S}^N. If not, there is a sequence {>l„}f ^ M(X) such that A(X„, //) -• 0 as n -• oo and yet X„ ^ N for any « > 1. On the other hand, from the definition of A in (1.1), the denseness of {(p,,: k > 0}, and the equivalence of (/) and (ii) of Theorem 1.1.1, it would follow that
lim \fdX„ = n-*
00
\fdp
*•
for a l l / 6 Q(X). Since N must contain a set of the form /I: max fjdX-\fjdp
< &.
for some choice of g, / and {f^\ ^ Q(X), we have a contradiction.
D
Remark. With a little more work it is possible to show that there is a metric A on M(X\ equivalent to A, such that (M{X\ S) is a Polish space. We will not be needing this fact in what follows. 1.1.3 Theorem. Let T be a compact subset of M(X). Then for each e > 0 there is a compact K ^ X such that p(X\K) < sfor all p e F. In particular for any p e M(X) and e > 0, there is a compact K ^ X such that p(X\K) < s.
10
1. Preliminary Material: Extension Theorems, Martingales, and Compactness
Proof. Let {Xj-. j> 1} be dense in X, and for /c > 1 and n > 1 put
OI-(J^B(.,1). By Theorem 1.1.1, the map fi -• ^(GJJ) is lower semicontinuous, and clearly
as n -• 00 for each k. Since T is compact, it follows from Dini's theorem that for each e > 0 and A; > 1, there is an n^ such that inf/z(GJ^)>l-^. S e t K = f]^^, GJ*. Then inf/i(X)> l - £ . Moreover, K is closed in X and is therefore complete. Finally, for any /c > 1
and so K is totally bounded. Thus K is compact and the proof is finished. Q 1.1.4 Theorem. Let T ^ M(X) be given and assume that for each e> 0 there is a compact set K '^ X such that inf ^(K)
>l-e.
Then F is precompact in M(X) {i.e., T is compact). Proof We first recall that if X itself is compact, then by Riesz's theorem and standard results in elementary functional analysis, M(X) is compact. In general we proceed as follows. Choose a metric p on X equivalent to the original one such that (X, p) is totally bounded and denote by X its completion. Then X is compact and we can think of M(A') as being a subset of M(X). Thus it remains to show that if {/i„}f ^ r and n„-* p in M(X\ then p. can be restricted to A" as a probabihty measure //, and /i„ -• /i in M(X). But, by our assumption on F, there is a sequence of compact sets {X,}, / > 1 in X such that p„(Ki) > 1 — 1// for « > 1. Since K, is compact and therefore closed in X, it follows from Theorem 1.1.1 that p(Ki) > \im„^^p„(Ki)>\l//.Thus
4'JK.) =
1. Weak Convergence, Conditional Probability Distributions, and Extension Theorems
11
and so we can restrict fi to (X, ^x) ^s a probability measure //. Finally since /i„ -^ /i in
M(X) lim ^(pd^i„ =
^q)d^
for all (p e Up{X\ and by Theorem 1.1.1 this implies that ^„ -• /i in M(X). Remark. Note that whereas Theorem 1.1.3 relies very heavily on the completeness of X, Theorem 1.1.4 does not use it at all. 1.1.5 Corollary. Let F ^ Cj,(X) be a uniformly bounded set offunctions which are equicontinuous at each point ofX. Given a sequence {//„}f ^ M(X) such that /!„-•// in M(X\ one has lim sup U (pdn„-
j (p dy,
0.
n-*ao 0 e F
Proof Suppose there exists £ > 0 such that lim sup sup \ (P dn„- ^ (p dfi > e. By choosing a subsequence if necessary, we can assume that for each n there is a (p, in F such that I (Pn dfi„ - j (p„dn > £ > 0 . Let M = sup^gf \\(p\\ and choose a compact set K in X such that sup„>i fi„(X\K)<s/SM. It follows from this that fi(X\K) <E/SM. From the Ascoli-Arzela theorem and the Tietze extension theorem, we can find snj/ in Ci,(X) such that I I/A I < M, and a subsequence (p^. = ij/j of{(p„} converging to ij/ uniformly on K. But then J ijjj d^„. - I il/j ^/i| < |J (ij/j - il/) dfij + J (il/j - iP) d^ J ^ dfi„. - ^ip dfi = \\(iljj-il/)d^J-h\\
(il/j-il/)d^„. X\K
i^j - ^) dfi ''X\K
+
J ^ dfi„. - J ij/ dfil
12
1. Preliminary Material: Extension Theorems, Martingales, and Compactness
Therefore I 0 < £ < lim I lAj d^nj - j il/j dfi\ <0 + 2 M . ^ + 0 + 2 M . ^ + 0
which is a contradiction.
Q
We now turn to the study of conditional probabihty distributions. Let (£, ^, P) be a probability space and I e J^ a sub cr-field. Then the conditional expectation of an integrable function / ( ) is another integrable function g(•) which is Z-measurable and satisfies: (1.2)
j g{q)P{dq) = j f{q)P(dq) A
for all A e Z.
A
The function g(') exists and is unique in the sense that any two choices agree almost surely with respect to P. (See Halmos [1950], for instance, to find a proof of the existence and elementary properties of conditional expectations.) The function g is denoted by E[f \ I ] . If we want to call attention to the measure P that is used, we use E^[f \ Z] in place of E[f \ L]. In the special case when the function / ( • ) is the indicator function XBIQ) of a set B in J^ we refer to the conditional expectation as the conditional probability and denote it by P ( B | I ) . It has some elementary properties inherited from the properties of conditional expectations. For instance if Bj and B2 are disjoint sets in J^ P(B, u B211) = P(B, 11) 4- P(B211)
a.s.
Since P(B \ I ) can be altered on a set of measure zero for each B e J*^ it is important to know if one can choose a " nice " version of P(B | L) such that P(B | Z) is a countably additive probability measure on J^ for each q e E. Such a choice, if it can be made, will be called a conditional probability distribution and will be denoted by {Q,{B)}. Definition. A conditional probability distribution of P given L is a family Q^ of probability measures on (£, .J^) indexed by q e E such that ii) For each B e .^, Qq(B) is I-measurable as a function of q (a) For every A ET and B e ^ P(A n B) = I
QmP(dq).
1.1. Weak Convergence, Conditional Probability Distributions, and Extension Theorems
13
In general a conditional probability distribution need not exist. However if we replace (£, J*^) by a Polish space X and its Borel cr-field ^ , then for any P on (X, iM) and any sub cr-field Z ^ ^ a conditional probabihty distribution of P given Z exists. We state and prove this as a theorem. 1.1.6 Theorem. Let X bea Polish space and 3 its Borel o-field. Let P he a probability distribution on (X, ^ ) and Z ^ ^ a sub a-field ofM. Then a conditional probability distribution {Q^} of P given I always exists. Proof. Since Z is a Polish space, there exists on X an equivalent metric p such that (X, p) is totally bounded. Therefore the space Up(X) of uniformly continuous bounded functions on (X, p) is separable. Let {fj}f be a countable subset of Up(X) such that/i(-) = 1, {fj]f are linearly independent and the linear span W of{fj}f is dense in Up(X). We denote by ^, some version of £ [ / | Z]. We can assume, without loss of generality, that ^i(*) = 1. For n > 1 let A„ be the set of n-tuples ( r i , r 2 , . . . , r„) with rational entries such that ^ i / i ( ^ ) + •• + ^ / „ ( ^ ) > 0
for all
xeX.
The set A„ is clearly countable. From the properties of conditional expectations, ^1 ^i(^) + r2g2(x) + • • • + r„g„(x) > 0 for almost all x. Thus if we let, for each ( r , , . . . , r„) e A„, ^(ri,...,r„) = {x- riOiM + • • • + r„g„(x) < 0} then F(^j
^^) e L and P[F^r^
^ J = 0. Therefore if
F=0
f(M
U
r
Then F G I and P{F) = 0. Choose X e X\F and fix it. We now define the linear functional L^ on W by L,(f) = tigi(x)-^''
+ t„g,(x)
w h e r e / e W is written uniquely as /=fi/i-f
••• + ?„/„
for some n and real numbers fj, (2, . •, ?„• We want to show now that L^{f) is a nonnegative linear functional on W. Suppose feWis non-negative. Then hfi + • + fn /n = / ^ 0- Given any rational number e > 0 we can find rationals (ri, ..., r„) such that rJi
+'••
+ rj„>
-E=
-£/i
14
1. Preliminary Material: Extension Theorems, Martingales, and Compactness
and Irj — tj\ < e for i <j < n. Therefore (ri+a)/i + - - + r „ / „ > 0 . From the definition of the set F, it follows that (r,+s)g,(x)-^''-^r„g„(x)>0. By letting e -• 0 over the rationals, we conclude that ti9i(x) +
--+t„g„(x)>0,
or equivalently Lj,(/) > 0. Since L;,(/i)= 1 and W is dense in Up(X), Lj^f) defined on W extends uniquely as a non-negative linear functional on U^{X). We continue to call this extension Lj^f). We can view the space Vp{X) as the space C(X\ where X is the completion of the totally bounded space (X, p). Note that X is compact. By the Riesz Representation Theorem, there is a probabihty measure Q^ on (X, edx) such that
Uf)=\f{y)QMy) for a l l / e C(X). Thus we have shown that for all x e X\F, there is a probability measure Q^ on (X, % ) such that (1.3)
gi(x) = f fi(y)QMy)
for all i. (Here we use the notation/to denote the extension of a n / e Up(X) to X.) This shows that the mapping
f(y)QMy) on X\F is Z[X\F]-measurable for a l l / e W, and therefore for a l l / e (7p(X). Moreover, it is easy to see that 1.4)
j ny)QMy), A n {X\F) = £l/(-M]
for all/e U^{X) and ^ G I . Given a compact set K in X, choose {
E'lQ.iK), X\F] = P{K).
1.1. Weak Convergence, Conditional Probability Distributions, and Extension Theorems
15
Next choose compacts K„ in X so that K„^ K„+i and P(K„) > 1 — (1/n). Then D=[jf K„ism^^ and, from (1.5), £^[e.(D), X\F] = 1. Thus there is a Z[X\F]-measurable P-null set F' such that Q^(D) = 1
(1.6)
for xe X\(F u F).
In other words, for x e X\(F KJ F'), Q^ can be considered as a probabiUty measure e^ on (Z, ^;^). For X e F u F , define Q^ = P. Then
f f(y)QMy) is a X-measurable mapping on X for all/e Up(X). Moreover, by (1.4) and the fact that P(F u F ) = 0, we have f/(y)e.(^yM|=£l/(-M],
AeZ,
for a l l / e (7^(X). Thus (1.7)
\f{y)Q.(dy)==E[f\:L]
a.s.
for a l l / e C/p(X). Since the class of bounded i^;^^-measurable/for which
x-^\f(y)QMy) is Z-measurable and (1.7) holds is linear and is closed under bounded point-wise convergence, it follows from Exercise 1.5.2 that this class coincides with the bounded ^;^-measurable functions. In particular, Q.(B) is Z-measurable and is equal to P(B \ 1) a.s. for all B e ^x- This completes the proof that {Q^ is a conditional probability distribution. D One can now use the conditional probability distribution to construct a version of the conditional expectation for any integrable function/(•) on {X, ^ , P). 1.1.7 Corollary. Letf(') be any integrable function on (X, ^, P). Then for almost all X, I I f(y) I QMy) < «
and j f(ymdy)
= E[f 11]
a.s.
16
1. Preliminary Material: Extension Theorems, Martingales, and Compactness
Proof. Clearly the class of non-negative functions for which (1.7) holds is a monotone class containing bounded measurable functions on (X, ^). A routine appHcation of the monotone convergence theorem yields Corollary 1.1.7 for non-negative integrable functions. By separating any integrable function into its positive and negative parts, we can complete the proof of the corollary. D If the sub c-field Z is countably generated, then we can say something more about the conditional probability distribution. If x e A" is any point we define the atom A{x) containing x by A{x) = n {A:xe A, A e l } . It follows from Exercise 1.5.3 that A(x) is in fact an element of the cr-field Z. 1.1.8 Theorem. // Z is countably generated then there is a P-null set N (i.e., P{N) = 0) in I such that for x e X\N
e.(^W)=i. Proof If /I e I , then = P(M^)
QM)
a.s.
= XA(X)
If ZQ is a countable field generating I , then outside a single null set N e Z (1.8)
e.(^) = Z.W
for all
Aei:„.
Since both sides of (1.8) are probability measures, it follows from (1.8) that Q^(A) = XA(X) for X e X\N
and A el..
In particular, if /I = A(x\ then for x e X\N
QAMx)) = XAU^) = 1This completes the proof. D A conditional probability distribution satisfying Qx(A(x)) = 1 a.s. is called a regular conditional probability distribution. Let (£, J^) be a measurable space and J*^„, « > 0, an increasing family of sub cr-fields of J^ such that ^ is generated by |J„ . ^ „ . If we have a consistent family {Pn} of probabihty measures on (£, J^,) (i.e., Pn-i = Pn \ ^n-\), under suitable conditions one can obtain a probabiHty measure on {E,^). The problem of course is to show that F, which can be defined naturally on U« ^n, is countably additive on it, and therefore can be extended uniquely to ^. We need a basic assumption
1.1. Weak Convergence, Conditional Probability Distributions, and Extension Theorems
on the nature of the a-fields #"„ c ^ . For each q^E K{^) = n {B: Be J^„ and
17
and n we define qe B}.
We make the following hypothesis: N
(1,9)
For every sequence {q„}f ^ E such that f] A„(q„) i= 0 M= 0 00
for every AT, we have f] A„(q„) ^ 0 . n=0
If E is a product space of the form E = Y\r= i ^i ^^^cl #^„ is the a-field generated by the first n coordinates, the condition is always satisfied. On the other hand, if E = C[0, 1], the space of continuous functions on [0, 1], and ^„ is the a-field generated by the path on [0, 1 — (1/n)], then the condition is not fulfilled. This is because a consistent sequence of continuous functions on [0, 1 — (1/n)] determine a continuous function on [0, 1), but not necessarily on [0, 1]. The first theorem we prove is a variation of Tulcea's extension theorem. It is particularly well suited for Markov processes. 1.1.9 Theorem Let (E,^) be a measurable space and {#'„ :n>0} a non-decreasing sequence of sub o-algebras whose union generates ^ . Let PQ be a probability measure on (£,#0), and, for each n > 1, let n'^(q,dq') be a transition function from {E.^n-i) to (£,#'„). Define Pn on (E,J^ri)for n>l so that Pn{B) = Jn^q,B)Pn-i(dq)
for
B G ^nl
and assume that, for each n>0, there is a Pn-null set Nn G J%, such that q^Nn.
Be #-„+!, and 5 n An(q) = 0 => n"+\q,B) = 0.
Then there is a unique probability measure P on (E,^) P I ^Q= PQ and, for each n>l, PiB) = fn"{q,B)P(dq')
for
with the properties that
B G J^.
For each n, F„ is probability measure on ^„ and P„+i agrees with P„ on J^„. Hence there is a unique finitely additive P on Q„ ^„ such that P agrees with P„ on ^ „ . We will show that P is countably additive on |J„ J^„ and therefore extends uniquely as a probabihty measure to J^. We have to show that if B„ e ij„ J^„ and J5„i0 then P(B„)iO. Assume that P(B„) > e > 0 for all n. We will produce^a point qef]„B„. We can assume, without loss of generality, that B „ G J ^ „ . For 0 < M < m and B G J^„ we define TT'"- "(q, B) to be XB(Q)- For n > m and B e J^„ we define TT'"' "(q, B) inductively by n"''"{q,B)=
\n"{q',B)n'"'"-'(q,dq).
18
1. Preliminary Material: Extension Theorems, Martingales, and Compactness
Clearly P(B)==jn'''"(q,B)Po(dq)
for
B e ^„.
We also have for n> m 7r'"'% B) = j 71^"^^'V. B)7f"^'(q, dq') for Be ^ „ . Define F« = L : 7 r « " ' t e 5 „ ) > |
n>0.
Then F^+ j ^ F^ and a
F;;'+^/
^
F;;'^
^ and for n > m TT < ""• %\ i m „ ,' B„) < ; ^ s ^ + rr*'(9», - ^ r ' ) —n/ — y
om+1 —
Hence n-'^'l^qm,
nFrO>£/2'"^^
We can therefore conclude that A„(q„) n (~]Q F ^ ^ ^ =/= 0 ; and so there is a qn+i € An(qn)\Nn with the property that 7r'"+^'"(^;„+i,5„) > £/2'"+2 for all n > 0. By induction on m, we have now shown that a sequence {qmjo' exists with the properties that qm+i e ^m(^m)\^m and inf„7c'"'"(g^,B„) > 0 for all n > 0. In particular XBjqm) = n"'''^(qm,Bm) > 0 and therefore Am(qm) <= B Yn^ S i n c e Byn € Thus rim^/ntem) CI ^^Bm. Finally, qN e nS'^m(^m) for all AT > 0, and so (1.9) imphes that Ho^ Amiqm) i^ 0- This completes the proof. D We will now establish Kohnogorov's extension theorem for product spaces. Let / be an infinite index set and for each a e / let X^ be a Polish space with its Borel
^m-
1.2. Martingales
19
c-field ^^. For sets F c: / we denote by Xp the product space Ha e F ^ a ^^^ by ^p the product (T-field Ha e F ^a oi^ ^F • For sets G ZD F i=^ 0 let (T^ denote the canonical projection from XQ onto X;r. We denote o\, by Op. 1.1.10 Theorem. Suppose that for each finite set F we are given a probability measure Pp on (Xp, ^p) such that, for any two finite sets 0 =^ F c:G,Pp = PG(<^F)~ ^- Then there is a unique probability measure P on {Xj, ^j) such that Pp — Pap^ for all finite Fi= 0. Proof Uniqueness is obvious from the fact that ^j is the smallest a-field generated by s^ = [jp (Tp^(^p). To prove existence, we observe that there exists a finitely additive P on s/ such that Pp = Pcp ^ Now suppose that {A^Q 6 J^/ is a nondecreasing sequence such that A„ 1 0 . Without loss of generality, we will assume that A^ e ^p^, n>0 where 0 =f^ FQ a F^ ••' a F„-' and {F„} are strictly increasing finite sets. Define ^„ = or^/(^pJ for n>0. It is easy to check that {J^„, n>0} satisfy (1.9). Next, let {Q"} be a regular conditional probability distribution of Pp^ given (ap"^ _ i) ~ ^ (^F„ _ i) ^^^ define
for all qe Xj and J3 e J^„. It is easily checked that TT" is a transition function from {Xi, ^n-i) to (Xi, J^„) and that it satisfies the condition of Theorem 1.1.9. Thus by that theorem there is a unique probability measure P on {Xj, (T(IJS ^n)) such that P equals Pp^ (jp^ on ^Q and P{B) =
\n%q,B)P{dq)
for all B e J^„. By induction, we see that P equals P on (Jj* J^„. In particular P(A„) = P(A„\ the countable additivity of P implies that P(A„)[0, and the theorem is proved. Q
1.2. Martingales Throughout this section, E will denote a non-empty set of points g. J^ is a c-algebra of subsets of £, and {J«^,: r > 0} is a non-decreasing family of sub aalgebras of ^. Given s > 0 and a map ^ on [s, oo) x £ into some separable metric space (X, D), we will say that 9 is (right-) continuous if ^( •, ^) is (right-) continuous for all ^ e £. If P is a probability measure on (£, ^) and 0: [s, oo) x £ -• X, we will say that 0 is P-almost surely [right-) continuous if there is a P-null set N e ^ such that 9{', q) is (right-) continuous for q 4 N. Given s > 0 and 0 on [s, oo) x £ into a measureable space, 6 is said to be progressively measurable with respect to {J^,: f > 0} a/rer ^ime 5 if for each t > s the restriction of 9 to [s, t] x E is ^[j, f] ^ ^^rn^^asurable. Usually there is no need to
20
1. Preliminary Material: Extension Theorems, Martingales, and Compactness
mention 5 or {^^i t > 0}, and one simply says that 6 is progressively measurable. Note that 9 is progressively measurable with respect to {J^,: t > 0} after time s if and only if 6^, defined by 6^(1, q) = 6{s -I- r, q), is progressively measurable with respect to {^t+s'- ^ ^ ^] ^^er time 0. Thus any statement about a progressively measurable function after time s can be reduced to one about a progressively measurable function after time 0. This remark makes it possible to provide proofs of statements under the assumption that s = 0, even though s need not be zero in the statement itself. Exercises 1.5.11-1.5.13 deal with progressive measurabihty and the reader should work them out. The following lemma is often useful. 1.2.1 Lemma. IfO is a right-continuous function from [s, 00) x £ into (X, D) and if 6(ty ') is ^t-measurable for all t > s, then 6 is progressively measurable. Proof Assume that 5 = 0. Let r > 0 be given and for w > 1 define e„(u,q) = e\^^
At.qj.
Clearly e„ is ^[0, t] ^ ^f-measurable for all n. Moreover, as « -^ 00, 6„ tends to 6 on [0, t] X E. Hence 6 restricted to [0, r] x £ is ^[o, t] ^ ^r^QSiSurable. n Given a probability measure P on (£, ^), 5 > 0, and a function ^ on [s, 00) x £ into C, we will say that (9(t), ^,, P) is a martingale after time 5 if 0 is a progressively measurable, P-almost surely right-continuous function such that 6{t) = 0(t, ') is P-integrable for all t > s and (2.1)
£1^(^2)1^, J = ^(^1)
(a.s., F),
s
The triple (6(t), ^^, P) is called a submartingale after time s if ^ is a real-valued, progressively measurable, P-almost surely right continuous function such that 0(t) is P-integrable for all r > 5 and (2.2)
E'[e(t2) IJ^,J > 0(t,)
(a.s., P),
s
In keeping with the remarks following the definition of progressive measurabihty, we point out that s plays a rather trivial role here and that any statement proved for the case s = 0 can be proved in general simply by replacing 6 by 0^ and {^f-. r > 0} by {^t+s- t > 0}. Thus, although theorems will be stated for general s, they will be proved under the assumption that s = 0. Usually, it will not even be necessary to mention when the (sub-) martingale begins and we will simply say the (6(t), ^^, P) is a (sub-) martingale. We begin our study of martingales and submartingales with the following lemma. Like nearly everything in this theory, the original version is due to J. L. Doob. Thus we will be somewhat lax in the assignment of credit before each theorem.
1.2. Martingales
21
1.2.2 Lemma. Let (6(t\ ^^, P) be a submartingale after time s with values in a closed interval / £ R. Assume that g is a continuous, non-decreasing, convex function on I into [0, oo) such that g o e(t) is P-integrable for all t > s. Then (g o 0(t), ^^, P) is a submartingale after time s. In particular, if (9(t), ^^, P) is a martingale or a non-negative submartingale after time s and r is a number greater than or equal one such that 16(t) [ is P-integrable for all t > s, then (| B(i) |^ ^^, P) is a submartingale after time s. Proof Assume that s = 0. By the version of Jenssen's inequality for conditional expectation values: E[g(0(h))\^,,]>g(E[e(t^)\^„])
a.s.
Thus, if 0 < f 1 < t2, then E[g(e(h)) I J^.J > g(E[e(t,) IJ^,J) > # ( t O )
a.s.
This completes the proof of the first assertion. The second assertion is immediate in the case when {9(t\ ^ t , P ) is a non-negative submartingale; simply take g(x) = x' on [0, oo). Thus the proof will be complete once we show that (|^(01» ^t, P) is a submartingale if (0(t\ J^,, P) is a martingale. But if (9(t\ J^,, P) is a martingale, then
E[ I ^(^2) I I i ^ J > I E[e(t,) I ^ J I = I d(t,) I and so (10(t) |, i^,, P) is a submartingale.
(a.s., P),
D
1.2.3 Theorem. If(6(t\ <^,, P) is a submartingale after time s, then for any A > 0 and allT>s: (2.3)
p( sup
e(t)>A<\E e(T),
\s
I
X
sup e(t) > X s
In particular, if 9 is non-negative, then
(2.4)
pi sup 0{t)>x)<-E[0(T)l \s
J
^
and for all r> 1 nl/r
(2.5)
E J f sup^^W^
<
r
-E[9iTy]
1/,
r—1 (in the sense that the right-hand side is infinite if the left-hand side is).
22
1. Preliminary Material: Extension Theorems, Martingales, and Compactness
Proof. Assume that 5 = 0. Relation (2.4) is immediate from (2.3) and (2.5) follows from (2.3) by the nice real-variable theory lemma mentioned in Exercise 1.5.4. Since 6 is P-almost surely right-continuous, (2.3) will be proved once we show that for any n>\ and 0 = r© < • • • < t„ = T: PI max^(rfc)
>x\<-E e(T\ max0(rfc)>/l
\0
/
^
0
To this end, define AQ = {^(to) > A}, and for 1 < /c < n set A = \^(k) > ^ I
and
max 6(ti) < A . 0
Clearly A^ nAj = 0 if i # ; , {maxo^fc^„ 0(tk)> X] = Ijo^fc^ and A^ e ^^^ for 0 < /c < n. Hence p( max 0(t») > A ) = X P{A^) < \ f £[0(t»), A,\ \0
/
0
^ 0
<\iE[e(nA,]=\E
e{T), m a x % ) > A 0
and the proof is complete. D We now want to introduce the important concept of a stopping time. A function T: £ -• [0, 00) u {oo} is called a stopping time (relative to {J^,: ^ > 0}) if, for all t >0,{T
nC^
0}.
In this connection, observe that the last time that 6 leaves C is not a stopping time since one has to know the entire history of the path 0( •, q) in order to determine if it never visits C after time t. The question of when the entrance time TA =
{Mt>0:e(t)eA}
is a stopping time is a difficult but interesting problem. If the trajectories are continuous and A is a closed set, then T^ is easily seen to be a stopping time. For the general case, see Dynkin [1965], Chapter IV, Section 1 for a discussion of the measurability of various entrance times relative to the completed ^r-algebras. Since we will be working exclusively with processes that are right continuous and almost surely continuous for every closed set C, the contact time T = {inf r > 0: [%): O < w < r ] n C # 0 } is a stopping time and agrees almost surely with the entrance time TC
1.2. Martingales
23
Given a stopping time T, we define ^, = {AeJ^:An{T
for all
t > 0}.
It is easy to check that ^^ is a sub o--algebra of J^ and that i^, = J^, if T = r. Intuitively, J^^ should be thought of as the set of events " before time T ". (Lemma 1.3.2 in the next section makes this intuitive picture precise in the case of a path space.) The following lemma collects together some elementary facts about stopping times. 1.2.4 Lemma. Ifx is a stopping time, then x is ^^-measurable. Ifx is a stopping time and 6 is a progressively measurable function, then 6(x) = 0(x('\ •) is ^ ^-measurable on {T < oo}. Finally, given stopping times a and x: (i) (T + T, a VT, and a AX are stopping times, (a) if A e ^ „ , then A r\\G <x\ and A r\{G <x\ are in ^^^t» (n7) if(T<x, then ^^^ ^,. Proof The proof that x is .^^-measurable is left to the reader. Suppose that 6 is progressively measurable and let r > 0 be given. Define/, on ({T < t}, ^t[{^ < t}]) by f(q) = (T(^) A t, q). T h e n / is measurable into ([0, t] x E, ^^o,t] >< ^t)- Since 9 restricted to [0, t] x E is ^^o, t] ^ ^r^^^^surable, it follows that Oof is ^t[{'^ < t}]-measurable on {T < t}. But 6 of is just the restriction of 6 to {x < t}, and so the assertion that ^(T) is J^^-measurable on {T < oo} has been proved. The proof of (i) is easy and is left to the reader. To prove (ii), we first show that {a < x] and {T < a} are in J^^ n J^^. By an obvious complementation argument, it suffices to prove that {a <x}e ^„ n ^^. Given t > 0, let g, denote the rational numbers in [0, t]. Then {(7 <x] n {x
[j {a <s} n {x> s} n {x
and the right-hand side is certainly a member of J^,. This proves that {(7 < T} e J^,. To show that {a <x} e ^^, note that {(7 <x} n {(7
T})
n {(T A T < r} = (^ n {(7 < t\) n ({a < T} n
{
< t}) G ^t
because {o- < i} G J^^, and {A r\ {a <x\) r\ {G /\x
{[G
< T} n {T < t\) e ^^
24
1. Preliminary Material: Extension Theorems, Martingales, and Compactness
since {CF <x} e ^^. Now (ii) is proved. Finally, (Hi) is obviously just a special case of(«). D We now turn to Doob's optional stopping time theorem. Let a and T be stopping times with values in the finite set {to, ..., r^v}, where 0 = to <"• < tf^ — r, and assume that cr < r. Given a martingale (6(t), J^,, P) and A e ^„,we have: E[e(x\ A] = X E[e(t,), ^ n {T = t,}] k=0
= i E[0(T), An{x = U] fc = 0
= I E[e(t,), An{a
= h}] = E[e{c), A],
k=0
since A e ^ „ ^ ^^ and therefore ^ n {T = f^} and A n {o" = r J are in J^,^. This proves that (2.6)
E'[e(x)\^^] = e(a)
a.s.
Next suppose that (9(t), ^^, F) is a submartingale and define ^(0) = 0 and Mh)-A(t,.,)=(E[e(t,)\^,^_J-e(t,.,)) for l
Ait) = A(tk),
and ^, = #;,
if t € [ffe, tfe+i),
for 0 < /c < N. Finally, set M(r) = ri{t A 7) - ^(r A T). Then (M(t), J^,, P) is a martingale, and so, by (2.6), £ [ M ( T ) | J ^ J = M((7)
a.s.
Since ^(T) > yl(a), it follows that (2.7)
£[e(T)|.^J>%)
a.s.
1.2.5 Theorem. Let a and x he hounded stopping times such that s
E[e(x)\:F,]>e{a)
a.s.
and if (9(t), #",, P) is a martingale, then (2.9)
£ [ 0 ( T ) | J ^ , ] = 0((7)
a.s.
25
1.2. Martingales
Finally, if(0(t\ ^ , , P) is a non-negative suhmartingale after time s and T > s, then {6(xvs): T is a stopping time bounded by T} is a uniformly P-integrable family. Proof Assume that 5 = 0. Let (d(t\ J^,, P) be a non-negative suhmartingale. Choose 7 > 0 so that G <x
fj = i—'-
^
and
[m] + 1
T„ = "^-^
.
Then (j„ and T„ are stopping times and cr„ < T„ < T. Since (T„ and T„ take only a finite number of values, we have from (2.7) that: (2.10)
£[^(t„)|^J>e(0
a.s.
(2.11)
E[e(T)\^,„]>e(x„)
a.s.
(2.12)
£[^(r)|i^J>e((T„)
a.s.
From (2.11) and (2.12), respectively, we know that
E[d(x„%e(x„)>x]<E[e(ne(x„)>k] and £[0(ff„), 0(c„) > A] < £[^(T), 0(ff„) > X]. Since 0(r) > 0, these imply: E[e(x„), 0ix„) >X]< E[e{n
sup 0{t) > X]
0
and £ [ % „ ) , % „ ) > A] < £ 6(7),
sup 0(r) > /I 0
But, by (2.4),
p( sup0(t)>A)<|£[e(T)], \o
/
^
and so we have proved that {6((7„): n> 1} and {^(T„): « > 1} are uniformly Pintegrable. Since 9 is P-almost surely right continuous, we now know that 6((7„) -• 6((7) and 9(x„) -• 9(x) in L^(P). Finally, ifAe^„, then A e J^^„, since a
Letting n -• oo, we now get (2.8).
26
1. Preliminary Material: Extension Theorems, Martingales, and Compactness
The proof that {^(T): T a stopping time bounded by 7} is uniformly integrable when (9(t), J^,, P) is a non-negative submartingale is accomphshed in exactly the same way as we just proved that {^(T„): W > 1} is uniformly integrable. The details are left to the reader. Finally, suppose that (9(t), ^t, P) is a martingale. Then (|^(0|» =^r» P) is a non-negative submartingale. Thus if (7„ and T„ are defined as in the preceding, {16(x„) \: n> 1} and {16(a„) |: n > 1} are uniformly P-integrable families and so 0(T„) -^ ^(T) and e(a„) -^ 0((T) in n(P). Since, by (2.6), E[e(T„)\^J
= e(a„)
a.s.
for all « > 1, the rest of the argument is exactly like the one given at the end of the submartingale case. D 1.2.6 Corollary. Let T: [0, oo) x £ -• [5, 00) be a right-continuous function such that T(t, ')isa bounded stopping time for allt >0 and T(% q) is a non-decreasing function for each q e E. If(0(t), J^,, P) is a (non-negative sub-) martingale after time s, then (6{T(t)), <^t(t)» P) is a (non-negative sub-) martingale after time 0. 1.2.7 Corollary. Ifr > s is a stopping time and (6(t), #",, P) is a (non-negative sub-) martingale after time s, then (^(r AT), J ^ , , P) is a (non-negative sub-) martingale after time s. Proof By Corollary 1.2.6 (6(t A T), J^,^^ , P) is a (sub-)martingale. Thus if ^ e J*^,^, then for ^2 > ti: (>) £[^(^2 AT), A n {Z> r j ] = E[e(ti AT), ^ n {T > tj}], since A n {r > t^} e ^ti^x{t < fj}, and so:
On the other hand, 9(t2 AT) = ^(T) = 0(ti AT) on
(>) £[^(^2 AT), A n{x < r j ] = E[e(ti AT),An{x< Combining these, we get our result,
t^}].
n
The next theorem is extremely elementary but amazingly useful. It should be viewed as the " integration by parts " formula for martingale theory. 1.2.8 Theorem. Let (6(t), J^,, P) be a martingale after time s andrj: [s, 00) x £ ->• C a continuous, progressively measurable function with the property that the variation \rj\(t, q) ofrj(', q) on [s, t] is finite for all t > s and q e E. If for all t > s (2.13)
s u p | e ( u ) | ( | „ | ( t ) + \„(s)\)
< 00,
then {0{t)ri{t) — j ^ 0{s)rj{ds), 9',, P) is a martingale after time s.
27
1.2. Martingales
Proof. Assume s = 0. Using Exercise 1.5.5, one can easily see that jo 0(u)r](du) can be defined as a progressively measurable function. Moreover, (2.13) certainly implies that 0{t)rj{t) — Jo 6(u)rj(du) is F-integrable. Now suppose that 0 < f j < ^2 and that A e J^^j. Then ^(^2)^(^2) - ^(^1)^(^1) - fo(u)rj(dul
A \ = E\ \ (e(t2) - 0(uMdul
A
Since
Emt,)-eit,Mt,iA]
= o.
But if A = r2 - ti, then
(\e(t2) - e(u))r,(du) = lim X ieit^) - ^ i + ^ ^ ) )
and, by (2.13) and the Lebesgue dominated convergence theorem, the convergence is in L^(P). Finally,
^[(^(t2)-^(r.+^A))(,(t, + ^ A ) - , ( r , + ^ for all n > 0 and 1
fie(t,)-e{uMdu),A and the proof is complete.
Q
The final topic of the present section is a theorem which will serve us well in what follows. Basically, this result shows that the martingale property is invariant under certain ways of conditioning a measure. Before we state the theorem, we need the following lemma, which is often useful. 1.2.9 Lemma. Let 6: [s, 00) x E-^Cbea progressively measurable, P-almost surely right continuous function such that 6(t) is P-integrable for all t > s. Let D ^ [s, 00) be a countable dense set. If 9 is non-negative and (2,14)
£[e(f2)|^„]>e(t.)
a.s.
for all ti,t2 € D such that t^ < t2, then (0{t), J^,, P) is a non-negative submartingale after time s. If (2.15)
E[e(t2)\^t,]
= e(t,)
a.s.
for all ti, tje D such that t^ < tj, then {6(t), J^,, P) is a martingale after time s.
28
1- Preliminary Material: Extension Theorems, Martingales, and Compactness
Proof. Assume that s = 0. Clearly the proof boils down to showing that in either case the family {| e(t) \: te [0, T] n D} is uniformly P-integrable for all T e D. Since (2.15) implies (2.14) with |^(*)| replacing ^(•), we need only show that non-negativity plus (2.14) implies that {9(t): t e [0, T] n D} is uniformly Pintegrable. To this end, we mimic the proof of (2.4), and thereby conclude that pi
sup \te[0,T]
e(t)>A<\E[e(T)l nD
J
/l>0.
^
Combining this with
E[e(t), e(t) >x\< E[e(T% e(t) > x\ <E
e(Ti
sup e(t) > X
t e [0, T] n D,
f 6 [ 0 , 7 ] r^D
we conclude that {9(t): t e [0, T] n D} is uniformly P-integrable.
D
1.2.10 Theorem. Assume that for all t >0 the o-algehra ^^ is countably generated. Let X >s he a stopping time and assume that there exists a conditional probability distribution [Q^] ofP given ^ , . Let 6: [s, co) x E -^ R^ be a progressively measurable, P-almost surely right-continuous function such that 9(t) is P-integrable for all t > s. Then_(6(t), ^ , , P) is a non-negative submartingale after time s if and only if (9(t A T), ^ , , P) is a non-negative submartingale after time s and there exists a P-null set N e ^^ such that for all q' ^ AT, (9(t)X[s, t]('^), ^t^ Gg) ^^ ^ non-negative submartingale after time s. Next suppose that 9: [s, oo) x E^Cisa progressively measurable, P-almost surely right-continuous function such that 9{t) is P-integrable for all t > s. Then {9(t), ^^, P) is a martingale after time s if and only if(9(t A T), ^ , , P) is and there is a P-null set N such that (9(t) - 9{t AT), J^,, Q^) is a martingale after time s for all q' i N.
Proof Assume that s = 0. We suppose that {9(t), ^ , , P) is a martingale. Then by Corollary 1.2.7 so is (9{t AT), ^t^P)- Let 0 < tj < f2» ^ e ^^ and A e ^^^ be given. Then E^[E'^[9{t2\ A\ B n{x<
r j ] = E^[9(t2\ A n B n {t < t^]] = £^0(ri)M n 5 n {T
E\EQ[9(t,\AlBn{x
Here we have used the fact that A r\ B n {x
1.2. Martingales
29
Taking a single null set for a countable subalgebra of sets A generating ^t^ we obtain a null set ^^1,^2 such that, for q' ^ N^^^^^ ^^^ '^W) ^ h E<^^[e(t2)\^,,] = e(t,)
a.s. Q,,.
We now take a countable dense set D in [0, 00). We can then find a single null set N such that for q' ^ N E<^^[e(t2)\^,,] = e(t,)
a.s. Q^,
provided tj, ^2 e D and ^2 ^ ^1 ^ '^(^')From Lemma 1.2.9 we can now conclude that for q' ^ AT, (6{t), J^,, Q^) is a martingale for r > T(^'). This can of course be restated as {9(t) — 6(t AT(q% J^,, Qq>) is a martingale for r > 0. Since (2.16)
P k : e j ^ : T ( ^ ) = T(^')] = l ] = l
we are done. Now suppose {6(t), ^^, P) is a non-negative sub-martingale. Then by Corollary 1.2.7 so is {^{t/\x\ #',,P). By replacing equalities by the obvious inequalities we can conclude that there is a null set N in ^^ such that (0(f), J^,, Q^) is a non-negative submartingale for t > x(q') provided q' ^ N. We note that this is equivalent to X[o, t]('^(Q'))H^) being a non-negative submartingale for t > 0. Again by (2.16) we are done. We now turn to the converse proposition. If 0
E''[E^^[e{{t,Ax{q'))vt,),A]] E''[E
= E''[e(til An{x<
ti}] + £'"[0(r Afj), /4 n {t > tj}]
= E''[e{ti), An{T<
ti}] + Ele{x A(,), A n{x> t,}]
= E-m,); A]. The submartingale case is proved in the same manner by replacing the equalities by inequalities at the relevant steps. D 1.2.11 Remark. It is hardly necessary to mention, but for the sake of completeness we point out that everything we have said about almost surely right-continuous martingales and submartingales is trivially true for discrete parameter martingales and submartingales. That is, if (£, J^, P) is a probability space, {^„: n >0} a non-decreasing family of sub
a.s.
30
1. Preliminary Material: Extension Theorems, Martingales, and Compactness
for all n>0. The obvious analogues of (1.2.4) through (1.2.10) now hold in this context. Indeed, if one wishes, it is obviously possible to think of this set-up as a special case of the continuous parameter situation in which everything has been made constant on intervals of the form [n, n -\- 1).
1.3. The Space C([0, oo); R"^) In this section we want to see what the theorems in Section 1.1 say when the Polish space is C([0, oo); R^). The notation used in this section will be used throughout the rest of this book. Let Q = Qj = C([0, 00); R'^) be the space of continuous trajectories from [0, oo) into jR'^. Given r > 0 and co e Q let x(u co) denote the position of co in R'^ at time t. Define Dico co')= y ^ supo<,<,|x(f, a j ) - x ( r , co^)| [co^co) 2^^ 2"l-hsupo^,^„|x(r,co)-x(t,co')| on Q X Q. Then it is easy to check that D is a metric on Q and (Q, D) is a Polish space. The convergence induced by D is uniform convergence on bounded rintervals. We will use M to denote the Borel a-field of subsets of (Q, D). Clearly the map x{t) given by co -• x(r, o) is D continuous and therefore measurable, for each t > 0. Thus
(y[x(t):
t>Qi\^Ji.
On the other hand, if co^ G Q, f > 0 and e > 0 are given then {co: sup |x(s, co) — x(s, co^) | < e} 0
<s
= [J{co : |x(s,co) -x(s,co^)| <s(l
j
for all rational s in [0,t]}. The set in question is therefore clearly in o-[x(f): t > 0]. Since sets of the form {w: sup |x(5, co) - x(5, co^) I < e} 0 <s
generate the topology of Q, we conclude that (3.1)
Jf = (7[x(t): r > 0 ] .
Next we define ^^ for r > 0 by (3.2)
^ , = (7[x(s): 0 < 5 < r].
1.3. The Space C([0, oo); R")
31
Clearly Ji^ ^ Jl^iox s
M, =
G[\]JI\
for t > 0.
By (3.1) we also have that (3.4)
^ = (7|U^^r).
The following theorem is a handy form in which to have Theorems 1.1.3 and 1.1.4. 1.3.1 Theorem. A family ^ of probability measures P on (Q, J^) is precompact if and only if lim inf F[|x(0)| < .4] = 1
(3.5)
and for each p > 0 and T < co (3.6)
sup
lim infF
\x(t) — x(s)\ < p = 1
0<s
Proof By the Ascoli-Arzela theorem a closed set K cz Q is compact if and only if (3.7)
s u p I x ( 0 , co) I <
00
(oeK
and for each T < co (3.8)
lim sup ^iO (ae K 0
sup
\x(s, oj) — x(t, oj)\ =0.
<s
We can therefore obtain immediately the necessity of (3.5) and (3.6) from Theorem 1.1.3. We now prove the sufficiency. Let us pick p — 1/n and T = n and find S„ — S„(e) such that for given £ > 0 I
inf P
sup
\x(t)-- x(s)\ < l/n\ > 1
p >«+l
0<s
We pick A from (3.5) such that infP[|x(0)| < ^ ] > 1 P e^
If we define
K. = r\
(jo\ sup 0<s
Ix(t, co) — x(s, co) I < - n [co: |x(0, co)| < A] ^
32
1. Preliminary Material: Extension Theorems, Martingales, and Compactness
then clearly P(K^) > 1 — £ for all P e ^. Moreover by the Ascoli-Arzela theorem Kg is compact in Q. So the sufficiency part of our theorem now follows from Theorem 1.1.4. D For a sequence {P„} of probability measures on (Q, ^ ) it is sometimes more convenient to have the following alternate form of the condition for precompactness. 1.3.2 Theorem. Let P„ on (Q, J^) satisfy the following
liminfP„[|x(0)| < .4] = 1 and for any p > 0 and T < oo lim lim sup P„
sup |x(r)-x(5)| >p
0.
0<s
Then {P„} is precompact. Proof Define for each fixed T < co and p > 0
US)-Pr
sup
\x(t)-x(s)\
>p
0<s
We then have for each n, il/„(S)lO as 3iO and lim lim sup il/„(S) = 0. Let e > 0 be given. There exists S(e) > 0 and no(s) < 00 such that for n > no(e) IA„(<5(E)) < 8.
Since each il/„(S)^0 as ^-•O, we can find ^i(e) such that for 3 <^i(£) and n < no(e) W^) < £• If we now take S2(E) = S(e)ASi(s) we conclude that il/„(d) <e for all n and 3 <S2(s). Therefore lim sup il/„(S) = 0 and Theorem 1.3.1 is now applicable,
n
1.3. The Space C([0, oo); /?")
33
We next want to see \vhat Theorem 1.1.8 looks like when the conditioning (T-algebra is ^ , , t being any stopping time. In order to do this we need the next lemma. 1.3.3 Lemma. Let x he any stopping time. Then (3.10)
^,
=
s>0].
In particular, M^ is countably generated. Proof. We first note that X(SAT) is ^^ measurable for each s >0. In fact, by Lemma 1.2.4, X(5AT) is ..^s^^-measurable, and M^^^^ Ji^. To prove M^ ^
/(ca) = F(x(ri, co), x(r2, w), ..., x(f„, w), ...).
Moreover,/is M^ measurable if and only if the {r„} can be chosen from [0, t\. Thus /is Jit measurable if and only if/is M measurable and/(co) =/(ca,) for all co. Now suppose that/is J(^ measurable. For r > 0, define/(co) = X{j}(T((y))*/(a;). Then/ is Jt^ measurable and therefore
In particular (3.12)
/(tt>)=/(e.)(co)=/(^)(co,(J
for all ca e Q.
Applying this to/(-) = X{o(^(*))» we get X{r}(T(^)) =
Zt(a,)(TK(a»)k{r}(T^K(ca)))-
If we now take t = x{oS) in this, we see that T((O) = we conclude that for any M^ measurable/ (3.13)
/(co)=/(co,(J
T(CO^(^)).
Using this fact in (3.12)
for all co e Q.
Finally given a Ji^ measurable function/ choose F so that (3.11) holds. Then by (3.12) /(co) = F(x(ri A T(ca), co), ..., x(r„ A T(CO), CO), ...) for o) e Q, which explicitly displays/as a o'[x(s AT): S > 0] measurable function.
D
34
1. Preliminary Material: Extension Theorems, Martingales, and Compactness
Given a stopping time T and a measure P on (Q, ^) we can now use Theorem 1.1.8 to find a conditional probability distribution {Q^>} of P given J^^ such that off of some P-null set N G M^ , we have (3.14)
eco'M) = Z > ' )
for all
AGM,.
Suppose we define Q^- so that outside iV, Q^ = Q^ and for co' e N and A e ^
where co'^^oi is defined as in Lemma 1.3.2. Then it is clear that Q^ is again a conditional probability distribution of P given Jt^.. In addition, for the new version (3.14) holds for all co'. 1.3.4 Theorem. IfP is a probability measure on (Q, ^) and xisa stopping time, then there exists a conditional probability distribution {Qo,} of P given M^ such that (3.14) holds for alio)'. If {Q^] is as in the preceding theorem we will call it a regular conditional probability distribution ofP given Ji^ and we will abbreviate this phrase by r.c.p.d. of P\M^. Notice that an equivalent way to express (3.14) is (3.15)
QioK^{^) — ^(<^') and
x(s, co) = x(s, a>') for
0 < s < T(CO)) = 1.
The version of Theorem 1.1.9 which is most suitable for the study of measures on (Q, M) is the following: 1.3.5 Theorem. Let {T„: n > 0} !?e a nondecreasing sequence of stopping times and for each n suppose P„ is a probability measure on (Q, ^r„)- Assume that P„ +1 equals P„ on J^r„for each n>0. If\im„^^ P„(T„ < t) = Ofor all t > 0, then there is a unique probability measure P on (Q, ^) such that P equals P„ on M^^for all n > 0. Proof. If P exists, it is obvious that for Ae M^ (3.16)
P(A)= l i m P „ [ / l n { t „ > t } ] . n->oo
Thus uniqueness of P on Jti for all r > 0 is proved; and therefore P, if it exists, is unique on M. To prove existence, first assume that T„ = n for each n > 0. Clearly (Q, M^ is isomorphic to (C([0, n\\ K% ^(C([0, n\\ K^))) and therefore we can find a r.c.p.d. {Q^.} of P„ given Ji^-1 for each n>\. Moreover if An(p) = n {B e M^, co € B}
1.3. The Space C([0, oo); R")
35
then A„(a)) = {(JL>': x(s, co') = x(s, co) for 0 < s < n].
Therefore if {a)„}o ^ Q has the property
for all iV > 0 then the co determined by x(t, co) = x(r, co„),
0
and w > 0
is in P)* ^„(co„). We can now apply Theorem 1.1.9 to conclude the existence of a P on (Q, J^) such that P equals PQ on J/Q ^^^ P(A)=\QUA)P{d(o') for all n > 1 and A e J^„. By induction on n > 1, it is easy to see that P equals P„ on J^„ for all « > 0, and so P is the desired measure. In general, we first define P„ on J^„ by PM) = lim Pk(A n
{T,
> «}),
Ae^„,
fc-»oo
Note that the limit exists, since P,,(A n {Tfc > n}) = Pfc+ ,(A n {T^ > n}) < P„^ ,(A n {T^^ i > n}). Also it is clear that P„ is afinitelyadditive probability distribution on J^„. To see that P„ is countably additive, suppose that {A^}f ^ J^n and v4„i0. Then for all k
< Pfc(Tfc < n) + Pfc(X, n {Tfc > n}). Letting m and then /c -> oo, we see that P„(^,„) -• 0 as m -• oo. A similar argument shows that P„+i equals P„ on ^ „ for n > 0. Thus, by the preceding paragraph, there is a P on (Q, ^) such that P equals P„ on ./#„ for all w > 0. Finally we must check that P equals P^ on J/^^. Given ^ e Jf^^ n ^ „ , we have that P(A) = P„(^)=limP,(An{T,>n}). But for / > /c I P,(^) - P,(A n {T, > n}) I = P,(^ n {t,
36
1. Preliminary Material: Extension Theorems, Martingales, and Compactness
as 1-* CO. Thus P(A) = Pk(A) for A e Ji^^ n Mn- But, by Lemma 1.2.2, M^^ is generated by the maps x(r AT^), for t > 0. Therefore
Hence P equals P^ on ^^^, and we are done.
D
1.4. Martingales and Compactness In the preceding section we developed necessary and sufficient conditions for the compactness of measures on (Q, ^). Like most general results, these conditions are not particularly useful when applied to special situations. It is the purpose of this section to develop a useful condition for compactness. The condition that we have in mind is ideally suited to the study of Markov processes and, more generally, processes for which there is a plentiful supply of associated martingales. Given p > 0 and co e Q, define TQ((O) = 0 and for n > 1: T„(co) = mf{t > T„_ i(co): |x(r, w) - X(T„_ i(a)), co) \ > p/4}. Here it is understood that T„((O) = oo if either T„_ i(co) = oo or there fails to exist a t > T„_i(a)) such that |x(r, co) — X(T„_I(CO), a))\ > p/4. Since co is a continuous path, it must always be true that either T„_I(CU)= oo or T„_I(CO) < T„(a>) and x„(co) -> 00 as n -• 00. Thus for 7 > 0 (this T is arbitrary but fixed throughout), we can define N = N(co) = min{w:
T„+I(CO)
> T}
and (4.1)
SM
= min{T„(co) -
T„_ ,((O):
l
iV(co)}.
We need the following lemma. L4.1 Lemma. Let t^ and ^2 be any pair of points in [0, T] such that 1^2 — ti | < S^(p). Then \x(t2, (o) — x(ti, (o)\ < p and so sup{ |x(r2, co) - x(ri, co) 1: 0 < ti < ^2 < T and
| (2 - tj | < 5Jp)} < p.
Proof. Consider the partition of [0, T] into the subintervals [TO(CO), Ti(co)), ..., [T^(„)_I(CO), T^(e«)(co)), aud [XN(CO)(O^\ T]. All of these subintervals, except possibly the last one, must have length greater than S^(p). Thus, either both ti and t2 lie in the same subinterval, or they He in adjacent subintervals. Since over any subinterval the distance of the path from its position at the left hand end never exceeds p/4, the oscillation of the path over any subinterval must be less than or
1.4. Martingales and Compactness
37
equal to p/2. Hence the oscillation over the union of two successive subintervals cannot exceed p. In particular, \x(t2, co) - x(fi, a))\ < p. D The preceding lemma shows that the problem of estimating P(sup{|x(f2) x{t^) \:0
P({co: SM
< S}).
The method that we are going to use to estimate the latter quantity depends on the following two hypotheses about P: 1.4.2 Hypothesis. For all non-negative f e C^lR'^) there is a constant Af >0 such that (f(x(t)) + Aft, Ml, P) is a non-negative submartingale. 1.4.3 Hypothesis. Given a non-negative f e Co{R% the choice of Aj- in (1.4.2) can be made so that it works for all translates off Under these hypotheses, we are going to develop an estimate for the quantity in (4.2) which depends only on the constants Af. Let e > 0 be given and choose /, € C^iR"^) so that /,(0) = 1, f(x) = 0 for |x| > e, and 0 0 ^(Tn+1 - T„ < ^ I Jt,^) < SApi4.
(a.s., P) on {T„ < oo}.
Proof Let e — p/4 in the preceding discussion and let {Qc^} be a r.c.p.d. of P given M^^. Then we can choose a P-null set F G M^^ SO that
((/r(x(f)) + '4.f)Z[o,,i(T„(a)')), A , e„.) is a non-negative submartingale for all co' ^ f, where/f (x) =/j(x - X(T„(CO'), CO')) if T„(co') < 00 and/f (•) = 1 otherwise. In particular, by Theorem 1.2.5, £«"[/r(x(T„,, A (T„(CO') + ,5)) + A,S\ > 1
38
1. Preliminary Material: Extension Theorems, Martingales, and Compactness
for cd ^ F. In other words, £^-[1
But 0 < 1 -ff
-/f
(X(T„^I A ( T > ' ) + (5))] < A,d, CD' ^ F.
< 1, and T„+ i < T„(a>') + 3 implies that + S))=l
l-ff(x(x„^,A(x„(co') if T„(a)') < 00. Thus
QA^n^i<^M) for co' ^ F such that F(T„..I
-
T„{CO') T„
+ s)
< oo. Since
< ^|^,J =
Q.(T„^I
< T„(-) + ^)
(a.s., F),
this completes the proof. D 1.4.5 Lemma. L^t (£, ^ , P) 6e a probability space and {^„: n > 0} a nowdecreasing sequence of sub a-algebras of ^. Let {(^„: n> 1} be a non-decreasing sequence of random variables on (£, #") ta/cmgf values in [0, oo) u {oo}, and assume that ^„ is ^„-measurable. Define ^o = ^ ^nd suppose that for some A < 1 and all n>0:
^[exp[-(^„^i-U]|«^J<^a.s. Iffor some T > 0 one defines N(q) =
mf{n>0:i„^,(q)>n
then N < oo a.s. and in fact: P(N >k)< e''X\
k>0.
Proof First note that: E[e-^^^^\^„] = e-^"E[cxp[-(Ui
-
U]\^n]
Q
1.4. Martingales and Compactness
39
We are at last in a position to prove the compactness criterion to which we referred in the introduction to this section. 1.4.6 Theorem. Let ^ he a family of probability measures on (Q, J^) such that Hm supP(|x(0)| >/) = 0. Assume that each P e ^fulfills hypotheses 1.4.2 and 1.4.3 and that the choice of the constants Af in 1.4.2 and 1.4.3 can be made independent of P e ^. Then ^ is precompact. Proof. In view of Theorem 1.3.1, we need only check that lim inf pj b^Q P e ^
sup
| x{t) - x(s) \ < p j = 1
\o<s
/ I
for every T > 0 and p > 0. Because of Lemma 1.4.1, this will be done once we show that (4.3)
limsupP({co:^»<^}) = 0.
Note that, from the definition of ^^(p) in (4.1), P(^.(p) < ^) < P( min T, - Ti_ 1 < (51 + P(N > k) \i
J
<^P(T,-x,.,<S)-\-P(N>k) < kdA^,^ + P{N > /c), where we have used Lemma 1.4.4 to get the last line. Thus, the proof will be complete once we establish that (4.4)
lim sup P(N >fe)= 0.
But, by Lemma 1.4.4, we know that for any to>0 and P e ^:
< P(T,^ 1 - T, < to I ^,,) + e-''P(Xi^ 1 - T, > fo I ^u) < e-^o + (1 _ e-'o)P(x,^, - T, < rol-^J <e-''-^(l-e-'')toA^,^a.s.
40
1. Preliminary Material: Extension Theorems, Martingales, and Compactness
Choosing IQ in a. suitable manner, we can make
Thus, by Lemma 1.4.5, sup P(N >k)< e'^X\ and this certainly guarantees (4.4).
D
Although Theorem 1.4.6 is well-adapted to the study of continuous time Markov processes having continuous paths, it is not suitable, as it now stands, for the approximation of such processes by discrete time parameter processes. We will now make the necessary modifications to get a theorem which covers this situation. Given an /i > 0, let Q^ stand for the subset co eQ such that x(%co) is linear over each interval of the form [/7i, (j + l)/i], ; = 0, 1 , — Given coeO;,, define T?(a)) = 0 and for ^i > 1, TJ(CO) =
inf{r > T*_ i(co): t = jh for some; > 0 and |x(r, co) - x(x*- i(a)), co| > p/4}.
We again adopt the convention that T*{CO) = OO if TJ_ i((o)= oo or if |x(r, co) — a>)\ < p/4 for all t>x*-i(a)). Once more, either TJ_I(CO)= OO or r*((o) > TJ_ i(co) for all « > 1 and
X(T„-I((O),
lim T*{(o) = 00. n-»oo
Define for 7 > 0 (arbitrary but fixed): N*(co) = inf{n > 0: T*+ ^((O) > 7}, SZ(p) = min{T„% ,(co) -
T:(CO):
0 < n < iV*(co)}
and et(co) = max{|x((; 4- l)/j, co) - x(jK co)\:0<jh<
T}.
In place of Lemma 1.4.1, we now have the following lemma. L4.7 Lemma. Ift^, t2 e [0, T] and 0 < ^2 - tj < S%(p), then I x(t2, co) - x(ri, co) I < p + 2ei(co). In particular, sup{|x(r2,co)-x(ti, co)|: ti, t j e p , T] and 0 < t^ - t2 < SZ(p)}
1.4. Martingales and Compactness
41
Proof. Given t^ and ^2, let rf and rf be, respectively, the smallest and largest multiple of h such that tf >ti and f^
^l^t'J
< SAp/4
a.s.
We are now ready to prove the analogue of Theorem 1.4.6. 1.4.11 Theorem. Let {h„: n>0} be a non-increasing sequence of positive numbers such that h„-^0 as n^ oo. Let {P„: n > 0} be a sequence of probability measures on (Q, Jt) such that P„ is concentrated on Qf,^. Assume that each P„ satisfies hypotheses L4.8 and L4.9 (with h = h„) and that the choice of the constants Af can be made independent of n. If for each T > 0 and e> 0
(4.5)
lim n-»QO
X ^»(l*(0' + i)K) - AiK)I > e) = 0, 0<jh„
and (4.6)
limsupP„(|x(0)| > / ) = 0, I >"oo
n>0
then {P„: n > 0} is precompact. Proof The proof here goes by analogy with the proof of Theorem 1.4.6. We must show that (4.7)
limlirnTPJ d^O n-*ao
for all r > 0 and p > 0.
sup
I 0 <s
|x(f) - x(5)| > 2p | = 0 J I
42
1. Preliminary Material: Extension Theorems, Martingales, and Compactness
By Lemma 1.4.7, pj
sup
I \
Ix(0 - x{s)\>2p\<
P„id*(p)
0<s
The term P„(ST(p) < S) is estimated in exactly the same way as P(S.(p) < ^) was in Theorem 1.4.6, only one now uses Lemma 1.4.10 instead of Lemma 1.4.4. In this way, one proves that lim sup P„{S*(p) <3) = 0. 5^-0
n>0
On the other hand, by hypothesis,
P„{ei>P/2)<
I
Pn(\x{j+l)K)-x{jh„)\>p/2)
0<jh
^0 as n - • 00
D
L5. Exercises 1.5.1. Let £ be a non-empty set and ^ a collection of subsets of E such that if A and B are in ^ then A n Bisin^. Show that the smallest class j ^ of subsets of £ such that (0 ^ c j ^
(a) Ee^ (Hi) A, B e ^, A cz B implies B\A e ^ \iv) A, Be^, AnB=^0 implies A KJ B e ^ (v) {A„}f e^.An^ A„^ 1 for « > 1 implies {Jf
A„e^
coincides with a(^), the smallest or-algebra containing ^. The proof of this fact is very much like the monotone class theorem (for details see Dynkin [I960]). As an application, show that if j f is a class of bounded functions on E into R such that (i) Xc e -^ for all C e ^ (a) 1 e Jt {in) Jf is a vector space (iv) If {/„}f ^ Jf is a nondecreasing sequence of non-negative functions with sup„,, /„(^) < 00, then/= lim„^^ /„ is again in Jf then J^ contains all the bounded measurable functions on (£, o'(^)). 1.5.2. Let (X, p) be a metric space and denote by Up(X) the class of all bounded, uniformly continuous functions on {X, p) into R. Show that the smallest class of functions from X into R which contains Up(X) and is closed under bounded pointwise convergence is B(X), the set of bounded Borel measurable functions on X.
1.5. Exercises
43
1.5.3. Let (£, ^) be a measurable space. Let ^ e £ be a point. The atom A{q) containing q is defined by A(q) = n [B: B e ^, q e B]. Give an example to show that A(q) need not be in ^ . However if ^ is countably generated and j^/ is a countable subalgebra generating J^ show that A(q) = n [B: B e ^, q e B], and is therefore always an element of ^. 1.5.4. Let (E, J^, P) be a probability space and l e t / ( - ) and g{') be two nonnegative random variables on it. Assume that g is integrable and that P[f(')>X]<jE[g('):f{')>Xl
A>0.
Show that if r > 1, then g e E(P) implies/e E(P) and that
The proof of this inequality turns on a clever application of the formula
£[/'] = j imYPidq) = r IJA-•?[/(•) > X\ dX which is valid for any non-negative/and any r > 1. See Theorem 3.4 in Chapter 7 of Doob [1952] for details. 1.5.5. Let (£i, J^i), (£2, ^2) and (£3, ^^3) be three measurable spaces. Let F: £1 X £3 -• R be a measurable map. Let ^l{q2, dq^) be a signed measure on (£3, ^2) for each ^2 ^ ^2 and a measurable function of ^2 for each set in ^ 3 . Show that the set AT c: £1 x £2 defined by N= [(quqi)' \ 1/^(^1,^3)1 1/^1(^2, ^^3) < 00 is a measurable subset of (£1 x £2, J^i x ^2) and the function J £(^1, ^3)/i(^2» ^^3) is a measurable function of (^1, ^2) on the set N. Deduce from this that if 6(t, q) is a progressively measurable function and rj(ty q) is a progressively measurable continuous function which is of bounded variation in t over any finite interval [0, T], then Z(r, q) is again a progressively measurable function where Z(r, ^) = f'^(5, q)tj(ds, q) if •'0
= 0 otherwise.
f' I ^(5, ^) I I ;y I (ds, q) < co •'0
44
1. Preliminary Material: Extension Theorems, Martingales, and Compactness
1.5.6. Let £ be a non-empty set and rj a map of [0, oo) x £ into a Polish space (X, d). Define on E the a-field J^ = (T[t]{s): s > 0]. Show that if/is an ^ measurable map of E into a Polish space M, then there exists a ^^z+ -measurable map F of X"^^ into (M,^) and a sequence {t„}f £ [0,oo) such that /(^) = F(^(fi,^), ...,^/(r„,^)•••)»
qeE.
Next, assume that ^/(^ (?) is right continuous for each qeE and define ^t = o-[^(5): 0 < 5 < r]. Given a measurable function 6: [0, oo) x E^M such that for each t, 6(u •) is J^,-measurable, show that there exists a ^[o,x) X ^j^2+-measurable map F of [0,oo) x X^^ and a sequence {t„}5° ^ [0,oo) such that for all t > s and q £ E 0(s At,q) = F(s A r, rj(t^ A r, ^),..., ^(r„ A r, ^),...). In particular conclude that for each fixed r, 6(s A t, ^) is measurable in (5, q) with respect to ^[o, n x -^r» ^i^d hence ^( •, •) is progressively measurable. 1.5.7. Let (£, ^, P) be a probability space and I c J^^ be a sub tr-field. Let (7, ^ ) be a measurable space and F: E x Y-^ R a measurable function (relative to J^ X ^ ) such that sup^^^y £^[|F(*, y)\] < 00. Show that a version G(q, y) of £^[F(', y)|I] can be chosen so that G(*, •) is L x ^ measurable. Suppose now that we have a map /: £ -• 7 which is I-measurable. Assuming that £''[|^(%/(;))|]< 00, show that £ V ( % / ( - ) ) | 2 ] = G(.,/(.))
a.e.
1.5.8. Suppose (9(t), i^,, P) is a martingale on (£, i^, P). Let J^,+o = C]s>t ^s where .^^ is the completion of #"5 in (£, .^, P). (That is J e .^^ if and only if there is an .4 in ^, with A AA a B where Be^ and P(B) = 0.) Show that {6(t), ^t+o, P) is a. martingale. 1.5.9. Use Theorem 1.2.8 to show that if (9(t), #",, P) is a continuous real valued martingale which is almost surely of bounded variation, then for almost all q, 6(t) is a constant in t. Note that this conclusion is definitely false if one drops the assumption of continuity. 1.5.10. Suppose (9{t), ^,, P) is a martingale on (£, J^, P) such that sup,^o £(^(0)^ ^ ^' Show that E{9{t))^ is an increasing function of ? with a finite limit as r -• 00. Use this to show that d{n) tends in mean square to a limit ^(00) as n-* CO. Next, use Doob's inequality to prove
sup 1^(0-^(5)1 >e
<-,Em^)-e'(s)i
In particular 6{t) -• 0(oo) a.e. as t ^ 00. Finally, show that if T is an extended stopping time relative to {#"{}, then £^ [O(co)\^t] = 9{x) a.e. P. This is an especially easy case of Doob's Martingale Convergence Theorem.
1.5. Exercises
45
1.5.11. Let (£, ^) be a measurable space and J^,, f > 0 a non-decreasing family of (T-fields such that ^ = o-(lJ, J^,). Given /I £ [0, oo) x £, we say that A is progressively measurable if XA{'^ •) is a progressively measurable map from [0, oo) x £ into R. Show that the class of progressively measurable sets constitute a (T-field and that a function/: [0, oo) x £ -• (X, ^) is progressively measurable if and only if it is a measurable map relative to the (T-field of progressive measurable sets. 1.5.12. With the same notation as above, let i : £ -> [0,oo] be an extended non-negative real valued function such that for each r > 0, {^: T(^)
< a} = ([0, t] X {q: x(q) < a}) u [0, a] x £.)
1.5.13. Let A(t\ r > 0 be a non-increasing family of subsets of £ such that A(t) G ^t for each t > 0. Consider the set A=
[j(t,A(t))
=
{(t,q):qEA(t)}.
f>0
Define B(t) = A(t - 0) = P|,<, A(s) for r > 0 and B(0) = A(0). Show that if A is progressively measurable then B = [j(t,B(t))
=
{(t^q):qeB(t)}
t>0
and B\A are progressively measurable too. (Hint: Consider the function/(r, q) = ;f^(r, q). From the fact that/(r, q) is progressively measurable show that/defined by /(t,q)=/(t-0,9)=lini/(i(l-^|,<j| is again progressively measurable. Identify/as XB)
Chapter 2
Markov Processes, Regularity of Their Sample Paths, and the Wiener Measure
2.1. Regularity of Paths Suppose that for each n > 1 and 0 < tj < • • • < r„ we are given a probability distribution F„,...,,„ on the Borel subsets of (R^f. Assume that the family {Ptu...,tJ is consistent in the sense that if {sj,..., 5„_ j} is obtained from {t^, ..., t„} by deleting the kth element t^, then P^^^ ..,s„-i coincides with the marginal distribution ofPf^^ j^ obtained by removing the kth coordinate. Then it is obvious that the Kolmogorov extension theorem (cf. Theorem 1.1.10) applies and proves the existence of a unique probability measure P on (R'^f^- "^^ such that the distribution of (il/(ti), ..., lA(^n)) under P is Pti,...,t„- (Here, and throughout this chapter, i/^ stands for an element of (R'^f^' "^^ and V(0 is the random variable on (R'^f^^ "^^ giving the position of ij/ at time t.) As easy and elegant as the preceding construction is, it does not accomplish very much. Although it establishes an isomorphism between consistent families of finite dimensional distributions and measures on a function space, the function space is the wrong one because it is too large and the class of measurable subsets is too small. To be precise, no subset of (R^f^' "^^ whose description involves an uncountable number of fs (e.g. {ij/: supo
2.1. Regularity of Paths
47
The question now is: how does one go about determining when the P associated with {Pti,...,t) gives Q outer measure one? The answer is contained in the next lemma. 2.1.2 Lemma. Let P be a probability measure on ((R'^f^' *\ ^(/jd)[. oo). Then Q has P-outer measure one if and only iffor every bounded, countable set 5 ^ [0, oo): (1.1)
P({^'- ^ \s is uniformly continuous}) = 1.
{Here ij/ \s stands for the restriction ofij/ to 5.) Proof First suppose that Q has P-outer measure one. Given S, note that As = {il/: il/\s is uniformly continuous} is a Borel set and that Q^ As. Hence P(As) = 1. Next assume that (1.1) holds for all 5 and let /I be a Borel set containing Q. By exercise 1.5.6, there exists a measurable function F on ((R'^Y^) and a countable set T = {t„: n > 1} ^ [0, oo) such that
Z^(^) = F(^(rO,...,^(0,...),
i^e(RT"^'-
Thus if 5^ = r n [0, iV], f)N=i {^'- ^ |s^ is uniformly continuous} £ A. But, by (1.3), P({^'. ij/ \s^ is uniformly continuous}) = 1 for all N; and so, by the monotone convergence theorem, P(A) = 1. This shows that Q has P-outer measure one. D We are now going to develop a criterion on {P,^ ^J for testing whether the associated P satisfies (1.1). The basic method which we will use goes back to Kolmogorov, although the elegant approach that we employ here is due to Garsia, Rademich, and Rumsey [1970]. It must be admitted that as a tool for studying Markov processes, Kolmogorov's criterion is rather crude by comparison to the machinery which we developed in Section 1.4. On the other hand, it has the important feature that it depends only on the two dimensional marginals P^ ^, and not on any higher order structural properties of the process. This fact makes it more ubiquitous than more refined results. 2.1.3 Theorem. Let p and ^ be continuous, strictly increasing functions on [0, oo) such that p(0) = ^(0) = 0 ami lim, , ^ ^(r) = oo. Given T>0 and (pe C([0, 7], R'l if
then for 0 <s
(1.3)
|#)-^(5)|<8jJ V ' ( ^ j p W .
48
2. Markov Processes, Regularity of their Sample Paths, and the Wiener Measure
Proof. Define d.^ = T and
'^yn^h Since /Q I{t)dt < B, there is a to € (0,^_i) such that I (to) < B/T. We are now going to choose a non-increasing sequence {t„:« > 1} ^ [0,to] as follows. Given f„_ 1, define d„^ i by p(d„_ i) = ip(r„_ J. Choose t„ e (0, ^„_ i) so that (1.4)
I(t„)<2B/d„.,
and
This can be done because the set of f G (0, ^„_ j) on which either one of these inequalities can fail must have measure less than d„-1 /2, and so there is a point in (0, ii„-1) at which they both hold. Clearly
Thus d„^ 1 < ^„+1
p(t„) = 2p(d„) = 4(p(^„) -
Wn))
<4(p(d„)-p(d„^,)). Combining (1.4), (1.5), and (1.6) with the fact that d„ < ^„_i, we see that mn) - Cl)(tn+l)\ < '¥-H2I{tn)/dn)p(tn - tn-fl)
p(d„^,))
<4^-'(4B/d'„)(p(d„)-p(d„^,)) dn
<4
'V-'{4Blu^)p{du). ^A
Summing over n > 0, we now get (1.7)
|^(to)->(0)| < 4 f
^-\ABIu^Wu).
Replacing >(r) by (j)(T — t) in the preceding argument, we conclude that (1.8)
\(t>(T)-(l>(to)\ <4
\\-'{4B/u')p(du).
2.1. Regularity of Paths
49
Adding these, we arrive at \ct>(T) - 0(0)1 < 8
(1.9)
j\-'(4B/u')p(du).
Note that (1.9) now holds for any T, p, and 0 such that (1.2) is satisfied. In particular, if 0 < s < r < 7 are given and we define 4>(u)==(pis-\-^-^-u\
M6[0, T],
and PW = P ( ^ - « 1
we [0,7],
then
Thus, replacing 0 by 0, p by p, and B by B in (1.9), we have
|^(r)-^(0)|
<8\\-'(4B/u')p{du), •'o
which becomes (1.3) after the obvious change of variables. D
2.1.4 Corollary. Let (£, ^, P) be a probability space and 0: [0, co) x E-^ R^ a ^[0,00) ^ ^-measurable function such that d(*, q) is continuous for all q e E. If for each T >0 there exist number a = aj > 0, r = r^ > 0, and C = C7- < oo such that
(1.10)
E[\eit) - 9{sW] < c\t- s\^+\
0 < s,r < r,
then for any 7 = 7^ e (2, 2 + a^) and X > 0: (1.11)
P(
sup
IMziM,J)_(4Ar)^C.A
where p = Pj, = (y^- 2)/r^ andA = Aj = ^l^l
\t - sP""""^ ds dt.
50
2. Markov Processes, Regularity of their Sample Paths, and the Wiener Measure
Proof. From (1.10) we know that (t) - 0(s) I \^
1
It(1fc1!^["|
Thus,
cn^f^f"-)-^'^ T
.T
and by Theorem 2.1.3,
1^(0-^(5)1 < 8 [ (^J du^i^ ^^ (4Xy''\t-sf, y-2
0<s
if
cn^r^)-—• 2.1.5 Corollary. Let Pbea probability measure on ((R'^f^' '^\ ^(Rd^io,«)) and suppose that for each T > 0 there exist numbers a = ar > 0,r = r j > 1 + a j , and C — CT < oo such that (1.12)
E[ Iil/(t) - il/(s) n < CI r - 5 P•'^
0 < 5 < r < r.
Then Q has P-outer measure one. Proof. According to Lemma 2.1.2, we must show that if T > 0 and S is a countable subset of [0, r ] , then P({il/: yff \s is uniformly continuous}) = 1. Given 5, choose for each N >\ points 0 = to, ^ < ' *' < ^N, N = 7" so that if S^ = {r.^: 0 < i < N}, then S^ c 5^+1 and S c Q * 5^. Then' (1.13)
P({^'' ^ \s is uniformly continuous}) ^lin,.imp(Lsupl^f:if)i^-%(4AH), A/QoN^Qo\! s,teSN r~'^l y —-^ !/
where 7 and )9 are chosen as in Corollary 2.1.4. But if ^^^\') is defined by Nu\ (^f+l,Ar - t)\p{ti,f^) + (t - ti,N)w{U+UN) \p [t) — , ti+l,N — ti,N
^ ^ , ^ , ti^N < t < ti+iN ,
2.2. Markov Processes and Transition Probabilities
51
and \l/^^\t) = il/(T) for t > T, then ij/^^^') is continuous and an easy calculation shows that (1.12) implies E[ I iA<^>(0 - il/^''\s) n < 3TI r - s P ^^
0 < 5 < f < 7.
Thus, Corollary 2.1.4 applied to ij/^^^ yields PIL.
sup
lMz:^^A(4Ap|)
and so we see that the right hand side of (1.13) is one.
D
We now have all the ingredients necessary for the following theorem of A. N. Kolmogorov. 2.1.6 Theorem. Let {Pti,...,t} ^^ ^ consistent family of finite dimensional distributions. If for each T > 0 there exist numbers a = a r > 0, r = r^ > 1 + OCT, and CT < oo such that (1.14)
j ^ \y - xfP,^,(dx
X dy) < C^^l? - s\^^\
0<s
then there is a unique probability measure P on (Q, Ji) such that
(1.15) for alio
p(x{t,) G r „ ..., x(0 E r„) = p,,.... j r , x ••• x r„) ••• < t„ and Tj, ..., r„ e ^R,.
Proof Since J^ = a[x(t): t > 0], the uniqueness is obvious. To prove existence, we proceed as follows. First construct P on ((R*'}^' °°\ ^^Rd)io, «>), corresponding to {^ti, ...,r„}» ^s in the first paragraph of this section. Then (1.14) implies (1.12); and therefore, by Corollary 2.1.5 and Lemma 2.1.1, P determines a restriction F to (Q, ^(/?^)[o, oo)[Q]) given by F(A n Q) = P(A), A e ^(^^[o. «>. Since ^ = J^(K
2.2. Markov Processes and Transition Probabilities In the theory of Markov processes, the consistent family {Pn, ...,tj arises in a special way. To be precise, we define a transition probability function as a function P(s; x; r, F), 0 < s < r, X e R**, and F e ^^d, satisfying: (i) P(s, X, t, •) is a probability measure on (R^, ^^d) for all 0 < 5 < t and
52
2. Markov Processes, Regularity of their Sample Paths, and the Wiener Measure
(a) P{s, •, r, r ) is ^;jd-measurable for all 0 < s < r and V e ^^d, (Hi) ifO<s
P(5, x;u,r)
= j P(u y\ w, T)P(s, x; r, dy).
Equation (2.1) is known as the Chapman-Kolmogorov equation. A transition probability function should be thought of as giving the conditional distribution of a process at time t given that at time s the process was at x. It turns out that, even if one specifies the initial distribution of the process, a stochastic process is not uniquely determined by insisting that it have the preceding property relative to a given P{s, x; t, •). However, if one goes one step further and demands that P{s, x; r, •) be the conditional distribution of the process at time t given the process before time 5 and that at time s the process was at x, then the process is uniquely determined as soon as its initial distribution is given. We now have the following formal definition. 2.2.1 Definition. Let P(s, x; r, •) be a transition probability function and /i a probability measure on (R^ ^^j^). A probability measure P on ((R**)^®' *^ ^(Rd)io.oo)) is called the Markov process with transition function P(s, x; r, •) and initial distribution fx if (2.2)
F(iA(0) G r ) =/z(r),
Te^Rd-.
and for all 0 < s < r and r e ^^^i (2.3)
P(iA(r) e r | G[^(U): 0 <
M
< 5]) = P(s, 1/^(5); t, Y) (a.s., F).
If P(s, X; r, •) = P(t — 5, X, •), the corresponding Markov process is said to be time-homogeneous. We now turn to the question of the existence of a Markov process with given transition function and initial distribution. 2.2.2 Theorem. Let P(s, x;t,')be a transition probability function and /j, a probability measure on (R*^, ^^d). Define (2.4)
Fo(r) = //(r),
Te^Rd,
and for 0 < f ^ < • • < r„+1: (2.5)
P^„...,w.(A) = J.^- JP(tn,yn;tn+udyn+i)Ptu...,tn(dyi X ... X dy^),
where A e ^^Rd)N+\. Then^ {F,j ^J is consistent; and a probability measure P on ((/?'*y°' ®\ ^(j?d)io.oo)) is the Markov process with transition function P(s, x; r, •) and
2.2. Markov Processes and Transition Probabilities
53
initial distribution fi if and only if P is the probability measure on ((R^f^' ®\ ^(Rd)io,oo)) having {Pti,...,t„} as finite dimensional distributions. In particular there exists one and only one Markov process with given transition function and initial distribution. Proof The consistency of {P,j J is immediate from the Chapman-Kobnogorov equation (2.1). Now suppose P on ((R^f^''^\ ^^Rd)io,oo)) has finite dimensional distributions given by (2.4) and (2.5). Given t > 0 and T e ^^i, it is clear that P(iA(0) G A, ^(t) e r ) = [ P(0, y; r, r)Fo(^>') •'A
= £^[F(0,tp(0);^,r),tp(0)GA], and therefore: P(iA(r) e r I (7[iA(0)]) = P(0, ,A(0); ^ r )
(a.s., P).
Next let 0 < 5 < r and r e ^^a be given and suppose that 0 < MJ < • • • < M„ = 5 and T i , . . . , r„ e ^^d are chosen. Then
p(^A(t/i)eri,...,^A(«„)er„,^(r)6r) = ^«,....,„„,r(ri x " x r „ x r ) \\
Pis;y„;t,r)P
rix...xr„ 'PI = £''[P(s, ^s); t, n
il,(u,) € r „ ..., ^(u„) 6 r j .
Thus an easy application Exercise 1.5.1 implies that (2.3) holds. Finally, assume that P on ((R^f^' *\ ^^Rd)io,oo)) satisfies (2.2) and (2.3). We want to check that its finite dimensional distributions {Q,j^ ^,J are {P^^^ J. Clearly (2.2) implies Qo = PQ- TO complete the identification we use induction on n. If n= 1 and t^ = 0, we have aheady checked Q^^ = P^ j. If n = 1 and ^i > 0, then for T i e ^Rd:
Q,,(r,) = P(iA(ri) e r j = £^P(0, iA(0); r^, T,)] = J P(0, >;; r„ r,)Qo(^y) = | m
yi h. r,)Po(dy)
= ^r.(r,). Now assume that Q,, / : (Ry -• K, we have
t = Pti
t - Then for any bounded measurable
54
2. Markov Processes, Regularity of their Sample Paths, and the Wiener Measure
In particular, if r^, ..., r„+ ^ e ^^d and
/(yi. •••. yn) = xu(yi) ••• XrSyn)P(tn^ 3^n; t„+u r„+i), then, by (2.3),
\\
P(tn^ yn; ^«+i, ^yn+i)^ri
J^yi x ••• x ^y„)
rix...xr„+i
= Pu,...,t„.A^i x . . . x r „ ^ i ) . Thus the proof can be completed by another application of Exercise 1.5.1.
D
Of course, there is no reason why a Markov process should always have to be realized on ({R^f^- °°^ ^(R4)io.ao)). In fact, we want the following definition. 2.2.3 Definition. Let (£, J^, P) be a probability space and {J^^: r > 0 } a nondecreasing family of sub
P ( ^ ( 0 ) G r ) = /i(r),
Te^Rd.
and (2.6)
F((^(r) G r IJ^,) = F(s, ^(5); r, F)
(a.s., P)
for all 0 < 5 < r and T e ^^d. Notice that if £ = (R'^y^- °">, J^^ = (^R^T' '^\ and J^, = (T[II/(U): 0 < M < r], then the preceding definition is consistent with the one given in 2.2. L The case in which we will be most interested is when E = Q^^ = Ji,^^ = ^^, and ^(t) = x(t). In fact, if (x(r). Jit, P) is a Markov process on (Q, Jt\ we will call it a continuous Markov process. 2.2.4 Theorem. Let P(s,x; t, •) be a transition probability function such that for each T > 0 there exist a = a j > 0,r = r^ > 1 + a r , an^ C = CT for which (2.7)
sup [ \y - yil'Pit,,
yu t2. dy)
0
2.2. Markov Processes and Transition Probabilities
55
Given a probability measure ^ on (R^, ^Rd\ there is a unique probability measure P — P^ on (Q, J^) such that (x(t), M^, P) is a continuous Markov process with transition function P(s, x;t,') and initial distribution fi. In particular, for each s > 0 and X E R^, there is a unique probability measure P^ ^ on (Q, Jt) such that (2.8)
P,, ^(x{t) = X for all
0
and (2.9)
P , , M h ) e r I ^ , J = P(t, x(t,); t2, r )
(a.s., P,, J
for all s
Proof Let Q on ((R'^f^' *^ ^(Rd)[o, «>) be the Markov process with transition function P(s,x;t,) and initial distribution fi (cf. Theorem 2.2.2). By Theorem 2.1.6, Q admits a restriction P^ to (Q, J/) having the same finite dimensional distributions. Since J^^ = ((T[\I/(U): 0 < M < t])[Q], it is obvious that P^ is the desired measure. (The uniqueness of P^ is clear, since its finite dimensional distribution must be those of Q.) Finally, suppose 5 > 0 and x e P** are given. Define
I
Sy^
if0
Pis,yi;t2.-)
ifti
P(ti.yi;t2,
•)
<s
if5
Then F(ti, ^i; ^2, •) satisfies (2.7). It is easy to check that the desired P^^c is the unique probabihty measure on (Q, J^) such that (x(r), J^^, P^^^) is a continuous Markov process with transition function F(sy x; t, •) and initial distribution S^. • A particularly important apphcation of Theorem 2.2.4 is to the situation in which (2.10)
P(s, x; r, r) = f g,(t - s, y - x) dy, •'r
where (2.11)
J.
^.(^'^) = 7^^"'^''^'^-
It is an elementary exercise to show that (2.7) is satisfied in this case with a = 1, r = 4, and C = d. Thus we have the following famous theorem of N. Wiener.
56
2. Markov Processes, Regularity of their Sample Paths, and the Wiener Measure
2.2.5 Theorem. For each s > 0 and x e R*^ there is a unique probability measure iirf\ on (Q, M) such that irjf,(x(r) = x,0
G r I ^ , J = f g,(t2 -h,y-
x(t,)) dy
(a.s., iTf)
for all s
Then TTg = iT^^^ o T'^K Also, if {ex, ...,ed} is an orthonormal basis in R^ and %i : R^ -^ R is defined by ntx = (x,e,), 1 0] are mutually independent under P and P 071^^ = i^^^^ for 1 < i < d. Finally, the following are equivalent statements about a probability measure P on (Q, J^) : (i) p = ^c^), (a) £^e'<«'^<'»] = e-^'^l'/^^ for all t>0 and Oe R^ and M, is independent of G\x\f) — x{s)\ t > s] under P,for all 0 < s < 00, (Hi) for alln>UO
2.3. Wiener Measure This section is devoted to the development of some of the important properties of Wiener measure. For reasons which will become clear in Chapter 4, we want to couch our discussion in slightly more general terms. Thus we introduce now the next definition. Definition. Let (£, ^, P) be a probability space and {J*^,: t > 0} a nondecreasing family of sub tr-algebras of ^. Given 5 > 0 and a function j8: [0, 00) x E -• R^, we will say that (P{t), ^^, P) is a d-dimensional s-Brownian motion (alternatively, when there is no need to emphasize d or {J^,: r > 0}, ^(•) is an s-Brownian motion under P) if
2.3. Wiener Measure
(/) (a) (in) (iv)
57
P is right-continuous and progressively measurable after time s, p is P-almost surely continuous, P(P(t) = 0 for 0 < t < s) = 1, for all 5 < ^1 < ^2 and F e ^^d P(p(t2) 6 r I J ^ , J = f g,(t2 - t , , y - Pit,)) dy (a.s., P), •'r where g^ is given in equation (2.11).
If s = 0, we will call {P{t\ ^ , , P) a Brownian motion, Clearly (x(r), J^t, i^%) is an s-Brownian motion. In fact, (x(t\ M^, 1^1%) is the canonical s-Brownian motion in that if (P(t\ ^t, P) is any s-Brownian motion and P o j?~ Ms the distribution o(p(') under P on Q (note that by (n), q^ p(', q)isa. map of a set having full P-measure into Q, and therefore P o p~^ {$ well-defined on (Q, e/#)), then P ^ jS" ^ = T^^f\). The next lemma gives a partial answer to the question of why one likes to consider other versions of Brownian motion besides the canonical one. 2.3.1 Lemma. Let (£, ^, P) be a probability space and (p(t), ^^, P) an s-Brownian motion. Denote by #'(J^,) the completion of^(^t) under P and use P to denote its own extention to ^ . For t > 0, set J^,+o = n<5>o ^t+d- Then (P(t), ^ , + 0 , P) is again an s-Brownian motion. Proof. Obviously, all we have to check is that P{P(t2) e F | ^^j+o) = jr 9d(h — h^ y - Pih)) <^y (a.s., P) for all s < tj < ^2 and F e ^j^^. To do this, it is certainly sufficient to show that for all 0 e CQ(R^) and A e ^tx + o^ (3.1)
E\(P(t2)\A] = E' \gd(t2-t,.y-P{t,))(i>(y)dy,A
Note that (3.1) is obvious if ^ € ^ij. If .4 € #ii+o, then, since A e J^,+e for all £>0,
E'[(t>{P{t2 + a)), A]- =T7PE \ 9d(t2 -tuy-
P(ti -f e))(P(y) dy, A
Since j5(-) is right-continuous, we can now let e \ 0 and thereby get (3.1).
D
We now want to prove one of the basic properties of Brownian motion, namely: "it starts afresh after a stopping time." The first step is the following lemma.
58
2. Markov Processes, Regularity of their Sample Paths, and the Wiener Measure
2.3.2 Lemma. Let (P(t), ^^, P) he an s-Brownian motion. Given to > s, define PtoiU Q) = P(t + ^0» ^) ~ P(h, q\t >0 and q e E. If
£"[4) o ^,„ I ^ , J = E^"\
(a.s., P).
Proof. We need only prove (3.2) for
CO G Q ,
where 0 < fj < ••• < t„ and 01, .., 0„ e Q(i^'^). That is, we must show that if tQ>s, Ae J^jo' n>l, and (^i, ...,>„ e Q(R''), then for 0 < fi < • • • < r„: (3.3)
£l(/>i(Ao(^i)) • • • UPto(tn)\ A] = E^'\ct>Mh)) • • • >„(x(t„))]F(^).
To this end, let w > s and <j) e Q(/?'') be given; Then for D > M and A € ^ „ : E-'WM'
A] = f El
q))\^MP(dq)
= f j f 4>(y - P{u, q))gAv, y - P(u, q)) dyy(dq) = ( j
E'{cj>{pM \^u] = \
(a.s., P).
In particular, (3.4) proves (3.3) when n= \. Next suppose that (3.3) holds for n. Let 0 < fi < ••• < r„+i and (/>i, ..., (/>„+i e Cfc(R'') be given. Applying (3.4) to M = r„ + ro, u = r„+1 - r„, and 0(>') = 0„+ i(y + z), we have: £:''[^„+i(A„+J^n+i-0 + ^)l^r„+J
= I 0n+ i{y)9d(tn^ ^-t^,y-z)dy
(a.s., P).
59
2.3. Wiener Measure
Thus, since (3.3) holds for n: £l0i(/J,„(ti))-<^„.i(A„(t„.i)M] ''A
^UKKq))-
''A
X j (l>n+i(y)9d(tn+i -t„.y-
x ( 0 ) \p(A)
= £n>i(^(^i)) •• 0„(x(r„))(/>„,iWr„,O)]P(A) Thus the induction is complete.
D
2.3.3 Theorem. / / {P(t\ i^,, P) is an s-Brownian motion and T is a stopping time satisfying T > s define p^(') by: \P(t^
if T(g) < O)
10
if
PXUq)T(q)=oo.
Then for A e ^^ and O a bounded J^-measurable function on Q into K, we have (3.5)
E^[
n (r < oo}).
In particular, (3.6)
E^[a)o^Jj^^] = £^ <"»[(!)] on
{i < oo}
(a.s., P).
Proo/. It is certainly enough to check (3.5) when 0 is a bounded continuous function on Q. But in that case, £''[0) o j5,, A n {T < oo}] = lim f^O) o j5,„, /I n {T„ < oo}],
60
2. Markov Processes, Regularity of their Sample Paths, and the Wiener Measure
where s -\- ^^ T„=
if
X < CO
if
X = CO.
n ' 00
But A e ^^ implies A n {T„ = s -{• (k/n)} e ^s+(k/n) for all /c > 1, and so E'[
Y.E k=
fc=l
\
\
"1/
= £^^10]F(.4 n {T„ < oo}). Since {T„ < oo} = {T < oo}, this completes the proof. Q
2.4. Exercises 2.4.1. Extend Theorem 2.1.3 in the following way. Let p and ^ be strictly increasing continuous functions on [0, oo) such that p(0) = ^(0) = 0 and lim^i^o ^ ( 0 = ^ Let L be a normed linear space and/: R^ -^ La. function which is strongly contin uous on B(a, 2r) for some a e R^ and r > 0. Show that
Wr)Va,r)
\
p{\x
-
y\)
J
implies that 2\x-y\
\f(x) -f{y)\\ < 8 (^
Ud+lj^\
^\-^y(du\
X, y e B(a, r),
where 7 = 7^=
inf
inf
B(x, p) n B(0, 1) .
xeB(0, 1) 0
P
The proof mimics that of Theorem 2.1.3. Set
Given distinct points x, y e B(a, r) set p = \x — y\ and proceed as follows: (a) Choose c e B{(x + y)/2, p/2) n B(a, r) so that
and set
XQ =
)^o = ^»
2.4. Exercises
61
(b) Given x„_i and y„-i, define d„-i and e„_i by: p(4-i) = i p ( 2 | x „ _ i - x | )
and p(^„-i) = ip(2|y„_i - yl),
respectively, and choose x„ e B(x, X _ i) n B(a, r),
>;„ e B(};, ie„_ J n B(a, r)
so that /(x„) < 2"-'B/yd^,
and 4 ' ( M ? ^ ^ ^ ^ ^ j < 2"- '/(x„. 0 / K - .
/(y„)<2''-'B/y^..
and ^(''p}!;!,!^/;";;!*)^'' ) ^ 2"" '/(j'„- .V?^-.-
(c) Conclude that
ll/(y)-/(c)|| < 41^ ^ P - ' I ^ J P W 2p
/4''+iB\
||/(x)-/(c)||<4j^ ^-|_^jp(d„). 2.4.2. Using Corollary (2.14), show that if ^ is a set of probability measures on (Q, J^) such that lim supF(|x(0)| >L) = 0 and for each T > 0 there exist a^ > 0, r^ > 1 + ar, and Cr < oo with sup E^[ Ix(t) - x(s) l"^] < Cr{t - s)^ •'^
0 < s < r < T,
then ^ is precompact. This observation suggests the following derivation of the existence of W^'^K Let {X„: n> 1} be independent Z?*^-valued normal random variables with mean 0 and covariance / on some probability space (£, J^, P). Given w > 1, define 0„: £ -)• Q by:
x(r, 0„(^)) = ^ ( | x , 4- «(r - ^ ) ^ M . i ) , (where ^ i ^fc = 0), and set P„ = Po^;K
Check that
sup E^'i I x(t) - x(s) n < C(t - s)\
0<s
62
2. Markov Processes, Regularity of their Sample Paths, and the Wiener Measure
for some C < oo and that the finite dimensional distribution of P„ coincide with that of i^^'^^ Sit times k^/n, ..., ki/n, where / > 1 and 0 < /c^ < ••• < /c,. Hence conclude that P„ -• i^^'^^ as n -> oo. 2.4.3. Prove Theorem 2.2.6. 2.4.4 Prove that if P(') is any Brownian motion, then
E[\m-p(s)n==c,\t-s\\
o<s
for any r > 1. Deduce from this that for any 0 < a < 1/2, p(') is almost surely Holder continuous with exponent a on any finite time interval. 2.4.5. Given A > 0, define 5^: Q -^ Q by: x(r, 5^co) = A" ^'^x(h, (o). Show that ir^"^^ is invariant under S^ (i.e., iT^"^^ = iT^'^^ « S^^). Using this fact, prove that a Brownian motion is almost surely not Holder continuous with exponent 1/2 even at one time point. With a little more effort one can show that Brownian motion has an exact modulus of continuity: (231 log ^ | Y'^ (cf. McKean [1969]). 2.4.6. Let a > 0 be given. Using Theorem 2.3.3, derive the following equality. ir^^\x(t) >a) = ii^^^^\x(t) >a,T
= iiT^^^z < t)
where T = inf{t > 0: x(t) > a}. Conclude that ds. ir^'Hr
(inyi'l
Next, use this formula to derive the joint distribution under i^^^^ of x(t) and T. Finally, show that ir^'H sup |x(s)| >a\<2dl~\ \o<s
I
—e'^^/^rd
\7C/
a
2.4.7. Observe that by the strong law of large numbers, lim
p(n) = 0
(a.s., P)
if (P(t\ ^t, P) is a ^-dimensional Brownian motion. Combining this with 2.4.6, show that lim tp(l/t) = 0
(a.s., P).
63
2.4. Exercises
Use this, along with Theorem 2.2.6, to check that if
at) = tp(i/t\
t > 0,
then the distribution of (^(•) under P is again iT^'^K That is, ((^(r), J^;, P) is a Brownian motion, where ^[ =
fiS sup S-^0 0<s
P ^ ^ ^ l
= 1.
/')c i^_ A I
|r-.|^. ^2(51og-j
The difficult part of the result is to prove that r^
(4.1)
ir^'^ lim — \-*0
\x(t)-x{s)\
sup -"*-
= 1. 4-^—-f^ / . w / 2 <^ 1
0 < s < t < 1 I -)c i _ „
A I
|r-s|^<5 ^2^l0g-j
The other part is a simple estimation of the probability of f]i=i A,^(p), where A^(p) are the independent events A,(p) =
4)-^-^)
< (1-^X2-"^^ log2")i/2
and 0 < p < 1 is arbitrary, (cf. McKean (1969)). It has been pointed out to the authors by R. Wolpert that a slightly weaker version of (4.2), namely (4.2)
^(1)
iim imi
sun sup
^iO 0 < s < r < l
M L : --YJ2 i ^ <^ ^8 = 1, i^c
i^„
A I
\t-s\<S (2^l0g^j
can be deduced from Theorem 2.1.3. To this end, take il/(u) = e" - 1 and for fixed a > ^ take p{u) = (aM/(log l/u)y'^. Check that E^'
I0:*(^af)"'i
<
00.
Next, apply Theorem 2.1.3 to show that lim ir^^^
1 + log :(t) - x(s)\ < 4a^'- j ' " ^ -^ log(l + ^ ) ^ ^ dw
fltoo
0 < s < r < 1 = 1.
64
2. Markov Processes, Regularity of their Sample Paths, and the Wiener Measure
That is to say, there is a finite random variable B such that 1
lo -
'" ^ 'K) for all 0 < s < r < 1 with r - s < ^. By L'Hospital's rule 1 + log
"•"
/
l\l/2
(.logl)" and therefore — hm
sup
\x{t)-x{s)\ 16a"2 L1I—1^<
almost surely. Since a > ^ is arbitrary, this implies (4.2). If one is willing to replace 8 in (4.2) by 16, then one can take €>(«) = e"^'^ and p(u) = otu^'^ for some a > 1. The details are somewhat simpler in this case.
Chapter 3
Parabolic Partial Differential Equations
3.1 The Maximum Principle In Chapter 2, Section 2.2, we showed how one can start with a transition probability function P{s, x; r, •) and end up with a Markov process. The problem is: where does P(s, x; r, •) come from? The example we gave there, namely: (1.1)
P(s, x; t,r)=
\ QII
-s,y-x)dy
•'r
is a natural one from the probabilistic point of view because of its connection with independent increments and Gaussian processes. It turns out to be natural from another point of view as well: the theory of second order parabolic partial differential equations. The connection between the P(5, x; r, •) in (1.1) and partial / G CY,{R^) and differential equations is well-known and easy to derive. Namely, if <> /(s, x) = j glT -s,y-
x)(p{y) dy, s < T,
then (1.2)
\ lim /(5, ') = (p
whe re A is Laplace's operator
i'dxf That is, / solves the backward heat equation with terminal data (p at time T. Besides the kinetic theory of gases and Einstein's famous articles on Brownian motion, there is a purely analytic reason why it is not surprising that the heat equation should be the source of a transition probability function. This reason is the " weak maximum principle " for parabolic equations.
66
3. Parabolic Partial Differential Equations
3.1.1 Theorem. Let a and b be bounded functions on [0, oo) x R'^ with values in S^j and R'^, respectively. Define
IffeC^'
^([0,
T) X R^) is bounded below and satisfies:
(1.4)
|^ + L J < 0 , 0 < 5 < r lim /(s, •) > 0,
then f is non-negative. In particular, iffeC^' ^([0, c,ge Q([0, T)) satisfy: (1.5)
f^ + L J + c(s)f> -g(sl
T) x R'^) n C„([0, T] x R**)
0 < s < T,
then for 0 <s
/(5, x) < ||/(r, .)||exp|jjc(u) du^ + fj^(r)exp||^(u) t/t/j^L
Proof. First assume that / satisfies (1.7)
f^ + L J < 0 ,
0<s
OS
If (so, XQ) G [0, T) X R** is a point with the property that (1.8)
/(5o, xo) (5, x),
(5, x) e [so, T) x R'',
then we would have: df ^ (so, Xo) > 0,
V^/(5o, Xo) = 0, and
((aS;^^»'^»0),„.<_/'^' where V^^ denotes gradient in the x-directions. Since ^so/(5o. ^o) = X Trace(a(so, Xo)H^(so,
XQ))
+
V^/(SO, XO)>,
and
3.1. The Maximum Principle
67
where
«*""-((sS-'"')),.,„., is the Hessian matrix of/, we conclude that: ^ ( s o , ^ o ) + ^5o/(5o,^o)>0,
since Trace(^B)>0
if
A,BeS^.
But this contradicts (1.7), and therefore there is no (s©, XQ) e [0, T) x R*^ for which (1.8) holds. Next suppose that / satisfies (1.4). Given S >0 and g > 0, set fs,e(s.x)=f(s,x)-hS(T-s)-h8e-^\x
12
Then ^fs,s
Sf
^5
^5
. __-., .2 S — se ^\x
and J-sfs, e = -LJ+ se ^(Trace a(s, x) 4- 2(b(s, x), x » . Hence, ^
+ LJ,,, < - ^ + £e-^(Trace(a(s, x)) + 2 < 6 ( 5 , x ) , x > - |x|2).
Therefore, for each ^ > 0, we can choose e^ > 0 so that/j ^ satisfies (1.7) for e < £^. Now suppose that /^ ^(s, x) < 0 for some (s, x) € [0, T) x R'^. Then, since lim.^r f^ As, ')>0 and /^^^(s, x)-^ +oo as |x| -• oo, there must be a point (so, Xo) e [0, T) X R*^ at which (1.8) obtains, and this is impossible if (1.7) is to hold. We have therefore proved that for all S > 0 and all 0 < e < e^,/^ ^ > 0 in [0, T) X R'^. From this it is clear t h a t / > 0 in [0, T) x R'*. Finally, suppose / G C^'^{[0, T) x R'^) n ^([0, T] x R^) and c,g € ^([0, T]) satisfy (1.5). Then \\f{T,)\\-f{s,x)Qxp{-J satisfies (1.4), and so (1.6) follows.
c(u)du)-\- f g{t)Qxpf- f D
c{u)du\dt
68
3. Parabolic Partial Differential Equations
We will make repeated use of Theorem 3.1.1 throughout the rest of this chapter. However, the reader should notice that Theorem 3.1.1 is a uniqueness theorem, and therefore it is not very powerful except in conjunction with an existence theorem. Combined with an existence result one can then prove the following: 3.1.2 Corollary. Let L^ be defined as in Theorem 3.1.1. Assume that for each t>0 and (p e CS(R'') there is an fe C^'2([0, t) x R*^) n Cj,([0, t] x R'^) such that f(t, ')^(p and (df/ds) + LJ = 0, 0 < s < r. Then for each t>Oand(pe C^(R'^) there is exactly one suchf. Moreover, if for 0 <s < t and x e R'^we define A,, ^(x) on CQ^(R'^) into R^ by As,t(x)(p = /(s,x), then As,t{^) determines a unique non-negative linear functional on C(R') =Le
C(R'): lim (p(x) = 0 1^1-
such that \As,Mcp\ Finally, if we define T,t,0<s
(peC(R').
<\\(PI
C(R'^) by
then 7^, ^ is a non-negative contraction on C(R'^) into itself lim^^^ 7^ ,(/)(x) = (p(x)for allxeR'^^and Ttuh = Tti,t2 o Tt2,ti ,
0 < ti < ^2 < ^3 .
Proof. The uniqueness o f / i s immediate from Theorem 3.1.1. In fact, from that theorem we see that min<^(x) < / ( s , •) < max(/)(x),
0 <s
This proves that A5^(jc) is a non-negative linear functional on Co(R'^) and that
\K.{xyp\<MSince Co(R'^) is dense in C(R% this completes the proof of the assertions about
KMIn order to prove that 7^, < is a non-negative contraction on C(K'') into itself, all we have to show is that T.^.cp e C(R'') for cp e C^(R''). Since, if cp e CS(R% / ( s , x) = Tst(p(x) is the unique C^' ^([0, t) x K'^)-function tending to (p as s/'t, it only remains to check that such a function must tend to 0 as |x | -• oo. Thus, the proof reduces to showing that if M > 0 and cp^ e CoiR*^) satisfies: 0 <(pM
3.1. The Maximum Principle
69
/M e Ct 2([0, t) X R') n Q([0, r] x R') satisfying (a/M/^s) + L, /M = 0, 0 < s < r, with/jvf(r, •) = (PM» has the property that/M(s, •) e C(R'') for all 0 < s < r. To this end, note that we can find positive numbers A and B such that for all XQ G R*^ the function : il/Js, x) = Ae^^'-'\t - 5) + e'^'-^^ \x - Xo|^
(s. x) e [0, t) x R^
satisfies (dij/^^/ds) + L^I/^^Q < 0, 0 < 5 < r. Thus if p = ^ |xo | > M + 1, then, since P^
0<s
In particular, A
lA
0 ( t - s) = ^ «»<'-»(' - s) P 1^0 I for I Xo I > 2M -h 2, and this completes the proof that T^ , cpj^ e C(R''). Finally, we must show that
To this end, let (p e C*(R'*) be given and set/(5, •) = Tg, t
and lim (/(s, x) -/„(5, x)) =f(t2.
x) - (p„{xl
X 6 R'.
s/t2
Thus, by Theorem 3.1.1, sup
\\f{s,.)-f„(s,-)\\<\\f(t„-)-,p„(-)\\-^0
0<S
as n-^ 00. That is: 7;,. ra > = f(tu •) = lim Utu •) = lim 7;,,,, (^„ = 7;,,,, o 7;^ ^^ <^, n-»oo
which is what we needed to show.
n-*oo
D
70
3. Parabolic Partial Differential Equations
A two-parameter family {T^ ,: 0 < s < t} of operators on C(R'^) having the properties given in Corollary 3.1.2 is called a time-inhomogenous semi-group of non-negative contractions on C(R^). It is clear that if the linear functional A^ ,(x) given by A^ t(x)(p = 7] , (p(x) admits the representation (1.9)
A,^ t(x)(p = f P(s, x; t, dy)(p(y%
0 <s
and
x e R',
where P(s, x; r, •) is a probability measure, then P(s, x; r, •) is a transition probability function. Our next result shows that the {T^ ,: 0 < s < r} constructed in Corollary 3.1.2 admits a representation of the sort in (1.9). 3.1.3 Corollary. Under the same assumptions on L, as those in Corollary 3.1.2, there exists a unique transition probability function P(s, x; t, •) such that (1.9) holds. Moreover, there exist numbers A and B, depending only on the bounds on the coefficients of L^, such that: (1.10)
\ \y-^
r^(5, ^; t. dy) < Ae^^'-'\t
- s)\
0<s
and
x e R'.
Proof. As we said in the preceding paragraph, the existence of a transition probability function P{s, x; t, *) will be established once we show that for each 0 < s
A., tM(p = \ (p(yHdyl
q> e C(R').
We must show that /z(j?^) = 1. To this end, let cp^ e CS'(R'^) be chosen so that 0 < (PM < 1, <)Pjvf = 1 on B(xo, M), and
3.2. Existence Theorems
71
There is a choice of A and B, depending only on the bounds on the coefficients of L such that dy ^ 4- L,y < 0, 0 < 5 < r. Thus, if {(p„}f ^ Co{R^) is a sequence of non-negative functions such that (p„(x)/'\x — XQ |^ X G R**, then, by Theorem 3.1.1, Ts,t(pn
0<s
In particular, j (p„(y)P(s. xo; U dy) = r,,,(^„(xo) < Ae^^'-'\t - sf. The estimate (1.10) now follows from the monotone convergence theorem.
G
3.2. Existence Theorems Theorem 3.1.1, and its corollaries, appears to be a good mill with which to turn out transition probabihty functions. However, like any mill, it requires grist before it can produce; and the grist in this case comes from the theory of partial differential equations. The following theorem is of just the sort that we need; in fact, it gives us more information than is required. We state it here without proof, because the proof is quite intricate (the parametric method is the one usually employed) and we will not be relying on it for anything outside the present section. A good derivation can be found in the book of A. Friedman [1964]. 3.2.1 Theorem. Let a: [0, oo) x R** -* S^ and h\ [0, oo) x i?** -• R^ he hounded functions for which there exist numhers OL>0\0
^^-lif(-^)8^^^i''(-^)iThen there exists a unique positive function p(s, x; t, y), 0 < s < t and x, y e R**, which is continuous jointly with respect to all its variables and has the property that if (p e C^iR'^) and g e Cf ([0, 00) x R% then for each t > 0 the function (2.1)
f(s, x) = I p(s, x; r, y)(p(y) dy-\- j du ^ p{s, x; u, y)g(u, y) dy
72
3. Parabolic Partial Differential Equations
is in C5' ^([0, r] x R^) and satisfies
withf(t, ')=^ (p. In particular, L, satisfies the conditions of Corollary 3.1.2, An easy consequence of Theorem 3.2.1 is the following. 3.2.2 Corollary. Let L, be given as in Theorem 3.2.1 and let P(s, x; r, ') be the associated transition probabilityfiinctionguaranteed by Corollary 3.1.3. Then for any 0 <s
F
Rd,
where p(s, x; t, y) is the function described in Theorem 3.2.1. Also, if t >0 and feCt ^([0, t) X R') n Q([0, t] x R% then (2.2)
jf(t,y)P(s,x;t,dy)-f(s,x)
Proof The identification of p(s, x; t, •) as the density of F(s, x; r, •) is immediate from the fact that if ^ = 0 in (2.1), then the corresponding / satisfies (df/ds) + L,f=0, 0<s
\l/(s,x) =
jf(t,y)P(s,x;t,dy) d
f>„i(^^ + L„ |/(M, y)P(s, x; M, dy) is in C^' ^([0, t] X R"^) and satisfies aiA df ^ + L,iA = ^ + L,/, with i/^(r, • ) = / ( ? , •). Thus, by Theorem
0<s
0 < s < f, 3.1.1, iA(s, •) = / ( s , •) for all
D
The rest of this section is devoted to the derivation of a result due to Oleinik. The point of Oleinik's theorem is to get existence theorems when the coefficient a is degenerate (i.e., (i) in Theorem 3.2.1 fails) but there are more stringent smoothness requirements on a and b. We will need the following lemma.
3.2. Existence Theorems
73
3.2.3 Lemma. Let a: R^ ^ S^ be a function having two continuous derivatives. Assume that Xo = sup{|D^a''^(x)| : 1 < ij
and
xe R^}
and x e R^: \iDa''^)(x)\ < (2Ao)i/2(a"(x) + a^^M)^/^
Moreover, if u is any symmetric d x d-matrix, then: (2.4)
(Trace((Da)(x)w))2 < 4d^XQ Trace(Ma(x)w),
xeR"^.
Proof To begin with, let (p e C^(R^) be a non-negative function such that a = supj^g^i |(p"(x)| < 00. Then for any x E R^ and all y e R \ we have, by Taylor's Theorem 0<(p(x-^y)<
(p(x) + (p'(x)y + ^ y\
In other words, the quadratic a/2y^ + (p'(x)y + (p(x) is negative for no real y. Hence, from the elementary theory of discriminants, ((p'(x)y-2oi(p(x)<0, and so (2.5)
\cp'{x)\<{2ay'Hcp{x)r'\
We now apply (2.5) to the functions (p±(x) = a"(x) ± 2a'\x) + aJ\x). Since (p + (x) = <e, ± e^, a(x)(e, ± Cj)}, where {e^, ..., Cj} is the standard basis in R^ \(p'±(x)\ <4/lo. Hence
\cp\(x)\<(Uoy'HcpAx)y". But a'^(x) = i((p+(x) —
<4'%,W +
74
3. Parabolic Partial Differential Equations
To prove (2.4), we can assume that a(')is diagonal at the point x in question. We then have, by Schwarz's inequahty: (Trace((Da)Wt/))2 = [ ;^((Da)M)'^M'^
< d'lko i
(a%x)
+
a^\xW^f
d
= Ad^Xo Trace(wa(x)M). D Before presenting Oleinik's resuh, we introduce some standard notation from the theory of partial differential equations. Given cp on [0, oo) x R'^ into K\ let 'P.=j^
If a — (ai, ..., a^) is a multi-index of non-negative integers, we define | a | = Xi ^i and
Sometimes (p^""^ is used in place of D'^cp. Finally, for n > 0, define
i
3.2.4Theorem. L^r a: [0, oo) x R** -• S^,fe:[0, (x^)R^ -• R\ andc: [0, oo) x i?** -• R^ be bounded continuous functions, and set
Assume that a e C5''"([0, oo) x R*^) for some m > 2 and that b, c e for some n > 1. Given T > 0, cp e Q(R'^), and C^'"([0, (X))x R'^)
3.2. Existence Theorems
75
g E C?'"([0, T) X R'^l suppose that f e Cl ^([0, T] x R^) satisfies f^ + LJ+c(s,
')f=-g.
0<5
mthf(T, ') = (p. If, for some 0 < / < m A«, fe C?''([0, T] X R') n C^''-^2([0, T) x R% thenf G C^' '([0, T) x R**) an^ r/iere exist numbers Ai and Bi such that (2.6)
11/(5, Oll^^) < MWcpf^ + sup \\git,-)f^)e^'^^-'\
0 < s < r,
s
Moreover, the constants Ai and Bi in (2.6) can be chosen to depend only on /, d, and the bounds on the spatial derivatives of a up to order I v 2 and those ofb and c up to order I. Proof Given a with |a| < /, we use Leibnitz's rule to derive:
(2.7)
/?• + L, /<" + ^ X' «»<4 /:i? + E c, , /«« = - g^^\ ^
k
p
where ^ means summation over those I
Then, from (2.7), we obtain:
|a| = l /?
|a| = /
We must estimate
To this end, note that, by Lemma 3.2.3, 2
{<{ f'ijf
= | T r a c e | | ^ //<'*> j j < Ad^'X, Trace(//<'">«//<'">)
76
3. Parabolic Partial Differential Equations
where Z/^^*^ is the Hessian matrix
of /^^*> and X, = sup{ | {9, d^a/dxUs, x)ey \/\e\^: 0 e R'\{0}}. Thus
(s, x) e [0, oo) x R' and
where AQ = niaxj
\|a| = l
/
fc
where
Hence
since: t>0
We have now arrived at: (2.9)
w, + L,w + Ci w + 2 X Z Ca, /, /^^y^'^ + 2 ^ f V |a|=/ ^
|a| = /
Next, observe that
+ C3ii/(s,-jir"
> 0.
3.2. Existence Theorems
77
where C2 and C3 depend only on the bounds on the coefficients c^^ p. Also
|a|=/
+ W||g(5,-)||^'^
Using these relations in (2.9), we arrive at: (2.10)
w, + Lsw + (Ci + C2 + C3||/(s, Oll^'-^^ + \\g{s, ')f^)w
We can now use induction on / and Theorem 3.1.1 to get (2.6).
D
Theorem 3.2.4 is the basic result of Oleinik on which our existence theorem for degenerate parabolic equations turns. However, before we can apply it, we need the following addendum to Theorem 3.2.1 (cf. Friedman (1969)). 3.2.5 Theorem. Let a: [0, 00) x R"^ -^ S^ and b: [0, co) x R'^ -^ R'^ be bounded continuous functions having bounded continuous derivatives of all orders. Set
If for some a > 0, (6, a(s, x)e} > a I ^ 1^
(s, x) e [0, 00) x R''
then for each t>0 and cp e C^lR'^) there R'') n C°"([0, t) X R') such thatf(t, •) =
exists
and
0 e K^
an / e C ^ ^ ( [ 0 , r ] x
0<s
We are now ready to prove the main result of this section. 3.2.6 Theorem. Let a: [0, 00) x R'^ -^ S^ and b: [0, 00) x R'^ -^ R"^ be bounded continuous functions having two bounded continuous spatial derivatives. Set
Then there exists a unique transition probability function P{s, x; r, •) such that P(s, x; ',r) is ^(s ^fmeasurable, for all (5, x) e [0, 00) x R"^ and F e ^^d, and (2.11)
j / ( f , y)P(s, x; r, dy) -f(s,
x) = j ' ^u 1 1 £ + L„J/(w, y)P(s, x; w, dy)
78
3. Parabolic Partial Differential Equations
for all 0 < s < t, X G R*^, and f e C^' ^([0, oo) x R"^). Moreover, for each (s, x) e [0, oo) X R**, there is a unique probability measure Ps^j, on (Q, . # ) such that: Ps, xMt) = X, 0 < t < s) = 1 and
Ps,A4t2) e r | ^ , J = P(ri, x{t,); t2. n
(a.s., P,,J
for s
(e,a„(s,x)ey>oi„\e\' for all (s, x) e [0, oo) X R'^ and d e R^ (Hi) for all r > 0 ||a„(s, x) - a(s, x)|| -}- |fe„(s, x) - b(s, x) | -• 0 uniformly on [0, r] x R'' as n -• oo, (iv) there are bounds on the first two spatial derivatives of a„ and b„ which are independent of n. Define {L": n > 1} accordingly, and let {rj,,: 0 < 5 < f} be the associated timeinhomogeneous semi-group on C{R'^) given in Corollary 3.1.2. Given t > 0 and (p 6 C^(R'), set /„(s, •) = r;,,
uniformly on [0, t) x R''. This proves that for (p e CQ(R% and therefore also for (p e C(R'^), T"t(p(') converges uniformly on [0, t] x R** to a limit, which we denote by r,,,
on CiR'^y We now prove that at most one P(s,x;t,) satisfying (2.11) exists. Indeed, given t>0 and (p e C^iR^^l define /„(s,-) = T^^^^P as before. Then (2.11) holds with/„ replacing/. Since (df„/ds) -\- Lsf„-*0 uniformly on [0, t) x R'', we have
\(p(y)P(s,x;t,dy)-f„(s,x)-^0
3.3. Exercises
79
as n -^ 00, and therefore: (2.12)
j(p(y)P(s,x;t,dy)==T,,Mxy
That is, if P(s,x;t,') exists, (2.12) must hold for all cp € C(R'^), and therefore there is at most one such P{s,x;t, •). Conversely, to show that P(s, x; t, •) exists, let P„(s, x; r, •) be the transition probabihty function associated with E. By Corollary 3.1.3, there exist A and B, independent of n, such that (2.13)
j b - ^ \^Pn{s, x; t, dy) < Ae^^'-'\t - s)\
0<s
and
x e R'.
Thus {P„(s, x; r, •): w > 1} is (weakly) compact on (R^, ^^d) for each (s, x) e [0, oo) X R"^ and t > s. Since T:, ,(^(x) = I (^(^)P„(s, x; r, dy%
cp e C(R'l
and Tj" ^ T^^^ strongly, it follows that P„(s,x;r, ) converges weakly to a limit P(s,x;r, •); and for this Hmit, Equation (2.12) obtains. In particular P(s,x;r, •) is a transition probabihty function (since {T^^: 0 < s < r} is a semigroup), P(s, x; f, r ) is jointly measurable in s, x, and f (5 < r) for each F G ^^d, and [ |y - X |^P(s, x; r, dy) < Ae^^'-'\t - s)^,
0<s
and
x e R^
where ^ and B are the same as in (2.13). In view of these remarks, it remains only to prove that P(s,x;r, ) satisfies (2.11). But, by Theorem 3.2.1, (2.11) is satisfied when P(s, x; r, •) is replaced by P„(s, x; r, •) and L. is replaced by L". Since L" f-^ L. /uniformly on [0, t) x R"^, we conclude that (2.11) must be true for the P(s, x; t, •) just constructed. The existence and uniqueness of {Ps,x'' (s, x) e [0, 00) x jR**} is now an easy consequence of Theorem 2.2.4. D
3.3. Exercises 3.3.1. Let L be a linear operator from C*(P'') into CiR*^) having the following properties: (i) L is local (i.e., if (p(x) = 0 in some neighborhood of a point x^ in R'' then (a) L has the maximum principle, (i.e., ifcpe C'^(R'^) and (p has a local maximum at the point x^ then (Lcp)(x^) < 0).
80
3. Parabolic Partial Differential Equations
Under these conditions show that L must be of the form (L(p)(x]
2^
" w+z^x:^" dx
^ ^ dx; dx
where a^\x) and b^(x) are continuous functions on R'^ and {a'^(x)} is a non-negative definite matrix for each x. Proceed as follows for the proof: (1) First show that for any constant function c, Lc = 0. (2) If (pEC^(R^) has the property that for some x^ e R'^, \(p(x)(p(x^) I = 0( IX - x^ 1^) as IX - x^ I -• 0, then (by considering (p(x) + e | x - x^ |^ for suitably small e) conclude that (L(p)(x^) = 0. (3) Define ^ x ) = (L(pj)(x) and a'^(x) = (L(p,(pj)(x) - b^(x)(Pi(x) - b'(x)(pj(x\ where (p,(x) = Xj, i.e. the rth coordinate of x. Then verify by trying functions of the form Yfi ^ii^i — ^?)Y that {a'^(x^)} is positive semidefinite for each x^. (4) With the above definition of a'^(x) and b^(x) (using (2) and Taylor's formula) verify the form of L. 3.3.2. Prove the following extended maximum principle: Let a: [0, oo) x R'^ ^ S'^ and b: [0, oo) X R'' -* R'^ be bounded and continuous. Let 7 > 0 be given. Let/: [0, T] X R'^ -^ R he a. bounded continuous function such that there exists a sequence of functions {/„} in C^^([0, T] x R'^) such that lim sup I f„(t, x) -f{u x) I = 0 n-*QO
0
xeR''
and lim sup 0
0
dt
+ LJ„(ux)
<e
where L, is the operator associated with (a, b). Prove that sup I /(r, x) I < £(T - r) + sup | /(T, x) xeR''
xeRd
(Hint: Apply Theorem 3.1.1 to the functions {±f„{U x) -¥ 8'(t - T)} with
£' > £.)
3.3.3. Let a: [0, oo) X R*^ -)^ Sd and b:[0,cc)xR'^be bounded and continuous. Let a have two spatial derivatives which are uniformly bounded. Let b have one spatial derivative which is uniformly bounded. Construct a sequence b„ such that b„-*b uniformly on [0, oo) x R**, the first spatial derivatives ofb„ have a uniform bound independent of «, and each b„ has two uniformly bounded spatial
3.3. Exercises
81
derivatives. Denote by Lj"^ the operator corresponding to [a, b„] and by P„(s, X, r, dy) the transition probabihty corresponding to L\"^ constructed in Theorem 3.2.6. Show that for each 0 < 5 < r < o o a n d ( p G C(R'^) \im(r::\(p)(x) = hm [ (p(y)P„(s, x, t, dy) n-*oo
n-»oo
= (Tst(p)(x) exists uniformly in x; and show that 7^ , is an inhomogeneous semigroup on C{R^). Prove also that (7^ ,(^)(x) = J (p(y)P(s, X, t, dy) for some transition probabilities P(s, x, r, •), and the Markov process corresponding to P(s, x, r, •) can again be realized in the space C([0, oo): R'') of continuous functions on R*^. (Hint: Define for fixed (p e Co(R'') and 0 < r < 00 u"(s, x) = (P:i\(p)(x) for
0<s
Note that from the construction of P") in Theorem 2.3.6, the first order spatial derivatives of u^"^ are continuous and are bounded, independent of n. Observe that the maximum principle of 3.3.2 applies to each u^"\ The rest of the proof is similar to the proof of Theorem 3.2.6.)
Chapter 4
The Stochastic Calculus of Diffusion Theory
4.1. Brownian Motion Let (£, ^, P) be a probability space and (p(t\ J^,, P) a ^-dimensional Brownian motion. The purpose of this section is to point out some of the properties that (P(t), ^t, P) possesses in common with a much larger class of stochastic processes which we will be calling Ito processes. Since we are going to be giving a completely rigorous derivation of these properties in the more general context of Ito processes, our treatment here will be somewhat informal and proofs will not be complete. The first topic that we want to take up is that of equivalent characterizations of Brownian motion. Given 0 e R^, define ^^(x) = ^'^^' ^>, x e R*^. Then an easy calculation shows that (1.1) E'[ee(P(t2))\^n] =
eMt,))oxp
^-^-J^\0\2 0 < ti < t2 and ^ 6 R'^.
In fact, (1.1) is equivalent to (iv) in Definition 2.3.1. Another, more concise, statement of (1.1) is that (1.2)
(XiQ(t), J^j, P) is a martingale for all 6 e R'
where X,e(t) = ^xp[K9.my
+ -2\0\'tl
Thus (1.2) is equivalent to property (iv) in Definition 2.3.1. A second way of stating (iv) in 2.3.1 is the following. Starting with (1.1), we have j^E'[eMt))\^,;\ for t>ti. Thus for ta > ' i :
= -^e,(^(t,))exp
t - t
-\0
84
4. The Stochastic Calculus of Diffusion Theory
4.1.1 Theorem. Let (£, ^, P) be a probability space, {^,: t >0} a non-decreasing family of sub o-algebras of ^, and j5: [0, oo) x £ -> K** a right-continuous, P-almost surely continuous, progressively measurable function such that P(P(0) = 0) = 1. Then the following are equivalent: (i) (p(t), ^t, P) is a Brownian motion, (a) (Yf(t), ^^, P) is a martingale for allfe Ci(R% (Hi) (XiQ(t), ^t, P) is a martingale for all 6 e R^. Theorem 4.2.1'below shows that Theorem 4.1.1 is a special case of a general result. Next, assume that d — 1. We want to discuss the possibility of defining (1.5)
" (Ois) dp{s):
The problem is that j?(% q) is a.s. not a function of bounded variation. This fact can be easily seen from Exercise 1.5.6, since (P(t), ^t, P) is a continuous martingale. One can also see it from the following simple computation. Let t^^ „ = k/2", 0
k=0
Then E[W„] = 1, and a simple argument, using the Borel-Cantelli Lemma, shows that W„ -• E[W„] a.s. Hence the quadratic variation on [0, 1] of )?(•, q) over the dyadics is equal to one a.s., and therefore P(% q) is a.s. not of bounded variation. This means that we cannot take an entirely naive approach to the definition of (1.5). To see how one can get around this sticky point, suppose that 0 is a smooth function on [0, oo) and think of (1.5) as being defined through an integration by parts. Then i'o 0(s) dp(s) = p(t)e(t) - p{0)e{0) - ^'o P(s)0'(s)ds. An elementary calculation shows that E\j'o e(s) dp(s)] = 0 and E[(l'o e(s) dp(s)y] = ^'o e^(s) ds. Hence 6 -^ jo 6(s) dp(s) establishes an isometry from l3([0, t]. A) into I}(E, P): where X is Lebesgue measure. One can therefore extend the definition of jo 0(s) dp{s) to cover all 6 e L^([0, t], A). This procedure was first carried out, and used with great success, by N. Wiener. The situation is somewhat more comphcated when 9 depends on q as well as s. To illustrate the sort of phenomenon encountered here, consider 6(s, q) = ^(5, q). Suppose first that we attempt to define
(1.6)
f' p(s) dp(s) = lim ' x 'Pih, nMh^ 1. „) - Pih, „))• •'0
n^oo k = 0
85
4.2. Equivalence of Certain Martingales
By elementary manipulation, we have that
Xm,„Mr,,,,„)-^(r,,„)) k=0
fc = 0
i^ni)-i^^(0)
^
^
i^^(l)-iAO)-i
a.s.
as n -• 00. Thus the lack of bounded variation manifests itself here in the appearance of the term — ^. We will see later on in this chapter that — ^ results from the fact that dp(s) = ''(ds)^'' and therefore dp^s) = 2p(s) dp(s)-h (dp(s)f = 2p(s) dp{s) H- ds (cf. Ito's formula Theorem 4.4.1). It is important to notice that putting the increments of Brownian motion "in the future" as in (1.4) makes a difference (this is another manifestation of the absence of bounded variation). For instance, one can easily check that
(1.7)
lim I ^(h.umh^un) - P(Kn)) = i^'(i) - m o ) + i
n-^oo k = 0
Although the difference here seems to be small, it turns out that (1.6) is a far preferable way to define jo P{s) dp(s) than the one given by (1.7). Suffice it to say that the advantage with (1.6) is that the following relations hold:
j P{s)dp(s)=
0
and
|jJ/J(s)
4.2. Equivalence of Certain Martingales Let (£, ^, P) be a probabihty space and {<^,: r > 0} a nondecreasing family of sub (T-fields of ^. Let 5 > 0 be arbitrary and a: [s, oo) x E-*S^ and b: [s, oo) X E -^ R^ be bounded progressively measurable functions. For each (t, q) €[s, co) X E the components of a(r, q) will be denoted by {a'^(r, q)} and those of b(t, q) by {^(u q)].
86
4. The Stochastic Calculus of Diflfusion Theory
For any function/(x) in the space C^(R'^) of twice continuously differentiable functions we define L, / for r > s by
(LM)f)(x)-\ i ^'(t^^)A^(^) + ZHM) ^^ dXi dxj dxi 2.. We note that if (^(•, •) is any progressively measurable function from [s, 00) X E~^R^ then (Lj(q)f)(i(u q)) defines another progressively measurable function of t and ^ for r > s and q e E. Usually we will suppress the variable q and write (^(r), a(t), b{t\ L, for our objects. We shall suppose that we are given a function i(t, q) mapping [s, oo) x £ into R** which is progressively measurable, right continuous in r, and almost surely continuous in t. The main result of this section is the following theorem which proves that various types of relations between L, and ^(t) are equivalent. 4.2.1 Theorem. For any J( •), a( •),/?(•) satisfying the conditions described above, the following are equivalent: (i) f(i(t)) - J^ (L„/)({(w)) du is a martingale relative to {E, J^,, P)for t > sjor allfeC^(R'). (ii) f(t, i(t)) - jl ((d/du) + L„)/(w, i(u)) du is a martingale relative to (£, J*",, P) for t > sJor allfe C^ ^([0, oo) x R'^). (Hi) f(t, i(t))exp[~][((d/du)-\-L^)f/f)(u,i(u)du] is a martingale relative to (£, J^f, P) for t>s and for all f e C^ ^([0, oo) x i^**) which are uniformly positive. (iv) IfOeR'^ andge Cl ^([0, oo) x R'') then Xe,,{t) = exp U , ^W -^(s)-l j\e
+ Vg,a(u)(e +
-f^\~^ +
(v)
Hu)du) + g(t,Ut)) Wg)y(u,a^))du
L^yu,au))du
is a martingale relative to (£, ^ , , P)for times t > s. IfOeRUhen Xe(t) = exp <^, f(t) - ^(s) - j b(u) du}--\
(vi)
<0, a(u)ey du
is a martingale relative to (E, ^ , , P)for t > s. IfOeR^^then Xie(t) = exp K^, m
- ^(s) - \'b(u) duy + i j\o,
is a martingale relative to (£, J*^,, P)for t > s.
a(u)ey du
4.2.
Equivalence of Certain Martingales
87
Moreover if any of the above equivalent relations holds, then for each t > s (2.1)
P(^sup | { ( „ ) - ^ ( . ) - / " M . ) . . | ^ A )
^(2.)exp(-^^)
where A = sup,>s^^g£ sup|^l = i (6, a(t, q)6y. In particular, for any r > 0, exp rsup
(2.2)
\i(u)-i(s)
and the constant C depends only on t — s, r, A and B where B=sup \b(t,q)\. t>s qe E
Proof. Assume (i). Let/e C^{[0, oo) x R''). Then for s < ti < tj and /I e ^„ E[f{h,at,))-f{t„at,));A] = EUih, i{t,)) -fit,, ^(f,)) +/(t., ^(t^)) -fit,, ('{l^)iu,at^))du;A
I
iit,)); A]
(L,f)(t,,i(v))dv;A
= E [(l)iu,Ciu))du;A] +E
[C [(!)(«'«'^))-(IK ^^")) du: A
+ E \'\Kf)iv,iiv))dv;A •"(1
f\L^f)it„av))-iKf)iv,m)dv;A = £|n£+L„/j(u,^(u))
88
4. The Stochastic Calculus of Diffusion Theory
region, namely: t^ < w < D < ^2. It is obvious that the vahdity of the last equality for all fe CQ([0, CO) X R'^) implies the validity of the same equality for all feCt ^([0, 00) X R"^). Hence (i) implies (ii). Next assume (ii) and let/e Cl' ^([0, oo) x R**) be uniformly positive. Take
a(f) =/(t, i{t)) - j j l + L„/j(«, iiu)) du and ri(t) = exp
-i;(»ti^)(..*))*
Then a(t)n{t) - j\(s) dr,(s) =/(f, at)) exp f- | J ^ ^ ^ ) ± ^ J ( u , ^u)) dui Applying Theorem 1.2.8, we see that (ii) implies (Hi). Assume (Hi) and assume for the moment that {(s) = 0. Obviously (iv) would follow if we were allowed to take f(t, x) = exp[<0, x> -I- g(t, x)]. We cannot do this because/is unbounded and not uniformly positive. To circumvent this problem, we choose for each M > 1 a uniformly positive function f^ in Cl'^([0, co) x R*^) such that /Af(t, jc) = exp[<^, x> + g(t, x)] for X < M and define XM = (inf [t > s: sup \^(u)\ > M]) A (M vs). s
Then by (Hi) and Corollary 1.2.7, (Xe,g(r A TM), ^r» ^) is a martingale for r > 5. Clearly Xe^ g(t A TM) -• Xe^ g(t) a.e. F as M / " oo for each t > s. Thus we can establish (iv) if we can show that {XQ^ ^(f A TM): M > 1} is uniformly integrable for each t > s. But Xlg(tA.TM) = ^2e.2s(^ATM)exp <
| ' " " < ^ + V^, a(M)(^ + Vg)}(u, ^u)) du
CX2e,2g(t^'^M)
and £[^20,2g(t A T^)] = £[^20,2^(5)] = exp[2^(5, 0)]. This completes the proof when i(s) = 0. To remove this restriction, we define i'(t) = i(t) — ^(s). From (Hi) and Exercise 1.5.7 it follows that for uniformly positive functions / in
Cl\[0,^)xR<^)
/(U'W)exp[-((i^^/^^
))du
4.2. Equivalence of Certain Martingales
89
is an (£, ^^, F) martingale for t > s. From what we proved earlier, it follows that for all e eR" and g e C^ ^([0, oo) x R'^), X;,^(r) is an (£, <^,, P) martingale for t > 5, where XQ^ g(t) is the expression defined in the same way as X^ g(t) with ^'{t) replacing i(t). If one defines g^'(t, x) = g(t, x' + x) for x, x' e R*^ we have for s
m'e,,^t,)\3^t^
= X',,,^{t,)
a.e.
Again using Exercise 1.5.7
£W,...,('2)|^,J = ^^.,Jt.)
a.e.
Clearly X'Q^ g^^^^(t) — XQ^ ^(t) for all t > s, and we have therefore proved {iv). Obviously (v) is just the special case of (iv) when ^ — 0. To see that (v) implies (vi) it is enough to show that (v) implies (2.1) and (2.2). Indeed, if the estimates (2.1) and (2.2) hold, then both sides of the equality E[X,(t2);A] = E[Xe(t,);A] for s
(b(u)du}>A •'s
/
.exp(-pA + ^^4rA)j. Taking p = X/A(t — s) we get
Replacing ^ by — ^ we obtain a similar bound for the infimum. Combining the two, we have P ( sup \i(u) - ^(s) - rb(v)dv\ >yi] <2d,^-X'/2Adit-s) \s
Js
I
/
90
4. The Stochastic Calculus of Diffusion Theory
Since
p( sup b ) - {(s) \s<M
(b(u)ds\>A *'s
I
/
sup \ij{t) - ^s) - (b'{u) du\ > A/V5). \s
*^s
1
/
This proves (2.1), and (2.2) is readily obtained from (2.1) by estimating the integral in terms of the tail probabihties. It remains to show that (vi) implies (i). First we observe that, by elementary Fourier analysis, it suffices to prove (i) when f(x) is of the form /(x) = exp[K^, x}] for some 0 e R'^. Set a(r) = Xie(t) and r]{t) = exp
e.iis) + / b{u)du\-^
f (e,a{u)e)du
By Theorem 1.2.8 (x(t)rj(t) - I oc(u) driiu) is an (£, J*^,, P) martingale for r > s. It is easy to check that the last expression reduces to
mt))-j\L.f){i{f))du with/(x) = exp[i<^, x>]. This proves (i).
n
4.2.2 Corollary, Let ^(t) satisfy any one of the equivalent conditions in Theorem 4.2.1 and assume that £:^[e^l^<'^l] < oo, for all X>0. Suppose that fe C^'^([s, T) x R"^) n C([s, T] X R%for some T > s, and that there are constants Cj and Cj such that f its first time-derivative, and its first t\/\;o spatial derivatives are bounded by
f(tAT, i(tAT)) - j " ^ ( | ^ + L„j/(u, i(u)) du is an (£, J*^,, P) martingale after time s. In particular, for all 0 e R^: (2.4)
e,i{t)-i{s)-(b{u)du
4.2. Equivalence of Certain Martingales
91
and
(2.5)
U i(t) - i(s) -i^h[u)du\ - \\e, a(u)ey du
are (£, J^,, P) martingales after time s. Proof. Extend/to R x R"^ so that/(r, ')=f(s, *) i( t < s and/(r, ')=f(T, •) if / and i/^ e Co'((— 1, l))so t > T. Use/to denote this extension. Choose functions <> that [ and il/ = I on[-j,j].
(/>((t2+ \x\y^^)dtdx=
1
Define )„ and il/„ by
and
Finally, define/„ by fn(t.x) = il/„(t,x)((P„*f)(t,x). Then/„ e CS'iR x R'^). Moreover/„(r, x)-*/(r, x) for all (r, x) e [s, T] x R'', and /„, dfjdt, dfjdxi, and d^fjdx^dx^ are all dominated by Cj^^^'^'. Finally, for s < ^1 < ^2 < T' and A e ^^^\ £ U ( f 2 , c ( r 2 ) ) - / „ ( r i , ^ ( r O M ] = ^ ' j^''(|^ + L„)/„(^^(u))tit.,/l] Using these facts in conjunction with estimate (2.2) and the hypothesis that £P[^A|4(.)ij ^ 00 foj. all A > 0, we see that
/(r A T, ^ t A T ) ) - f " ' | £ + L„j/(u, ^i.))./!. is indeed a martingale. The rest of the proof is easy and is left to the reader. D
92
4. The Stochastic Calculus of Diffusion Theory
4.3. Ito Processes and Stochastic Integration Let (E, J^, P) be a probability space, {^t- ^ > 0} a non-decreasing family of sub
a: [s, oo) x E-^Sj,
b\ [s, oo) x £-•/?''
progressively measurable functions with the properties assumed in Theorem 4.2.1. If {, a, and h are related by one of the equivalent conditions given in Theorem 4.2.1, we will say that ^ is an ltd process (on (£, J^,, P)) with covariance a and drift b after time s. This sentence will be abbreviated by the notation ^ ^ ^d(^, b) when there is no need to emphasize (£, J^,, P). The purpose of the present section is to develop the theory of stochastic integration with respect to an Ito process. Since it is clear that a Brownian motion is an Ito process, the theory which we develop must take into account the pitfalls which we pointed out in Section 4.1. The approach which we will adopt is basically due to K. Ito, and, by modern standards, it is very classical. We begin our discussion with two simple observations. In the first place, the starting time s of an Ito process plays no essential role. Indeed, if ^ ~ ^d{^, b) on (£, J^,, P), then it is clear that | - J^(a, b) on (£, J^,, F), where: |(r, •) = (^(r + 5, •),
a(t, •) = ci(t + s, •),
i(r, •) = fe(t + 5, •), and ^, = ^ t + s
As with the theory of martingales in Section 1.2, this observation enables us to restrict our attention to the case 5 = 0 although we will state our theorems for general s. Our second observation is that if (^ ~ -^d(a, b) and we define ^'(t) = ^{t) — ji b(u) du, then ^' ~ ^^(a, 0). Thus, if we can assign a meaning to di(t) integrals when ^ ^ ^d(^, 0), then we can assign one to d^(t) for ^ ~ ^5(a, fe), namely: ''di(t) = d^'(t) + b(t) dt'' For this reason, until further notice, we will take b = 0. Let (£, ^, P) and {J^,: r > 0} be given. A function 0: [0, oo) x E^R^ is said to be simple if 6 is bounded, progressively measurable, and there is an integer n > 1 such that e(t) = e{[nt]/n) for all t > 0. If (^ - J^(a, 0) on (£, J^,, P) and 6 is a simple function, we define jo <^(w), d^(u)} by
(3.1)
| W . , ( u ) > . I (.(;-), (^f-±l.,)-^(^-M))
where n > 1 is any integer for which 6(t) = 9([nt]/n\ t > 0. Notice that in keeping with the warning in Section 4.1 the increment appears in the future. (It is easy to
93
4.3. Ito Processes and Stochastic Integration
check that the definition does not depend on n so long as ^(0 = 0{[nt]/n), t > 0.) If 0 < ti < t2, we define
Ui
-^0
'0
4.3.1 Lemma. Ifi^ ^J(fl, 0) and 6 is a simple function, then Jo <^(w), d^(u)} is a right-continuous, almost surely continuous progressively measurable function. Furthermore
•'o (j^'<e(u), di(u)}J
- j\e(u),
a{u)e(u)} du,
and 1 r' Xe(-)(t) = exp j <0(M), d^u)} - - J <%), a(u)e(u)} du are all (£, ^t, P) martingales. In particular, if
f(t)=f'<%w^(")> •'o and a(u) = {e(u),a(u)e(u)), then f ~ J^?(a, 0). Finally, if 6^ and 62 are simple functions and Xi and X2 ^^^ ^^<^l numbers, then Aj ^1 -4-^2^2 ^^ ^ simple function and:
\\x,e,(u) 4- A2^2(w), d^(u)y = ^,f<e,(u), d^u)} + A2|'<^2("), d^u)}. Proof That
Ut)=(<e{u),dau)^ is right-continuous, ahnost surely continuous, and progressively measurable follows immediately ft-om the corresponding facts about 1^. We will now show that
94
4. The Stochastic Calculus of Diffusion Theory
f ~ ^?(a, 0). Once this has been done, we can use Corollary 4.2.2 to complete the proof of everything except the final assertion. Since it is clear that if A e R, then
•'o we will know, by (v) of Theorem 4.2.1, that J ~ ^?(5, 0) once we show that Xe^.)(t) = Qxp
m-\fa(u)du
is an (£, J^,, P) martingale. To this end, choose n so that 9(t) = 6([nt]/n\ t > 0, and let k/n
^{t2)-^(h)=(e{^),Ht2)-i{h))Since 9(k/n) is a bounded ^^^-measurable function, we can apply estimate (2.2), Exercise 1.5.7, and (v) of Theorem 4.2.1 to conclude that: exp <e(k/nii(t2)-ati)} U'\e(k/nl
a(u)0(k/n)} du
1
a.s.
and therefore that E[XeUh)\^n]
= Xe^.it,)
a.s.
It is now an easy matter to remove the restriction that k/n < t^
^ > 0,
when n = Ml ^2. Thus X^ 6^ H- A2 ^2 is simple and it is easy to check that
One just has to write out jQ{ei(u),di{u)),i = 1,2, in terms of the partition determined by n. D
95
4.3. Ito Processes and Stochastic Integration
The next lemma demonstrates that a large class of functions on [0, oo) x £ into R"^ can be approximated by simple ones. To be specific, let a: [s, oo) x £ -• 5^ be a bounded, progressively measurable function after time s and define @ 5(a) to be the set of progressively measurable 9: [s, oo) x E^R^ such that j id{u\ a(u)e(u)y du
< 00
for all T > s. We now prove: 4.3.2 Lemma. Given 9 e @ J (a), there exists a sequence of simple functions {9„:n> 1} such that for every T > 0. (3.2)
i\e{u) - dM a{u)(e(u) - 0„(u))> du=
lim E M->00
/O
0.
J
Moreover, if C=
sup f >0,
\9(t, q)\ < 00,
qeE
then the 9^s can be chosen so that sup
\9,(t,q)\
t>0,qeE
for all n> 1. Proof First assume that
C=
sup 19{t, ^) I < 00 t>0, « e £
and that 9{\ q) is continuous for all qeE. Then we can take 9„(t, q) = 9([nt]/n, q), and clearly {9„: n > 1} has the desired properties, including the fact that suPt>o,«€£ |^n(f> q)\
(t>(n(t - u))9(u, du, q)
n > 1 and
t > 0.
Then, 9„ is progressively measurable, 9„(\ q) is continuous, and sup,>o,qeE |^n(^ ^)| ^ <^- Morcovcr, by elementary properties of approximate identities, lim [
\9„(t,q)-9{t,q)\'dt^Q
96
4. The Stochastic Calculus of Diffusion Theory
for all T > 0. Combining this with the preceding, we see that the lemma has been completely proved in the case that supj>o,,e£ \0(t, q)\ < oo. In order to remove this restriction, it is enough to show that if ^ € (§)?(«), then there are bounded e„ 6 @J(a), « > 1, such that (3.2) holds. But this is easy. Simply take 0„(t.q) =
Xi-n,n,mt.q)\)e(t.qy
Then e„ e @ J(a), 9„ is bounded, e„(t, •) -• e(t, •) a.s. for each t > 0, and <^„(r, q\ a(t, q)e„(t, q)} < <^(r, q\ a(t, q)0(t, q)} for all (r, q). Thus, by the Lebesgue Dominated Convergence Theorem, (3.2) holds. D Aside from one technicality, we now have the basic machinery needed to complete the definition of the stochastic integral. Let 6 e @J(a) be given. By the preceding lemma, we can choose simple ^„'s for which (3.2) obtains. Define Ut)=
(<eMdi(u)y,
t>0
and n > 1.
By Lemma 4.3.1, for all m, n > 1:
L{t)-^M and - j <e„(u) - 0 » , a(u)(0„{u) - 0„(«))> du
{L(t)-Ht)r
''o
are martingales. Hence, by Doob's inequality (Theorem 1.2.3):
sup
\L(t)-L(tn
<4E j ie„(u) - OM^ a(u)(9„(u) - OM)} du
0
and so, because of (3.2), lim E m->oo n-+oo
sup luo-uor
= 0
LO
for all T > 0. In particular, there is a sub-sequence {J„, (•)} such that ^„, (% q) converges uniformly on finite intervals for q outside a P-null set N. It is at this point that we encounter the aforementioned technicality. Namely, it would seem reasonable to define
) <%), di(u)y = lim e„, (t) 0
n'-^oo
4.3. Ito Processes and Stochastic Integration
97
off the set N, and let it be defined measurably, but otherwise arbitrarily, on N. We could even guarantee that the resulting function be right-continuous, and it would certainly be almost surely continuous. However, it would not, in general, be progressively measurable relative to the original family {J^,: t > 0} because the set N forces one to " anticipate the future." The usual way in which this difficulty is avoided is to complete the a-algebras J^, as suggested in Exercise 1.5.8. However, this solution is not entirely suitable for us, since the completion is a function of the underlying measure P and in our applications the measure P changes, whereas the (T-algebras do not. Thus we will work a httle harder and obtain a definition of jo <^(M), d^(u)y which, besides being right-continuous and almost surely continuous, is progressively measurable relative to {^^i t > 0}. The essence of the procedure that we have in mind is contained in the following lemma. 4.3.3 Lemma. Let {r]„: n> 1} be a sequence of right-continuous, almost surely continuous, progressively measurable function on (£, #^,, P) into R^. If lim sup PI sup \rj„(t) - ri„(t)\>e = 0 «-*oo n>m
\0
for all s>0 and 7 > 0, then there is a right-continuous, almost surely continuous, progressively measurable rj on (E, ^ , , P) into R^ such that lim Pi sup \rj„(t) - ri(t)\
>eUo
n-»QO
I
\0
for alle>0 and T > 0. Proof Without loss of generality, we can and will assume that for each t > 0, P lim sup sup \rj„(u, q) - //^(w, ^) | = 0 I = 1. \m-»ao n>m 0
I
Since each r}„ is progressively measurable, it is easy to see that A defined by A = {{t, q): lim^_.„ sup„>^ supo<„<, \rj„(u, q) - ri„(u, q)\=0] is progressively measurable (cf. Exercise 1.5.11), Moreover, if A{t) = {q: {t,q) € A}, then PiA(t)) = 1; and if Jq = {t: {t,q) G A}, then Jq is an interval and fl(%q)= lim^„(-,^) M-+00
uniformly on compact subsets of Jq. It is clear that rj is progressively measurable on A. Hence, by Exercise 4.6.8 below, t] admits a right-continuous, almost surely continuous, progressively measurable extension to all of [0,oo) x E. Finally, one easily checks that limPJ, sup \n„(t) - rj(t)\ > 8 l = o
n-^oo
\0
\o
for all e > 0 and T>0.
D
98
4. The Stochastic Calculus of Diffusion Theory
In view of the preceding discussion and Lemma 4.3.3, we have now proved the next result. 43.4 Lemma. If 6e@ J (a), then there exists a sequence {6„:n> 1} of simple functions and a right-continuous, almost surely continuous, progressively measurable function J such that hm E
j ie(u)-0„(u),4umu)-0„iu))ydu
0
and lim£
sup 0
for all T > 0. Obviously, it is our intention to take the function f described in the preceding lemma to be our definition of Jo <^(M), di{u)y. However, before we can do that we must check that this definition does not depend in an important way on the choice of the simple approximants 4.3.5 Lemma. Let 6 e @d(a). There exists a right-continuous, almost surely continuous, progressively measurable function J such that sup
lit) - \\e'{u), dau)>\ < 4£ f <.e(u) - e'(u),«(«)(%) - e'(«))> du
0
for all simple functions 6' and all T > 0. In particular, there is, up to a set of P-measure 0, exactly one such function and it is the one given in Lemma 43.4. Proof Let {0„: n>\) and J be given as in Lemma 4.3.3. If S' is a simple function, then sup l(t) - \ ie'(u), di(u)} 0
<{E
sup Ut)-\<0Au),d4(u)y\ ^ I
0
sup
f'<0'(«) - 0M di(u)y\
o
sup Ut)-(<,0Mdau)>, p
+ 21E f i0'{u) - 0M a(u){e'{u) - 0„{u))} du 2 £ J <0'(«) - 0{ul a(M)(e'(«) - 0(u))y du
•
D
99
4.3. Ito Processes and Stochastic Integration
As we noted at the beginning of our discussion, anything that we can do when 5 = 0 is immediately extendable to general s. Thus, if s > 0, we say that 6 : [s,Go)xE ^ R^ is simple after time s if 0(-fs) is simple relative to {^t+s : ^ > 0} and we define J^(9(u),di{u)) accordingly for ^ e J^J(fl,0). It then follows that if i ^ -^i(«,0) and 9 e ©J(a), then there is a unique, up to a P-null set, rightcontinuous, almost surely continuous, progressively measurable function | after time s such that sup s
f ie(u) - e'(u\ a(u){e(u) - e'(u))y du for all simple functions 0' after time s and all 7 > 0. This function 1 is what we call the stochastic integral of 6 with respect to <^, and we denote it by j ^ <^(M), d^(u)y. The next theorem is now immediate. 4.3.6 Theorem. For 6 e ®d(a) and ^ - J^5(a, 0),
and (j'<0(u), dC{u)}J - ^id{u), a(u)(u)> du are (E, J<^,, P) martingales after time s. Moreover, ifO^ and 92 ^^^ in @d(a), then for Ai, ^2^ R
\\x,9,(u) + X292Hdau)y ^,\\9M
d^My + ^2f<02H
da^)}
a.s.,
and so
sup s
j'iOM dau)} - [iOM dau)}\ <4E \ (0Au) - 02H a{u)(0,{u) - e^iu))} du
100
4. The Stochastic Calculus of Diffusion Theory
In the case that 6 e (H)a(a) satisfies (3.3)
sup <0(r, q), a(u q)0{U q)} < oo, t>s,qeE
we can say more about J^ <^(M), di(u)y. 4.3.7 Theorem. Suppose ^ - J\{a, 0) and 6 e ®%a) satisfies (3.3). Let ?(r) = J^ <0(M), d^(u)y and a(t) = <0(r), a(t)e(t)y. Then J - Jf\(a, 0) on (£, J^„ P). Proq/! It is enough to carry out the proof when s = 0. We will do sofirstunder the assumption that 9 is uniformly bounded. In this case, we can choose simple ^„'s having the same bound as 6 such that (3.2) holds. Then Lit)=(<9„{u),da'')> •'o
tends in probability to|'(t) as n-> oo; and, by Lemma 4.3.1, X'm
= exp
xUt)-^\'a„(u)du
is an (£, J^,, P) martingale for all X e R, where a„(t) = <^„(r), a(t)e„(t)y' Clearly X^J^^t) -> X;,(t) in probability, where X,(r) = exp Xm-^(a(u)du Thus, by (t;) of Theorem 4.2.1, the proof will be complete in this case if we can show that X^''\t)\ n > 1 is uniformly integrable. But (X^l\t)f = X5'l(r)exp
du
<m(r)exp[A2rCj, where C„ = sup,>o.,e£ <^«(^ ql E[X^;l(t)] = 1, it follows that
a(t, q)e„(t, q)}.
sup E[{X^"^(t))^] < 00.
Since sup„ C„ < oo and
4.3. Ito Processes and Stochastic Integration
101
To remove the restriction that 6 be uniformly bounded, choose bounded ^„'s from (g)S(a) so that (3.2) holds and sup <0„(t, ql a(t, q)e„(u q)y < sup {6(1, q\ a(t, q)e(t, q)} t>0,qeE
t>0,qeE
for all n (cf. the proof of Lemma 4.3.2). One can then repeat the preceding argument to complete the proof. D There are a few matters that we still have to discuss before closing this section. In the first place, suppose that ^ ~ J^i^, b), where b^O. Let @d(a, b) be the class of progressively measurable 9: [s, oo) x E-^ R'^ such that 9 e @d{^) ^^^ J
\(9(u\b(u)y\du
< 00
for all T> s. Given 9 e @3(a, b), we can choose a version (unique up to a F-null set) of jj <^(w)» H^)y du which is right-continuous, almost surely continuous, progressively measurable and is equal to the ordinary Lebesgue integral js <^(w, q\ b{u^ q)y du if JJ I (fi(u, q\ b(u, q)} \ du < co for all r > s. We can therefore define \\9(ul d^u)} = \\9(u\ di'(u)} + \\9(u\ b(u)y du, ''s
''s
•'O
where i'(t) = ^(t) - fo b(u) du, The next matter concerns the problem of defining f \9(u), di(u)}
for s
i^ Jta, b) and 9 G ®5(a, b).
There are three possibihties, all of which give the same solution. First, one might note t h a t ^ ~ ^5(a, b) and 9 e @5(ti, b) means, in particular, that ^ ~ ^^d^(a, b) and 9 e (H)a^(a, b). This would give one definition of JfJ <0(M), di(u)}. A second possibihty is to define f\9(u)M(u))=
f\9(u)M(u))-
Jt\
Js
f\9{uUi(u)). Js
To see that these two coincide is quite elementary and is left to the reader. Finally, one might adopt f<XuutMO(u\dau)\ s
for some T > t2,3LS the definition of JJJ i9(u), di(u)}. In Exercise 4.6.9 below, the reader is asked to show that this definition is the same as the preceding one.
102
4. The Stochastic Calculus of Diffusion Theory
The final topic to be taken up is the stochastic integration of matrix valued integrands. Let ^ ~ J^5(«, b) and suppose that
(3.4) and
I kMOl
(3.5)
dt < 00
for all r > 5. Then (7*6 e (g)5(a, b) for all 6 e R\ and we define j ^ ^(M) ^{(M) SO that (3.6)
e,\c7(u)di(u))
=
\(c7*(u)e,d^{u)}
for all 0 e /?". It is easy to read off the properties of J^ (7(u) di(u) from those of js
qe E
and sup \((jb)(t, q)\ < 00. t>s qe E
If^(t) = Js o-(w)
Proo/ Let 9 € R"he arbitrary. From (3.6)
= j<(r*{u)e, di'(u)} + j'ia*{u)0, = j'(c*(u)d, d4'(u)} + fie,
b(u)) du
(cb){u)y du
4.3. Ito Processes and Stochastic Integration
103
where i'(t) = ^(t) - J^ b(u) du and ^' - J'J[a, 0]. It follows from Theorem 4.3.7 that (3.7)
<0, lit)} - J\[e(7a(7*e, Oab].
From Theorem 4.2.1 (using i;), the validity of (3.7) for all 9 E R" is the same as f - J'„[(Taa*, ab]. D We close this section with the "chain rule" for stochastic integrals. 4.3.9 Theorem. Let ^ '^ -^|(a,^) and a : [s,co) x E -^ R^ ^ R'^ be a progressively measurable function with sup Trace ((Ta(T*){t, q) < oo t>s qeE
and sup \((Tb)(t, q)\ < 00. t>s qeE
Also, let p : [5, oo) X £ - • R*" ® J?" fee a progressively measurable function satisfying sup Trace (paacr*p*)(t, q) < 00 t>s qeE
and sup \(pab)(t,q)\ < 00. t>s qeE
Then if we take r,{t) =
(,T(u)dau)
and consider the stochastic integral with respect to tj, we have jy(u)dr,{u)=jy{uMu)d^{u). Proof. Consider the 2m dimensional process a(t) defined for f > s by (a,(t), a2(0) where ct,(t) =
j'p(u)a(u)di(u)
104
4. The Stochastic Calculus of Diffusion Theory
and ociit) = \ p(u) drj(u). We define y(t) for t > s as ad -h n dimensional process with components yi(t) and 72(0 where y,(t)^at)-i(s) and 72(0 = '/(O-
Clearly y(t:
1.1M)*'
and
'••>=i;rr'
;.,)^'H
We can now take a vector 6 e R"* and write <^, aj — a2> = <^, a) where 6 = (6, —9)e R^"". An elementary computation using Theorem 4.3.8 yields <^,ai(0-MO)-^1(0,0). Hence (Xi(t) - 0^2(1) = ai(0) - a2(0) = 0. This proves the theorem. D
4.4. Ito's Formula In Section 4.1 we saw that i( p(t) is a 1-dimensional Brownian motion, then
''0
The same argument as we used there yields the more general fact that
j'ii{s)dp{s)=mt)-mo)-{
4.4. Ito's Formula
105
As we indicated at the time, the appearance of the term r/2 on the right hand side results from the fact that "^j8(r)" is not a true differential, in the sense that (dp{t))^ i= 0. One can now ask how this fact manifests itself when one takes the " differential" of a more general function of a Brownian path. That is, suppose/is smooth. What is "i^/()5(r))"? When/(x) = x^, we have just seen that
df(m)=r{m)dm+¥"(m)dt. In other words, it is necessary to go out two terms in the Taylor's expansion of/in order to arrive at a true infinitesimal. This section is devoted to showing that the preceding is a general fact. A second way in which to view what we are about to do is " the identification of the martingale "
f(m)-f-2mp(s))ds as the stochastic integral
These preliminary remarks should become clearer as we proceed. 4.4.1 Theorem (Ito's Formula). Let ^ - ./^(a, b) on (£, ^,, P) with \i(s)\ bounded. Given a function fe C^-^([0, oo) x R"^) such thatf, df/dt, df/dxi, and d^f/dxidxj(l < i, j
(4.1)
f(t.m)-f(s,i(s)) = j\s/f{u,m)M'{u))
+ fil^-^
L^f(u,m)du,
t > s,
almost surely, where ^'(t) = ^(t) — \\ b(u) du. Proof Because of estimate (2.2), it is clear not only that V^/(-, (J(-)) e @d(a, b) but also that
[V./>.^(«))-v,/(«,?(«))pd« for all r > s as M -> 00, where
106
4. The Stochastic Calculus of Diffusion Theory
and
mj{t,m)),
B{t)=(Ht),[l+L,y{um)\
i
\la"{t)^{t,m) 1=1
if i=j=d + i if I
^-^/
if
l
Using (it;) of Theorem 4.2.1, one sees immediately that z ~ ^5+1(^, B). In particular, if 9(t) = {Vf(u (^(0), -1), then 0 is bounded; and so exp
A^ -' I {e(u\ dz'(u)} - - \ <0(u), A(u)e(u)y du
is a martingale for all k e R, where z'(t) = z{t) - \\ B(u) du. But {9(t), A(t)e(t)} = 0, and therefore j <^(w), ^Z'(M)> =0,
r > s,
almost surely. When this is written out, it is easily seen to be just equation (4.1). D Ito's formula in the form given by (4.1) above " identifies the martingale " in the definition of an Ito process. However, from a differential calculus point of view, (4.1) is not the most intuitive expression of Ito's formula. Indeed, if one wishes to emphasize the differential aspect of the formula, one should write: (4.2)
fit,i{t))-fis,m) = f'
where L^ = i Jj.J=i "'^(O S^/^^i ^^j 's the second order part of L,. Of course, (4.1) is equivalent to (4.2). The reason that (4.2) is more pleasing from a differential
4.5. Ito Processes as Stochastic Integrals
107
point of view is that it lends itself to the intuitively appealing differential formula:
(4.2') df{t, m=f((, m) dt+
dUt) dij(t) = ^'^(0 dt.
Equation (4.2') is, perhaps, the best and most concise statement of the "second order nature" of Ito processes. The reader may wish to compare this purely probabilistic statement with the analytic one given in Exercise 3.3.1. One more general comment about Ito's formula is in order. Equation (4.2') emphasizes the local nature of this formula and makes it seem unnatural that we have had to impose a global growth condition on/(cf. Theorem 4.4.1). The fact is that the growth condition is, in some sense, totally unnecessary from a path-bypath point of view; it is only there for reasons of integrability. The interested reader should consult the book of H.P. McKean Jr. [1969] or Exercise 4.6.10 below for a more expanded treatment of this point.
4.5. Ito Processes as Stochastic Integrals Let (P(t), ^t, P) be a ^/-dimensional s-Brownian motion and suppose that a: [s, co) X E^ R"^ (S)R'^ and b: [s, oo) x E-^ R'^ are bounded progressively measurable functions. An immediate consequence of Theorem 4.3.8 is the fact that any process i(t) satisfying (5.1)
i(t) - i(s) = I (7(u) dp(u) + j b(u) du,
t > 5,
is an Ito process having covariance a(') = (7a*(•) and drift b(•). In this section it is our purpose to prove the converse theorem. That is, suppose that ^ ~ ^d(^, b) on (£, J^,, P). What we want to show is that there exists an s-Brownian motion p(') for which (5.1) holds. Obviously, this will not be possible in general if we insist that P{-) live on {E,^t,P)' For instance, suppose that E consists of exactly one point q, ^ = J^^ = {(f), {q}}, and P = ^^. If {(r, q) = sin t, then ^ ~ J^?(0, b), where b(t, q) = cos t. On the other hand, there is no way in which one could put a 1-dimensional Brownian motion on (£, .i^,, P). Although this example may appear to be extreme, it demonstrates the necessity of distinguishing the situation in which a( •) may degenerate. For this reason, we are going to prove two versions of the converse theorem: one when a( •) is non-degenerate, and the other when it may degenerate.
108
4. The Stochastic Calculus of Diffusion Theory
4.5.1 Theorem. Suppose that a: [s, oo) x E-^S^ and b: [s, oo) x E -^ R'^ are bounded progressively measurable functions after time s on (£, ^) relative to { ^ , : r > 0 } . (Remember that S^ is the set of strictly positive definite d X d-symmetric matrices.) Let a: [s, oo) x E-^ R'^ x R'^ be a progressively measurable function for which a( •) = (T(7*( •). If^: [5, 00) x E-* R^isa progressively measurable function and P is a probability measure on (£, ^) such that ^ ~ ^%a, b) on (£, ^ , , P), then there exists a d-dimensional s-Brownian motion P(')on (£, ^^, P) for which (5.1) holds. Proof Note that o-(r, q) is non-singular for all (r, q) e [s, co) x E, and therefore (
r>5,
where ^'(t) = ^(t) - J^ b(u) du, then p - J^5(/, 0). Thus, if p(t) = 0 for r < s, then p(') is an 5-Brownian motion on (£, J^,, P) after time 5. Finally, by Theorem 4.3.9, fcT(u) dp(u) = fa(u)a-'(u)
= i(t)-i(s)-
and this completes the proof.
di'{u) =
(b(u)du
fdi'{u)
a.s.,
D
We now turn to the case in which a( •) is allowed to degenerate. The idea is Doob's (cf. [1952]). Intuitively speaking, what is going on here is that we first have to ramify the original sample space and thereby create a space on which we can fit a Brownian motion which is independent of the original ltd process. We then hold this new independent Brownian in reserve and only call on it to fill in the gaps created by lapses in " randomness " of the original Ito process (lack of randomness corresponds to degeneracy in a( •)). In this way we get a full Brownian motion for which (5.1) holds, although we will, in general, have had to introduce an external source of randomness in order to do so. 4.5.2 Theorem. Suppose ^ ~ Jd(^, b) on (£, J^,, P), and suppose that a: [s, 00) x £ -> P*^ ® P*^ is a progressively measurable function such that a — ao*. Then there is a probability space (£, ?, P), a non-decreasing family of sub o-algebras {J^,: t > 0}, and progressively measurable functions:
4.5. Ito Processes as Stochastic Integrals
109
[5, 00) X £ - > S ' '
[5, oo)x E-^R'^i^R'^ [s, co)x E^R"^
I
[s, 00) X E^R*^
and P: [s, 00) X E^R'' such that a, a, B, and J are jointly distributed under P in the same way as a, a, b, and ^ are under P, ^(•) is a d-dimensional s-Brownian motion on (£, ^t, P), and Ut) - e(s) = j'5(u) d^u) + J'B{U) du,
t > s,
P-almost surely. Proof. Let Q = Q^ ^s in Chapter 1, define x('), J^ and {J^^: r > 0} accordingly, and let i^^f^ be ^/-dimensional Wiener measure starting at (5, 0). Set £ = £ x Q, ^ = ^ x'j^,^^ = ^^ X ^ , , and P = P X ir%. Extend a, a, b, and ^ from E to E and x from Q to £ in the obvious way, and denote these extensions by 5, a, B, I, and X, respectively. It is easy to check that all these extensions are progressively measurable after time s with respect to {^,: t > 0} and that
-©--((o;)'©)
on
(£,#i,P)
(cf. Exercise 4.6.4 below). Let n(r, q) be the orthogonal projection onto the range of 5(r, q) and n(r, q) the orthogonal projection onto the range of cr*(r, q)cr(ty q). Note that n(r, ^) = lima(r, q)(el + a(r, q))-' e^.0
and n is therefore progressively measurable. Similarly, fl is progressively measurable. Define a by a(f, q) = lim {el + a(r, ^))- 'n(t, q) and set f = a*a. The following relations are easily deduced from elementary linear algebra: (5.3)
ra = n
4. The Stochastic Calculus of Diffusion Theory
110 and
at = n.
(5.4)
We now define
£(M)
to be the d x 2^-matrix
(T(M),
/ — n(w)) and
Ht) = ft{u)d^-{u), where
-(,)==w-i;('<"')*. Note that this integral is defined, since, in fact,
t(^ ^jt* = tar + / - n = (ta)(tay + / - n = / by (5.3). In particular,^ ~ ^d{L 0) and is therefore a ^-dimensional s-Brownian motion. Finally, fd(u) d^(u) = \'d(u)l(u)
^S'(M)
= fdt(u) d^'(u) + fd(I - n)(u) dx(u)
where l'(t) = f (r) - J^ B(u) du. But, by (5.4), at = n, and therefore
ht)-^'(s)-(at(u)d^'(u) j Trace[(/ - n)a(/ - n)(M)] du = 0. On the other hand, a(I — fl) = 0 since the null space of a coincides with that of aa*. Thus \(a(I - n)(u) dx(u) \ s
We can therefore conclude that
(a(u) dm=i'(t) - r(s)=m - us) - ["%) du, t > s, P-almost surely, and this completes the proof.
D
4.6. Exercises
111
4.6. Exercises 4.6.1. Let (^ - J^5(0, b) on (E, ^,, P). A special case of Theorem 4.5.2 is the fact that at) - ^s) = fbiu) du,
t > 5,
P-almost surely. However, one doesn't really need any of the machinery developed in this chapter to prove it. In fact, it is instructive to derive it directly from Theorem 1.2.8 and Leibnitz's rule. 4.6.2. Let ^ ~ Jl(a, b) on (£, J^,, P), where P is defined on (£, ^). For each t > 0, define ^^ to be the set oi A £ ^ such that there exists a B e ^t satisfying P(AAB) = 0. Next, take ^t^o = f]d>o ^t-^s- Show that (^ - J^5(a, ft) on (£, ^,+o» P) (cf. Exercise 1.5.8). 4.6.3. Show that the condition that an Ito process be almost surely continuous is superfluous in the following sense. Let (E, JF, P) be a probability space, {^t'- t >0} a non-decreasing family of sub (j-algebras of ^, and a: [s, oo) X E -^ Sd and b: [s, oo) x E-^R*^ bounded progressively measurable function after time s. Suppose i: [s, oo) x E -* R^ is a. right-continuous progressively measurable function after time s which satisfies any one of the conditions of Theorem 4.2.1. Note that C = {q: i(', q)is continuous on [5, 00)} is an element of J^, and show that P{C) = 1. Perhaps the simplest approach is to first note that almost sure continuity was never used in the proof of Theorem 4.2.1, and therefore that i must satisfy (ii) of that theorem. Conclude that E[ \ ^12) —
^(rOn < A(t2 - ri)V<^2-'^>,
s
4.6.4. Let (£, J^, P) be a probability space and {J^,: r > 0} a non-decreasing family of sub c-algebras of J^. Suppose ^ ^ J\(a, b) on (£, ^ , , P). Given a second probabiHty space (£', J^', P'), extend c^, a, and b in the natural way (i.e., ^(r, {q, ({)) = (^(r, q) etc.) to E" — E x E' and call these extensions J, a, and h. Define P" = P x P' on (£", ^"\ where ^" ^ ^ x ^'. Show that | - >5(a, h) on (£", J^, X ^\ P"). Use this result to show that if ^ - J\{d, b') on (£', ^\, P'\ where {J^J: r > 0} is a non-decreasing family of sub c-algebras of ^\ and if P'lf
r."\
( ^^^^ ^) \
^^''^^U'(t.q')j a"(Uq") = a(t,q)®a'{Uq')
for r > s and q" = (q, q')eE\ then i" ^-J'Aa", b") on (£", J^;', P"), where d" = d^ d' and J^;' = J^, x 3F[.
112
4. The Stochastic Calculus of Diffusion Theory
4.6.5. The following is an interesting, and sometimes useful, representation theorem for one-dimensional Ito processes. Let (E, ^ , P) and {J^,: r > 0} be as in the preceding and suppose that ^ ^ -^d(a, 0). Assume, for the moment, that there is an a > 0 such that a(t, q)>(x for all (r, ^) e [s, oo) x E. Then for each q e E, r -• 5 + js a(u, q) du is a strictly increasing function from [s, oo) onto [s, oo). Let i(%q) be the inverse function. Check that {x{u •): t > s} is a non-decreasing family of bounded stopping times and that T(*, q) is continuous on [s, oo) onto [s, oo) for all qe E. Apply Doob's stopping time theorem (Corollary 1.2.6) to show that if P(t, q) = (J(T(r, q\ q) — (^(s, q\ t > s, then p(*) is a 1-dimensional s-Brownian motion on (£, <^t(r)> P)- ^^ particular, one will have shown that at) = as) + p\(1>M a(u)dui
(6.1)
t>s
where the distribution of j5(*) under P is if^i]\). In other words, the paths of such a 1-dimension Ito process are the same as those of a 1-dimensional Brownian motion; the difference being entirely in the rate at which the process follows these paths. When one puts it this way, it is clear that there should be no essential reason for the insistence on the uniform positivity of a. We will now outline how to proceed when this condition is dropped. As in Theorem 4.5.2, the difficulty caused by allowing a to be zero is that there may not be sufficient " randomness " to support a full-blown Brownian motion. The development outlined below is based on the ideas in H. P. McKean's book [1969]. Suppose that ^ ~ ^\(a, 0) on (£, i^,, P). Because of Exercise 4.6.22, we can and will assume that J^, = J^^+o. Define C = s + jf a(u) du. The first step is to prove that limt^^ ^(t) exists almost surely on {C < oo}. To see this, let T(r) = sup{M > s: s 4- J; a(v) dv
n(t n\ = 1*^^^^^' ^)' ^) ~ ^('' ^^ ^^ ^(^) ^ ^
'^^'^^
\Uq)-as.q)
if
aq)
Note that r](', q) is right-continuous for all q e E. Using Corollary 1.2.6 and Exercise 1.5.10, prove that for all X e R: exp Ht)-j{{tAO-s)
-> ^x(t)^
P
is a martingale after time s. With Exercise 4.6.3, conclude from this that '/-^^l(X(..oo)(0,0)on(£,J^,(,),P).
4.6. Exercises
113
To complete the proof, let i^i]o t>e 1-dimensional Wiener measure on (Q, J/) starting at (s, 0). Let E = E x Q, ^ = ^ x J^, ^r =-^KO x ^ r . and P = P X irill .For q = (q, co) E E and t > s, define ^(^ q) = ri(t, q) + x(t, co) - x(tAi:(ql co). Using Exercise 4.6.4, show that j5 is a 1-dimensional s-Brownian motion on (JE, ^t, P). Finally, check that if ? and a are the natural extensions of i and a, respectively, to [s, oo) x £, then l(t)^^(s) + pl^s +
j\{u)du\
t > 5,
F-almost surely. This boils down to proving that ^(t) = ^(T(S + J^ «(«) ^M)) (a.s., P), which can be done most easily by calculating the second moment of the difference. 4.6.6. In Exercise 4.6.3, we pointed out that almost sure continuity is an unnecessary assumption in our definition of an Ito process. That is because we have adopted a very strong characterization of an Ito process. If, instead, we had chosen a characterization more in keeping with Levy's description of 1dimensional Brownian motion as the only ahnost surely continuous martingale p(t) such that P(0) = 0 and ^^{t) — t is a martingale, then the assumption of continuity would have been essential. Indeed, if rj(t) is a Poisson process with constant intensity 1, then tj(t) — t and (ri(t) — t)^ — t are martingales, but rj(t) — t is certainly not a Brownian motion. We will now outline how one can characterize Ito processes in a way which is consistent with Levy's idea. The procedure which we outline below is adopted from Kunita and Watenabe [1967]. Let (£, ^ , P) be a probability space, {J^,: r > 0} a nondecreasing family of sub (7-algebras of i^, and a: [s, oo) x E -^ S^ and /?: [5, oo) x £ -• /?** bounded progressively measurable functions. Let <^: [5, 00) x £-^P'^ be a right-continuous, almost surely continuous progressively measurable function. Then ^ ~ ^d(^, b) if and only if for each 6 e R^:
e,at)-as)-\Hu)du and
0, at) - as) - \HU) du\ - \\e, a(u)ey du are (£, ^t» P) martingales. The " only if " statement is immediate from Corollary 4.2.2. To prove " if ", it is certainly sufficient to treat the case when ^ = 0, since
114
4. The Stochastic Calculus of Diffusion Theory
Otherwise we can simply replace 4'(') by ^(•) - J^ ^(w) tiw. Giwen fe Co(R% s
sup
\i(u)-^(Ti"\)\v\t-zi"l,\>-\] At-
Since (J( •) is almost surely continuous, for almost all ^ e £ there is a k(q) such that 4"^(^) = h for k > k„(q). Thus
Assuming that b = 0, show that
£[/(^(4"'))-/(^W"^)M] 1
''
2i.pl
ay (^(4«>0)
[Sx,dxj
a^^u)du,A tfc-iC)
+ o(;^)£[tr-Ti"»j. From this it is easy to conclude that:
£[m^2))-mri)M] = £ J"L„/(^H)^M,/I 4.6.7. Let ^ ~ J^d(a, ^) relative to (£, J^^, P). Given a progressively measurable function V:[s,oo)xE^R which is bounded above, show that for any fECt'([s,oo)xR<^) exp (J' V{u)du\ fit,m)
- f e x p Q " ^((7)^(7') (^^ + L J + K/")(«, (^(M))^^
is a martingale after time s. The proof is a simple application of Theorem 1.2.8. As simple as this fact is to prove, it is nevertheless extremely powerful and leads immediately to a general form of the Feynman-Kac formula. 4.6.8. Let (£, (f^) be a measurable space and {J^^: r > 0} a non-decreasing family of sub (T-algebras of ^. Suppose that >1 ^ [s, 00) x £ is a progressively measurable set after time s (cf. Exercises 1.5.11-1.5.13) such that (ti, q) ^ A whenever ri > s and there is a ^2 > t^ such that (^2, q) e A, Define A(t) = {q: (f, q) e A\ J(q) — [t\ (r, q) G A), and T(^) = sup{r > s: (t,q)e A}. Show that {x
4.6. Exercises
115
/(r, q) = XQ for {t, q) E A A a n d / = / o n A. Show that/is progressively measurable on A Finally, define/on [5, oo) x £ by/(r, q) =/(? AT(^), ^) and check that/is progressively measurable after time s. Observe that if/(s q) is right-continuous on J(q) for all q, then/(% q) is right-continuous on [s, 00) for all q. 4.6.9. Let (^ ~ J^5(a, 0) on (£, J^,, ?) and suppose that 0 £ @d(a). Given stopping times Tj and T2 such that 5 < TI < T2 < T, for some T > s, show that Zi.,...)(-W-)6@5(«)andthat
= f^fc,..»<e(uMI(«)>
(a.s.,P),
where fJ{e(u)Ji(u)) = |(T/) when | ( ) = X(0(W),^C^(M)). This can be done without ever resorting to simple approximations. The idea is to use nothing but the basic properties of stochastic integrals plus Doob's stopping time theorem. By the same sort of reasoning, show that if T is a stopping time and 0: E^ R*^ is an J*^^-measurable function satisfying: r''
1 <0; a(u)e} ^M < 00
.s
J
for all r > 5, then X[; oo)('^)0 e ®d(«. 0) and \\XIU, oo)(t)^, d^u)} = <^, i(t) - ^(t At)).
4.6.10. Kunita and Watanabe [1967] gave an elegant and easy derivation of Ito's formula based on the results of Exercise 4.6.9. Define the stopping time T^"^ as in Exercise 4.6.6 with t^ = s and t2 = t. Then, almost surely,
fit, at)) -f{s, as)) = i (/(tr, ^K')) -mi,, mi o))One now expands / in a Taylor's expansion, up to first order in time and second order in space, around (xi"! 1, (^(T["1 J). By Exercise 4.6.9, one can see that the terms corresponding to the first order spatial derivatives tends, as n -• 00, to
j
116
4. The Stochastic Calculus of Diffusion Theory
in (3.2). This is entirely natural, since the approach is based on a differential technique. 4.6.11. Suppose that ^ - J^5(a, 0) on (£, ^,, P) and 0: [s, co)x E^R" gressively measurable. Does one really need the condition
I (e(ui a(u)e(u)y du
< 00,
is pro-
T> s,
in order to define jj (^(w)^ d^(^)}l The answer is no, as we will now see. In fact, suppose that instead (6.1)
•||^<%), a{u)e{u)y du < 00
for all
T>s\=l.
Define, for / > 1, Oi(t. q) = X[o,n\\ <^("»
Prove that f/+i(r) = J,(r) a.s. on {JJ <^(M), a(u)d(u)) du < /}. It then follows that there is a right-continuous, almost surely continuous progressively measurable J such that 1 = lim,_^ J, a.s. Of course I will not, in general, be a martingale, since l(t) need not even be integrable. Thus, the most suitable context in which to discuss this extension of Ito's integral is that of " local martingales ". To demonstrate that this extension is meaningful, the reader should use it to prove that Ito's formula holds for a l l / e C^'^([0, oo) x R**) without any growth conditions of/. Finally, it is important to notice that, in some sense, the condition (6.1) is the optimal one. In fact, one can show that the result proved in Exercise 4.6.4 can be extended to any stochastic integral i(') = is <^(")» ^^(w)>» where 9(') satisfies (6.1). That is, f(*) looks like a 1-dimensional s-Brownian motion run at the clock is ^0(u\ a(u)0(u)y du. Since a 1-dimensional Brownian motion has hmit superior equal + oo and limit inferior equal to — oo as time tends to oo, it is clear that our notion of stochastic integral breaks down completely if (6.1) doesn't hold. The interested reader is referred to the book of H. P. McKean [1969] for more details on this point, and also on the version of Ito's formula mentioned above. 4.6.12. Let (^ - J'a(a, 0) on (£, J^,, P) and assume that ^(s) = 0. Show that for each 2
sup \m'
iC
,p/2
Trace a(u) du
117
4.6. Exercises
for all T > s. Estimate (6.2) is one fourth of a family of beautiful inequalities discovered originally by D. Burkholder [1966] (See Ex A.3.3). The easy proof outlined below is based on ideas of A. Garsia. A more analytic approach can be found in the book of H. P. McKean [1969]. Here is Garsia's proof: sup \i(t)\' s
^{j^Jmirm
p rp(p-i)
[ Tracea(r)|{(Or'^t
p(p-i)
sup |(^(r)|''-2 J Trace a(t)dt s
2/p
21 p
^p(p-i)
sup \m\'
If Trace a(t) dt\
s
We have used here Doob's inequality, the fact that/((^(r)) — J^ ^w/(<^(")) du is a martingale with f(x) = | x |^ and finally Holder's inequality. 4.6.13. Let (J - J^i(a, 0) on (£, J^,, P) and assume that <^(0) = 0. We are going to outline a proof of the existence of a function /^(r, q)fort>0,xeR and q e E such that: (i) l^ is progressively measurable for each x e R, (ii) lx(% q) is right-continuous, non-decreasing, and /^^(O, q) = 0 for all x e R and qe E, (Hi) for almost all q e E, (x, t) -• /^(f, q) is jointly continuous, (iv) for almost all q e E, (6.3)
f Uf, q)dx = \ \'a(u)xr&u)) du for all r > 0 and r e ^ « .
In the case «(•) = h (6.3) shows that 2/.(t, q) is identifiable as the density, with respect to Lebesgue measure, of the occupation time functional, up until time r, of the path ^(% q). For this reason, l^ is called the local time at x. The existence of a local time was first observed by P. Levy [1948], but it was H. Trotter who first provided a rigorous proof. The tack which we adopt here is due to H. Tanaka. (See McKean [1975].) Although it is somewhat obscure in the present proof, what underlies the existence of a local time for a 1-dimensional Ito process is the fact that points have positive capacity in one dimension. The way in which this fact is used here is hidden in the boundedness at its pole of the fundamental solution for 3 d^/dx^. To be precise, the functional for which we are looking is given, at least formally, by
I j'aiu) 5{i{u]
x) du:
118
4. The Stochastic Calculus of Diffusion Theory
where S(') stands for the Dirac ^-function. If/(x) = x vO for x e R then/"(') = S( •), and then Ito's formula predicts that
J (a{u) d(m - x) du =/({(f) - x) -fi-x) - j'zi,. am) dm^ •'o
*'o
The terms on the right of this expression make sense and therefore indicate how we should go about realizing the left hand side. This observation is the basis of Tanaka's approach. We now outHne some of the details. For n> 1, let/„ be defined on K by: 1
X <
n
/„w=
1 1 - <x < n n
«(x+^)y4
X >
Of course /„ is not in C^(R), but an obvious mollification argument shows that Ito's formula can be extended to yield:
fMt) - x) -U-x) - (f'Mu) - x) dm n r ^ •'o almost surely. It is easy to check that for each x e R and T > 0: pi sup | ( / „ ( ^ ( r ) - x ) - / „ ( - x ) \o
(f'Mu)-x)dau)] •'o
/
- (fm - x) -f(-x) - \\^, ui(u)) ^^(«))| >fij tends to zero as n -• oo for each £ > 0. Here/(x) = x v 0. With this, together with the method used to construct stochastic integrals, one can conclude the existence, for each x e /?, of a right-continuous, non-decreasing, progressively measurable function T^ such that /^(O) = 0 and almost surely (6.4)
i(t}=mV X - 0VX + j'zu. am) dm,
«^ o.
119
4.6. Exercises
The next step is to show that we can modify T^ so that the resulting function satisfies (i), («*), and (iii) above. To this end, show that for each T > 0 and L > 0 there is a C = C(T, L) such that
sup \m-ut)[
0
for all X, y e R such that 0 < j; — x < L. This can be done most easily by observing that, by Exercise 4.6.12, sup i(xi..Am)dm)*] o
<36(4/3r£[(fa(«)zu.„(.^("))
L\ 0
/
One can now use Exercise 4.6.5 to estimate the right hand side in terms of a similar quantity involving a 1-dimensional Brownian motion, and the desired estimate then follows easily. We now define l':% q) = hn:.v2n{t. q) + 2"(x - [2"x]/2")\2n.,^,,2^(U q). It follows immediately from the preceding estimate that there exists C — C(T, L) such that sup E sup |/<;'(t)-/?'wr
for all 0 < j; — X < L. We can therefore apply Exercise 2.4.1 and the method used to construct stochastic integrals to conclude the existence of an /^^(r, q) satisfying (i), (H), and (iii) as well as the property that lx(') = L(') ahnost surely. Finally, note that i(t)s/x-Ovx-Ut) is an (£, ^ , , P) martingale for all x e R. Thus, if (p e Co(R) and F is given by F(x) = j(xvy)(p{y)dy, then Fm)-F(0)-\Ut)cp(x)dx is an (£, ^^, P) martingale. On the other hand, F" = (p, and so 1 c' F(^(r))-F(0)--|a(i/M<^(M))^t/ is a (£, J^,, P) martingale. It follows from Exercise 1.5.9 that (6.5)
~ j'a(u)cp(Hu)) du = j /.((V(x) dx,
t > 0,
120
4. The Stochastic Calculus of Diffusion Theory
almost surely. It is now easy to choose one P-null set N so that (6.5) holds off of N for all (p e Co'(R), and (iv) is immediate from this. 4.6.14. Let i^ be the one-dimensional Wiener measure. Wiener [1930] discovered a remarkable decomposition of the Hilbert space 1} (i^) into mutually orthogonal subspaces of "homogeneous chaos." This means that every O in 1} (i^) can be written as Z* ^^""K where O^"*^ comes from the subspace of "chaos of mth order." For each M > 1, let A„ be the set A„ = {se R";0 <Si <S2'-'<s„< oo}. Given / 6 L2(A„) (with respect to the Lebesgue measure) we define (6.6)
[ f(s)S"^x(s) = [ ^dx(s„) f"dx(s„.,)''•
Cm
dx(s,)
and show that (6.7)
(|^/(s)
if n^m,
lO f(s)g(s)ds
ifn
{i.
m.
To do this, we first define (6.8)
f
/ ( 5 ) S"^X(S) = (dx(s„)
rdx(s„.,)
'• • C m
dx(s,)
by repeated stochastic integration, moving from right to left. Here A„(t) = {5 € R" : 0 < si < S2... < s„ < r}. By using Ito's formula, we show, by induction, the analog of (6.7) with A„(0 replacing A„. We then use exercise 1.5.10 and let t -* 00. We can in this manner see that (6.6) makes sense and these multiple stochastic integrals satisfy (6.7). We can also check along the way that the left hand side of (6.8) can be defined as a progressively measurable, right-continuous, almost surely continuous process on Wiener space. We now define J^^^^ to be constants and Jt^"^ for n > 1 to be {/^^ /(s)J(">A:(S) : / G L2(A„)}. Using (6.7), one easily verifies that J^^^^ are for n > 0 mutually orthogonal subspaces of L^(ir). Wiener's theorem is that L^(ir) = ©^^ ^^"). To prove this fact, first define X^t) = exp
j'4>(s)dx(s)-U'<|,^s)ds
for 4> e L2P, 00). Show by Ito's formula that X^{t)=l
+ (4>(s)X4s)dx(s]. * 0
By induction, show that X^t) = 1 +
<^(Sj)(<«%(s) + i?„(t),
4.6. Exercises
121
where Rm{t)= I >(s„+i)^x(s^+i) '0
(t>(sjdx(sj... ''O
(l)(si)X^(s^)dx(si) ''o
Then estimate E[Rl^{t)] and show that E[Rl^(t)] -• 0 as m -> oo. Conclude from this that X(j)it) G 0 ^ J^^"^^ for all t. Finally, prove Wiener's theorem by observing that the only function in L^(i^) that is orthogonal to X^pit) for all (p and t is the zero function. Now if X(t) is a martingale on the Wiener space with suptE[X^(t)] < oo, then by Exercise 1.5.10, X(t) = E[X\Jii\ and use the representation X = Ig'O^")^ in terms of homogeneous chaos to conclude that X{i) is given by a stochastic integral X(t) = f e(s) dx{s\ JO
where E fe^(s)ds< CO, In particular, X{t) is almost surely continuous. This result has been generalized by Kunita and Watanabe [1967] and Jacod [1976]. See McKean [1973] for a more detailed discussion of Wiener's ideas.
Chapter 5
Stochastic Differential Equations
5.0. Introduction Starting with coefficients a(t, x) = ((a'^(t, ^)))i
^^^'" = ^
can be a source of a transition probability function on which to base a continuous Markov process. Although this is a perfectly legitimate way to pass from given coefficients to a diffusion process, it has the unfortunate feature that it involves an intermediate state (namely, the construction of the fundamental solution to (0.1)) in which the connection between the coefficients and the resulting process is somewhat obscured. Aside from the loss of intuition caused by this rather circuitous route, as probabihsts our principal objection to the method given in Chapter 3 is that everything it can say about the Markov process from a knowledge of its coefficients must be learned by studying the associated parabolic equation. The resuh is that probabihstic methods take a back seat to analytic ones, and the probabilist ends up doing little more than translating the hard work of analysts into his own jargon. In this chapter we present K. Ito's recipe for constructing a diffusion with given coefficients. Ito's method has the advantage for us, that it does not rely on the theory of partial differential equations; and, at the same time, provides probabilistic insight into the structure of diffusions with non-constant coefficients. The reader may be interested, if somewhat disappointed, to note that the class of coefficients to which Ito's method applies is (at least in the degenerate case) negligibly different from the class that We handled by the methods of Chapter 3. The origin of Ito's method is the following intuitive idea. Let coefficients a(t, x) and b(t, x) be given. The diffusion with these coefficients should be, essentially, a Brownian motion ^(t) which has been altered, at each instant, by changing its instantaneous covariance to a(t, ^(t)) and its mean to b(t, {(r)). That is, for small ^>0:
at+s)- m - ci% mm + s)- m)+Ht. m ^
123
5.0. Introduction
where P(t) is a Brownian motion with the property that {P(t + 6) — p(t): ^ > 0} is independent of ^(t). In other words, (^(•) ought to solve the stochastic integral equation: (0.2)
^(0 -as)=
f a^(u, m)dm
+ 2 h(u, ^{u))du.
There is a second, more formally acceptable but less intuitively appealing, way to arrive at (0.2). Namely, suppose that a and b are amenable to the techniques of Chapter 3, and let {P^, x'- (s» x) e [0, oo) x R"^} be the associated Markov family of measures on (Q, ^). Then, for/e Co(R'^) and s < fj < ^2 and x G R*^: E'^'if(x(t2))-f(x(t,))]
= E^
fL,f(x(u))du
where L^ is the operator having a(t, •) and b{t, •) as its coefficients. By the Markov property, one therefore has: E'''[f{x{h))
- / ( x ( f O ) | ^ , J = £''"'-'[/(x(r,)) =
^'^'l.^dD
-f(x{t,))]
f L,f(x(u))du
|X/Ww))^w|^ri But this means that f{x(t))-j'L,f{x{u))du is an (Q, ^t, P^, x) martingale after time 5. Because this is true for any/e C^(R% we conclude that x( •) - /J(a( •, x( •)), ^( % x( •))) on (Q, ^ , , P). If one sets (^( •) = x('), Equation (0.2) can therefore be seen as a consequence of Theorem 4.5.2. In this connection, note that, by Theorem 4.3.8, if ^ satisfies (0.2), then conversely Regardless of the manner in which one arrives at (0.2), the question remains as to how one can exploit this equation to construct a diffusion process. Following Ito, what we are going to do in this chapter is to treat (0.2) as a equation for the trajectories of the diffusion in much the same way as one would approach an ordinary differential equation. In fact, it has been recently discovered by H. Sussman [1977] that, in special cases, Ito's method can be interpreted as a natural extension of the theory of ordinary differential equations.
124
5. Stochastic Differential Equations
5.1. Existence and Uniqueness Let (£, J^, P) be a probability space, {^^1 ^ > 0} a n on-decreasing family of sub c-algebras of ^, and j5( •) is fi-dimensional Brownian motion on (£, ^^, P). Suppose that cr: [0, 00) x R^ ^ R"^ 0 R*^ are measurable functions which satisfy the conditions: (1.1)
sup \\(j{t, x)\\ < A,
sup \b(t, x) I < ^ t>0
r>0
and for all x, y e R*^ and t > 0: (1.2)
||(T(r, x) - (7(u y)\\ + \b(t, x) - b(u y)\ < A\x -
yl
where A
^(0 = rj + fa(u, ^(u)) d^(u) + f'%, ^(u)) du, t > s. s
s
Moreover, it will be shown that {(•) is uniquely determined by these properties up to a P-null set. 5.1.1 Theorem. Assume that G and b satisfy (1.1) and (1.2). Given s>0 and an ^s'^easurable tj: £ -> R** such that E[ | ^ |^] < oo, define ^„(t) = {„(f; s, r])for t > s by io(') = V and Ut; s,rj) = rj + j (T(U, ^„-,(U))
dp(u) H- | % , i„-,(u)) du.
Then there is a right-continuous, almost surely continuous progressively measurable function i(') = i('; s, rj) on[s, oo) x E into R"^ such that (1.4)
lim E «-*C0
sup
m-iM
0
s
for all T > s. In particular, (^(•) solves (1.3). Proof. We will first show that there is a progressively measurable function ?: [s, oo)x E-^R'^ such that (1.5)
lim E n-»oo
J \m-at)\'dt
= 0
125
5.1. Existence and Uniqueness
for all T > s.To this end, set
A„(t)=E[\a')-^n-dt)n
t>s.
Then: A„,,(t) < 2£ f {a(u, |„(u)) - c{u, i„.,{u))) dp(u) LI S
+
2E\\((b{u,Uu))-biu,i„.,(u)))du\
< 2A^ [ A„(M) du + 2A^t - s) f A„(w) du =
2A'(l-^(t-s))fA„(u)du,
and Ai(r)<2£ fa(u, rj) dp(u) + 2E I b(u, rj) du <2A^i-^(t-s))(t-s) Working by induction on /i > 1, we conclude that sup A„(r) < ^^ s
' n -t
^^^ ii
It follows immediately from this that for each T > s: lim sup£ m->oo
I
\Ut)-Ut)\'dt
n>m
By the Riesz-Fischer Theorem and Exercise 1.5.11, we conclude that there is a progressively measurable I: [s, oo) x E-^ R'^ for which (1.5) holds. We next set (1.6)
i(t) = ^ + \\(u, Uu)) dp(u) + | ' % , l{u)) du,
t > s.
Certainly (^(•) is right-continuous, almost surely continuous, and progressively measurable. If we can show that (1.4) holds for all T> s for this (^(•), then the
126
5. Stochastic Differential Equations
proof will be complete since we will then know that J( •) can be replaced by (J( •) on the right hand side of (1.6). To prove (1.4), note that
sup m-^m
s
<2E
sup s
+ 2 ( r - s)E \^ \b{u,l(u))-b(uAn-MW
du
< 2A\A + (T - s))E as « - • 00 by (1.5).
Q
5.1.2 Corollary. Let o, b, and r\ be as in the preceding. Then there is exactly one solution to (13); namely, the solution ^(*;s,rj) constructed in Theorem 5.1.1. Moreover, for T > s: (1.7)
sup
\i(t;s,rj)-i(t;s,rj')[
s
<3E[\r^-rj'
\']cxp[3A'(4 + (T - s))(T - s)]
for all square integrable ^^-measurable rj and rj' on E into R''. Proof Let rj and rj' be given and suppose (^(•) and (^'(•) are solutions to (1.3) for rj and ri\ respectively. Then:
sup im-mi '{A\2
<3£[i^-^'n
s
+ 3£ sup
j V(", i(u)) - c(u, ^'(«))) d^(u)
s
-}-3£ sup <3E[\r,-ri'\^]+
f{b(u, au)) - b(u, ^'(«))) du 12£ j {a(u, i(u)) - a{u, i'C))) dp(u)
+ 3(r - s)E j \b(u, i{u)) - b(u, i'(u))\Uu <3E[\„-r,'\'] + 3A^4 + {T- s)) f E[\m - iV)p] *•
5.1. Existence and Uniqueness
127
Putting A(r) = £[sup,<,<^ |c(r) - {'(r)!^], we have: A(t) dt.
A(T) <3E[\rj-r]'\^]^3A^4-^(T-s))\ An application of Gromwairs inequality now yields: A(T) <3E[\rj-rj'
\^]Qxp[3A''(4 -\-(T - s)(T - s)].
This proves, in particular, that {(•) = i'(') almost surely if rj = rj'. Obviously, since we now know that i(') and {'(•) must be {(•; 5, rj) and i('; s, rj') respectively, it also proves (1.7). D 5.1.3 Corollary. Let o and b he as before. For (s, x) e [0, 00) x R", let ^, (^(* v s ; s, ^/), where r\ = x. Then is,x(') ^^^ ^^ chosen to be (j(p{t)— P(s): t > s)-measurable. Moreover, if P'(') is a second Brownian motion on a second space (E\ ^\, F) and if^^^ ^( •) is defined accordingly for the primed system, then the distribution of i^, x(') ^nder P coincides with the distribution of ^'s^ ^( •) under P'. >S, JC\
Proof To prove the first part, simply note that if ^/ = x in Theorem 5.1.1, then, by induction, ^„( 'i s, rj) can be chosen to be (T(p(t) — P(s): t > 5)-measurable for each n>0. Thus the assertion follows from (1.4). To prove the second part, define in( S 5, ?7) with ?/ = X for the primed systems as in Theorem 5.1.1. By induction, the distribution of ^„(-; s, rj) under P and i'„('; s, rj) under P' coincides for all n>0. Therefore the same must be true of the distributions of 4 jc( *) and c^^ ^( •). D Corollaries 5.1.2 and 5.1.3 tell us that when a and b satisfy (1.1) and (1.2), we can unequivocally talk about the distribution of the solution to (1.3) when r^ is constant. Since 4 , x( *) is P-almost surely continuous, we think of its distribution as a probability measure P^ ^ on (Q, J^). We urge the reader to bear in mind that, because of Corollary 5.1.3, the measure P^ ^ is a function of a and b alone; and not of the underlying Brownian motion i5( •) or the space (£, ^,, P). 5.1.4 Theorem. Let {P^ ^: (s, x) e [0, 00) x R*^} be the family of measures on (Q, J^) associated with functions a and b satisfying (LI) and (L2). Then for bounded continuous E^'^'fO] is a continuous function. In particular, {Ps,x'(s, x) e [0, 00) X R^} is measurable in the sense that (s, x) -> £''"•''[0] is measurable for all bounded ^-measurable Q>: Q -> C. Proof Clearly it suffices to prove that lim d\0
sup 0<s-s'
sup E \x-x'\
sup 0
\is,x(t)-is',x'{t)\'
128
5. Stochastic Differential Equations
for each T > 0. Let 0 < 5' < s < T and x\ x e R'^ bQ given. For t > s, {5% At) i(t; s, rj') almost surely, where s
s
ri' = x' + j (T(M, {,<, A^)) dp(u) + J b(u, {,., A^)) du.
Thus , by (1.7), E sup \^s,x{t] s
-^•..•(oH
3£[ \t]'-x
|2]exp[3.l2(4 + (T - s)){T - s)\
and £[ I ^/' - x' |2] < 3 IX - X' |2 + ?>A\s - s') + 2>A\s -
s'f.
If 5' < r < s, then
^s',At) - 4,.(0 = X' - X + Ji CT(U, 4 , , , ( U ) )
^)5(W)
+ j i % . 4',x'(")) du,
and so sup s'
\UAt)-^s',Ai)\^
< 3 |x - x' p + 2>A^(s - s') + ^A\s - sy
Since (^^^ J(t) - 4 , x (0 = x - x' for 0 < r < s', this completes the proof.
D
We are now ready to prove the main result of this section. 5.1.5 Theorem. Let G and h satisfy (1.1) and (1.2) and define {Ps,x'- (s, x) e [0, 00) X R*^} accordingly. For 0 < s < t, x e R'^, and T e ^^d, define P(s,x;t.r)
=
P,^,(x(t)er).
Then (s, x) -• P(s, x; r, F) 15 measurable on [0, t) x R'^; and for s
e F | ^ J = P(t,, x(t,); t2. F)
(a.s., F , , J .
In particular, {Ps,x' (^» x) e [0, 00) x R"^} is a continuous Markov family and P(s, x; t, ') is its transition probability function. Finallyjorfe €^'^([0, 00) x R% 0 < s < t, and X e R'^: j f(t, y)P(s, x; f, dy) - / ( s , x) = j ' du \ i~
+ L„ j / ( u , y)P(s, x; u, dy),
5.1. Existence and Uniqueness
129
where
Proof. To prove the Markov property, l e t / e Cf,(R'^) and s
£M/Wf2)Mx(5i) e r „ . . . , x(5„) e r„}] Hence, if we can show that
E[m(t2; h, is..(h)))\^u] = j /(y)P(fi. is.Ahy, h,dy)
a.s.,
then the proof will be complete. To this end, note that for any z e R*^, itu^ih) is <j(p(t) — p(ti): t > ri)-measurable and is therefore independent of ^t^. Hence Elf(iUt2))\^tJ=EU(it,At2))] ==\f(y)P(h,z;t2,dy)
a.s.
Next, suppose that r]\ E-^ R^is J^,^-measurable and takes on only a finite number of values Zj, ..., z^. Then it is easy to see, from uniqueness and Exercise 4.6.8, that i(t2lt,,ri)==Y.Xir,=zj)in,zj(t2)
a.s.
1
Thus, for such an t], we have: E[mt2;
h, m^n]
= I fiy)nh.
ri; t2. dy)
a.s.
Finally, choose a sequence {^„}f of such ^'s so that £[|^/„ — 4, x(^i)|^]~^0 ^s w -• 00. Then ^{t2\ t^, rj„) -• i(t2; t^, is,x(h)) ^^ probability, and so
= lim I f{y)P{u, n„; h, dy) = j f(y)P(t„ ^,,,(tj); t„ dy)
130
5. Stochastic Differential Equations
almost surely. We have used here the fact that \f(y)P(ti,-;h,dy)
= E'''[f(4h))]
and is therefore a bounded continuous function. The final assertion is an immediate consequence of the fact that x( •) ^ /5(^(T*(% x(-)), b{', x(-))) on (Q, Jf,, P,, J. D 5.1.6 Remark. If c and b satisfy (1.1) and (1.2), then we have seen that the unique solution to (1.8)
i(t) = X + (<^{u, i(^)) dp(u) -f (b(u, (^(w)) du
is a measurable functional of p( •). This brings up a rather subtle point which we will have occasion to discuss in greater detail in Chapter 8. Suffice it to say now that when a and b fail to satisfy (1.1) and (1.2), there are situations in which (^(•) can satisfy (1.8) without it being true that (^(•) is a measurable functional of j5(-). See Exercise 5.4.2. 5.1.7 Remark. We have already observed that if cr and b satisfy (1.1) and (1.2), then the measures P^ ^ depend only on a and b. On the other hand, a has no intrinsic meaning, since it is a = aa* which is important and there may be a continuum of choices of a such that a = ca*. Thus we would like to know that P^ ^ really only depends on aa* and b. That this is indeed the case will be shown in Section 5.3. 5.1.8 Remark. The condition (1.1) can be considerably weakened. A good discussion of what one can say when (1.1) is abandoned can be found in the book of H. P. McKean [1969]. 5.1.9 Remark. Suppose that a and b satisfy (1.1) and (1.2) and that, in addition, they are independent of time. Let ^s,x(')^ (^^ x) E[0, OO) X R*^, be defined accordingly relative to )&(•) on (£, J^,, P). Given 5 > 0, set J^; = J^^+s, r > s, and P'(t) = p{t + s) - P(s). Then )?'(•) is aBrownian motion on (£, J^;, P). Moreover,
^(r) = X + fcj(^(u)) dp'(u) +\ b(i(u)) du,
t > s.
Hence ^(•) = io,x('l where (^o.x(-) is defined relative to )?'(•) on (£, J^;, P). From this we see that the distribution of ^s, x(' + ^) coincides with that of c^o, x(* )• In other words, if ^^i Q -• Q is the map given by x(r, 6^(0) = x(t + 5, co), t > 0, then Po,x = Ps,x ° ^7^' In particular, for 0 < s < r: P(s, x; r, F) = P,, ,(x(t) e F) = P , M ' M t - s) e F})) = Po, Mt - 5) 6 F) = P(0, x; r - 5, F).
5.2. On the Lipschitz Condition
131
Hence, if P(r, x, •) = P(0, x, r, •), then P(s, x; r, •) = P(r - 5, X, •),
0 < s < t.
The Markov process is therefore homogeneous in time.
5.2. On the Lipschitz Condition We saw in Section 5.1 that if we are given coefficients a and b such that b satisfies (1.2) and a can be written as a = aa* with a satisfying (1.2) then we can construct a diffusion corresponding to a and b. Unfortunately the assumptions on a are not directly transferable into assumptions on a. However we have the following theorems giving a wide class of coefficients a for which a exists with GG* = a and a satisfying (1.2). 5.2.1 Lemma. Let 0 <(x < Abe given, and suppose that ae S^ has all its eigenvalues in [a, A]. Then the non-negative definite symmetric square root a* of a is given by the absolutely convergent series:
A^ I c„(I - a/Af 0
where c„ is the n^^ coefficient in the power series expansion of(l--x)^ In particular, the map a-^ a^ is analytic on S^ into 5 / .
around x = 0.
Proof. Simply write a = A(I — (I — a/A)) and note that 0 < <^, ( / - aM)^> = \e\'-j{0,aO}
OeR'.
D
5.2.2 Theorem. Let a: [0, cc) x R*^ -^ S^ be a function such that {6, a(s, x)6y > ct 19 \^for all (s, x) 6 [0, oo) x jR**, 0 e R^ and some a > 0. If there isaC < oo such that \\a(s, x) - a(s, y)\\
y\,
s> 0
and
x, y e R'^,
then \\a^s,x)-a^(s,y)\\
s>0
and x, y e R*^,
Proof Clearly it is enough to show that if x -• a(x) is a C^-function on R^ into S^ such that <0, a(x)ey > a 101^ x e R^ and ^ e R^ then \\(aHx)y\\<^,\\(a(x))%
132
5. Stochastic Differential Equations
Moreover, we can assume that a( •) is diagonal at the point x in question. But (a(x))' = aHx){aHx))' + (ai(x))'a*(x); and therefore if a( •) is diagonal at x,
and this completes the proof.
D
We now want to look at the situation when a is allowed to degenerate. 5.2.3 Theorem. Let a: [0, oc) x R^ -^S^he a function having two continuous spatial derivatives. If max l
e,—A^.x)e dx;
<Xo\e\',
OeR'
for all {s, x) e [0, oo) X R'^, then \\a^(s, x) - a*(5, y)\\ < d(2Xo)^\x -y\,s>0
and x, y e K"^.
Proof It is sufficient for us to show that if x -• a(x) is a C^-function on R^ into Sj", then
sup \ie,(a(x)rey\ <^|^|^
OGR^
implies ||(ai(x))'||
The desired estimate now follows from Lemma 3.2.3. D
5.3. Equivalence of Different Choices of the Square Root Suppose p(') is a ^/-dimensional Brownian motion on some space (£, ^^, P). Given bounded measurable functions a: [0, oo) (x) R*^ -• K** x R'^ and b: [0, oo) x R'^ -• i?**, suppose that ^( •) is a right-continuous, almost surely continuous, pro-
5.3. Equivalence of Different Choices of the Square Root
133
gressively measurable solution to (3.1)
at) = ^ + f<^(u. a^)) dp(u) + \\(u,
^u)) du,
t >s.
Then, by Theorem 4.3.8, i(') is an Ito process after time s with covariance <7(% (^(•))
t > s.
Set U - ) = <^(- vs), and note that, by Theorem 5.3.1, ^, - /^Xis, ao)(-)«(% U*)). X[s, ao)(')H*y is(')))' Hence, by Theorem 4.5.2, there is another space (£', ^', F) and a non-decreasing family {^J: r > 0} of sub c-algebras of ^' on which there live a Brownian motion p'(') and a right-continuous, almost surely continuous progressively measurable function ^'(•) such that {'(•) is distributed under P' in the same way as is( •) is under P and
i'{t) = x+(xis.U"H'',i»dP'(u) + f X[s,oo){u)Hu,i'(u))du. But this means that P^ ^ is the distribution of (^'( •) under P', and therefore it is also the distribution of i(*) under P. This argument not only proves the assertion which we set out to check, it also proves the next theorem. 5.3.2 Theorem. Let a: [0, oo) x R'^ -* R'^ ® R'' and b: [0, oo) x R"^ -^ R'^ be bounded measurable functions satisfying (1.1) and (1.2). Define {Ps,x'- (s, x) e [0, cx)) x R'^} accordingly, and set a — aa*. Let (£, ^ , P) be a probability space, {^,: t >Qi} a
134
5. Stochastic Differential Equations
non-decreasing family of sub a-algebras of ^, and ^:[s,co) x E ^ R^ a rightcontinuous, almost surely continuous, progressively measurable function such that P{^{s) = x)=^landi - /3W% i(')l ^(% i('))) on (£, J^„ P). Then the distribution of i(' Vs) under P is P^^^•
5.4. Exercises 5.4.1. Let (T: [0, OO ) X R'^ -• i?'^ (X) R'^ and ^: [0, oo) x R'^ -• R*^ be bounded measurable functions satisfying (1.1) and (1.2). Given a space (£, J^, P), a non-decreasing family {^^i t > 0} of sub c-algebras of J*^, and a ^-dimensional Brownian motion j^(-) on (E, ,^,, P), define {(^,,^(-)- (s, x) e [0, oo) x R'^] as in Section 5.1. Using Exercise 4.6.12 show that for each 2 < p < oo and T > 0, there is a constant Ap J < CO such that sup |4,,,,(r)-i,,,,,(r)|^ < ^ p , r ( k i - ^ 2 | +
\x^-X2\'r'
0 < t < 7'
for all sj, S2 e [0, T] and Xj, X2 e P**. Next, apply Exercise 2.4.1, and the technique used in choosing a nice version of a local time (cf. Exercise 4.6.13) to find a new family {J5^('): (5, x) e [0, 00) x P*^} of right-continuous progressively measurable functions 5,, ^(-j such that P(is,x(') = 5s,x(0) = 1 ^^v all (5, x) e [0, cx)) x P'^and for P-almost all ^ e £ the map (5, x, t) ->• J^ ^{t, q) is continuous. 5.4.2. There is a distinction between the type of "path-wise" uniqueness proved in Corollary 5.1.2 and the "distribution" uniqueness proved in Theorem 5.3.2. In fact, this distinction will be the topic of Chapter 8. To prove that there really is a difference, let p(') be a 1-dimensional Brownian motion on some space (P,J^,,P). Define (T(X)
=
1
if
X> 0
1
if
X< 0
and Mt)=\ail}{s))dp(s),
t>0.
''0
Note that ^( *) is again a Brownian motion on (£, J*^,, P). Next observe that
and ^(t) = j'cT(-^(«))d^(M),
t>0.
5.4. Exercises
135
(The latter equation turns on the fact that
•(j^(0)(^M)^" = o) = for all t >0.) Thus we have here a a with the property that there is more than one solution to the corresponding Ito stochastic integral equation; and yet it is clear that all solutions have the same distribution, namely #^[)^/o • This example was discovered by H. Tanaka (cf. Yamada-Watanabe [1971]). Next observe that ^t) = J p{t) \ - 2/o(r), where /o( •) is the local time of j5( •) at 0 (cf. Exercise 4.6.13). Thus ^( •) is (T( | j?(r) |: t > 0)-measurable. On the other hand, p(') certainly is not cr( | P(t) \: t > 0)-measurable. Thus, in spite of the fact that
ii{t) = j'<^w{u))dm, t>o P(') is not (T(p(t): t > 0)-measurable. Here is the example promised in Remark 5.1.1.
Chapter 6
The Martingale Formulation
6.0. Introduction We have now seen two approaches to the construction of diffusion processes corresponding to given coefficients a and b. Let us quickly recapitulate the essential steps involved in these different methods and see if we cannot extract their common features. The approach presented in Chapter 3 involved solving certain differential equations in which a and b appear as coefficients. If enough of these equations can be solved, then we are able to interpret the " fundamental solution " as a transition probability function P(s, x; r, •). Moreover, P(s, x; r, •) automatically satisfies estimates which permit us to apply the results of Chapter 2 and thereby construct, for each (5,x) € [0,oo) x jR*^, a probability measure Ps,x on Q = C([0,oo),R'^) with the properties that: (0.1)
P,,^(x(r) = x, 0 < r < s ) = 1
and
(0.2)
Ps,Mt2) e r | ^ , j = p(t„ x(ri); t2, n
(a.s., p,,,)
for all 5 < f 1 < ^2 ^^^ r G ^j^d. The relationship between P^ ^ ^^^ the coefficients a and b is via P(s, x; f, •) and can be most concisely summarized by the equation: (0.3)
j P(5, x; r, dy)f(y) -f(x) 'R
= ( du j P(5, x; u, dy)L, f{y\ "s
'Rd
where
(0.4)
L. =
lia%.)^^^it>%.)±
6.0. Introduction
137
and Equation (0.3) holds for/e CS'(R'^)- Combining (0.2) and (0.4), we can eliminate P(s, x; r, •) and state the connection between F^ ^ and the coefficients a and b by the equation: £M/W^2))|^J-/(x) =£
fL,f(x(u))du\J^,,
or, more succinctly, (0.5)
f(x{t))-fL^f(x(u))du
is a P,, ^-martingale for a l l / e CJ'(i?'*). The second approach that we have discussed for constructing diffusions is Ito's method of stochastic integral equations. Here again there are intervening quantities between the coefficients a and b and the measures P^^ (described in the paragraph preceding Theorem 5.1.4). Namely, there is the underlying Brownian motion with respect to which the stochastic integral equation is defined and there is the choice of G satisfying a = aa*. That neither of these quantities is canonical can be seen most dramatically when one tries to carry out Ito's method on a differentiable manifold (cf. McKean [1969]). But, if one focuses on the direct relationship between P^^ and the coefficients a and b, matters simplify; and, in spite of the apparent differences between Ito's method and the construction via partial differential equations, one is led once again to Equations (0.1) and (0.5) as being the most concise description of the connection between P^ ^ and a and b (cf. Theorem 5.3.1). With the preceding discussion in mind, it seems only natural to ask if, in fact, (0.1) and (0.5) do not characterize what we should call "the diffusion with coefficients a and b starting from x at time s.'" Indeed, any measure with a legitimate claim to that title satisfies (0.1) and (0.5). Thus the real question is whether (0.1) and (0.5) are sufficient to uniquely determine a measure. That this may not be too much to hope for is already indicated in Theorem 4.1.1, from which it is easy to show that Wiener measure is characterized by (0.1) and (0.5) with a = I and ^ = 0. On the other hand, one must not allow oneself to be over-optimistic because it is easy to find coefficients a and b for which (0.1) and (0.5) do not uniquely determine a measure. For instance, let a = 0 and suppose that b is a continuous vector field which admits several integral curves through a given point. Then it is obvious that each integral curve determines a different measure satisfying (0.1) and (0.5). Thus we should guess that there will be cases in which (0.1) and (0.5) together determine a unique measure, but that it cannot be entirely trivial to recognize such cases. Obviously, an interesting class of questions come out of these and related considerations. Before investigating them, as we will be throughout the remainder of this book, we will formalize the statement of the problem around which these questions revolve.
138
6. The Martingale Formulation
Given bounded measurable functions a: [0, oo) x /?''-•Sj and b: [0, oo) x R*^ -• R'^, define L, by (0.4). Given (s, x) e [0, oo) x R'^, a solution to the martingale problem for L, {or a and b) starting from {s, x) is a probability measure P on (Q, J^) satisfying P(x(t) = x, 0
<s)=\
such that
f(x{t))-fL„f{x(u))du is a P-martingale after time s for a l l / e C^(R'^). The basic questions which arise in connection with the martingale problem for given coefficients are the following: (i) Does there exist at least one solution starting from a specified point (5, X)?
(a) Is there at most one solution starting from (s, x)? (Hi) What conclusions can be drawn if there is exactly one solution for each (5, x)?
Questions (/) and (ii) are, respectively, those of existence and uniqueness; while the answers to (Hi) presumably contain the justification for our interest in (i) and (ii). By analogy with Hadamard's codification of problems in the theory of partial differential equations, we will say that the martingale problem for a and b is well-posed if, for each (5, x), there is exactly one solution to that martingale problem starting from (s, x). To the student familiar with the modern theory of Markov processes via semigroups and resolvents the preceding formulation may seem somewhat odd. In particular when the coefficients do not depend expHcitly on t, the modern theories describe the diffusion process generated by L as a Markov family of measures {Pj,: X G R"^} whose transition probability function comes from a semigroup which is generated by an extension of L. On the other hand in our formulation each measure Px or, in the time dependent case, each measure Ps,x is described separately, and no mention is made about the connection between one P^ ^ and the others. There is cause for concern here, because unless one proves uniqueness, there will in general be no nice relation among the solutions starting from various points. (Although, as we will see in Chapter 12, under quite general conditions one can make a " Markov Selection " from among the multiplicity of solutions available for each starting point.) Thus, it should be clear that uniqueness is very vital for our formulation to yield a completely satisfactory theory. Once established, uniqueness will lead to many important consequences. For these reasons, after quickly estabhshing a reasonably general existence result, we are going to devote much of this chapter to the development of some techniques that will play an important role in estabhshing uniqueness results.
6.1. Existence
139
6.1. Existence We open this section with an easy construction which will serve us well in the sequel. 6.1.1 Lemma. Let s >0 be given and suppose that P is a probability measure on (Q, ^% where M' = a(x(r): t > s). Ifrj e C([0, 5], R'^) and P(x{s) = rj(s)) = 1, then there is a unique probability measure S^ 0 ^P on (Q, J^) such that S^ (x)5P(x(t) = rj(t% 0
^Ua,WW-|^(^j
if
r>5
is clearly a measurable map of X onto C([0, 00), R''), and therefore the restriction of P to X determines, via *F, a probability measure on (Q, J^). It is easy to check that this is the desired measure <5^ ^ ^ P . D 6.1.2 Theorem. Let x be a finite stopping time on Q. Suppose that co-* Q^^ is a mapping ofQ into probability measures on (Q, ^) such that (i) CO -• Qa)(A) is ^^-measurable for all A e J^, (a) Q^^(X(X((D), •) = X(T(CO), O))) = 1 for all a> e Q.
Given a probability measure P on (Q, Jf), there is a unique probability measure P(S)r(.)Q. on(Q., Jt) such that P®^.)Q. equals Pon (Q, Jt^)and{^^ 0t(co)6co} ^^^ r.c.p.d. of P ® t ( ) G . l-^f ^^ particular, suppose that x>s and that 0: [s, 00) X Q-)> C 1*5 a right-continuous, progressively measurable function after time s such that 6(t) is P<S)^^.)Q.-integrable for all t > s, (^(r AT), ^ , , P) is a martingale after time s, and (6(t) — 9{t A X((O)), J/^ , Q^) is a martingale after time s for each co e Q. Then (0(t), J/^, P (S)z{.)Q.) is a martingale after time s. Proof The second assertion is an immediate consequence of Theorem 1.2.10 once we have proved the first part. Moreover, the uniqueness assertion of the first part is obvious. Finally, to prove the existence of P. 0 T ( ) 6-»it is enough to check that (^-* ^oi ®r(a>)6a>(^) ^s ^.-measurablc for ad\ A e J^ and then set R(A) = E^[3.^,^.^Q.(A)l
AeM.
140
6. The Martingale Formulation
Once it is known that co -• ^^ <S)x{co) Qco(^) is e/#^-measurable, the proof that R has the desired properties of P. ® t ( ) 6 . is easy. But if ^ = {x(ti) e F j , ..., x{t„) E F J , and F j , ..., F„ e ^j^d, then where n> 1,0
nM^))QM)
n-l
+ I X[r.,rk.i)(i^N)Zri(^(^. to)) ••• Zr.(^(^fc, co)) X ea,Wt,+ i ) e F , + i, . . . , x ( g e F „ ) + X[r„, oo)(T(co))zri(^(^i, w)) • • XrSx(t„, co)), and this is clearly ^^-measurable.
D
We will soon use Theorem 6.1.2 to prove a quite general existence resuh. But before we do, we want to establish a very useful theorem which complements Theorem 6.1.2. 6.1.3 Theorem. Let a: [s, oo) x Q -• 5^ andb: [s, co) x Q-* R'^ be bounded progressively measurable functions. Suppose that P is a probability measure on (Q, J^) and that (J: [s, 00) X Q -• i?** is a progressively measurable, right continuous P-almost surely continuous function for which (^(•) ^ ^3(a(*), b(')) with respect to (Q, ^f, P). Given a stopping time x > s and a r.c.p.d. {Q^^} ofP \ M^, there exists a P-nullset Ne J/, such that i(') ^ J^j'^\a(•),/?(•)) with respect to (Q, .v#,, QJ) for each w^N. Proof Let @ ^ CQ{R^) be a countable, dense subset. For e a c h / e @ , we can apply Theorem 1.2.10 to find a P-null set Nf e M^ such that
Xy(t) =/(^(f, «)) - / ( a t ( c « ) , «)) - f' (L„/)(^(u, «)) du
is a martingale after T(CO) with respect to (Q, ^ , , Q^) for each co ^ Nf. Define N =[jfe(H)N. Then iV is a P-null set and NeJi,. Moreover, for a n y / G C^{R^) and CO ^ AT, we can find {/„}f ^ (g) such that /„ - • / in C?(P'^) and therefore XJ?jr)-• Xy(r) boundedly and point-wise. Hence {X%t\ Jt^, Q^ is a martingale after T(CO) for all / G Cg^R^) and co ^ N,By Theorem 4.2.1, this completes the proof. D 6.1.4 Lemma. Let c: [0, oo) -• 5^ and b: [0, QO)-^R^ tions and set
L. =
be bounded measurable func-
-,ifit)^^^'^^im^^
6.1. Existence
141
Then the martingale problem for L, is well-posed. Moreover, if {Ps,x- (5, ^) e [0, 00) X R'^] denotes the family of solutions determined by L,, then (5, x) -• P^^ ^(A) is measurable for all A e J^. Proof To prove the existence of a measurable family of solutions, we could invoke Theorems 5.1.4 and 5.3.2. Moreover, uniqueness could be proved by an application of Theorem 4.5.2 and Corollary 5.1.3. However, we prefer to take a different route. Define C(s, t) = J^ c(u) du and B(s, t) = f, b(u) duior 0<s
(a.s., P)
for all s < fi < r and 6 e R'^. Hence, if/e B(P^), then EV{x(t2))\J^n]
= j /(y)i^(ri, x(rO; ^2, ^y)
(a.s., P)
for s
= g{x)"'gm(x) J--J fliyi) rix-xr„ L(yn)P(s, x;t,,dy,)"
P{t„. 1, y„_ 1; r„, dy„).
142
6. The Martingale Formulation
This equation not only demonstrates that P is unique, it also guarantees measurability as a function of (s, x). Thus it only remains to prove existence. But this is easy, since P(s, x; r, •) certainly satisfies the conditions of Theorem 2.2.4; and if {Ps,x- (Sy x ) e [0, oo) X R"^} is the corresponding Markov family constructed in that theorem, then
exp
iUx(t2)-j
b(u)duW-\
<0,C(#> du
= (p(t,, x(t,); t2, e)Qxp[-Ke, B(s, t2)y + i<0, c(5, t2)oy] exp
iU x(t,)-
j ' ' h(u) du\ + ^ fie, c(u)e} du
(a.s.,F,.J for all s
Proof Define P<^> on (Q, Ji) by
P^yA) = I dMWx),
A e M,
where b^ is the measure on (Q, M) such that bj^x{f) = x, r > 0) = 1. For /c > 0, define a^{', (o) — a{', co^), b^{', w) — b{% co^), and r
—r
VS)kln ^ k/n, x(k/n, •)
By Theorem 1.3.5, there is a unique P on (Q, Jf) such that P equals P^^^ on Jf^/n for all /c > 0. G i v e n / e C^(K^), set
0(t)=f(x(t))-\
L^f(x(u))du
6.1. Existence
143
where
Certainly (e(t A 0), Jf^, P^^^) is a martingale. Assume that (0(t A k/n\ M^, ?^^^) is a martingale. Since
is a martingale for all co, it follows from Theorem 6.1.2 that (^(t A [(/C + 1)/«]), ./#,, p(fc+ i)j is a martingale. Thus, by induction, {d{t\ M^, P) is a martingale, and so we have completed the proof of existence. Clearly the uniqueness will follow if we can prove the last assertion. Given P with the prescribed properties and /c > 0, define \l — ^yy{k^\)ln
^
(A:+l)/n,jc(()t+l)/«,-) *
By Theorem 6.1.2, we see that x() ~ ^j^(a/t(),^fc(')) with respect to ( Q , ^ t , 0 . Thus by Theorem 6.1.3, if {Qoi\ is a r.c.p.d. of Q \ J^k/n^ then, for g-almost all co,x{') ~ J^j'^"(flfc(-,co),/?fc(',co)). Hence, by Theorem 6.1.3, for ^-almost all co, (o — ^w yS)k/n ^ kin, x(k/n, a)
Since P equals Q on J^k+ i/n» this finishes the proof
D
6.1.6 Theorem. Let a: [0, oo) x Q -> S^ and b: [0, cc) x Q ^ R"^ be bounded progressively measurable functions. If for each t > 0, a(t, •) and b(t, •) are continuous, then for each probability measures p. on R^ there is aP on (Q, J^) such that P(x(0) e r ) = p(r), r 6 J?«,, and x(-) -- ^d(«(•), H')) ^^th respect to (Q, ^ , , P). Proo/. For each n > 1, define a„: [0, oo) x Q -• S^ and /?„: [0, oo) x Q -• P*^ by a„(t, co) = a(r, cb) and
b„(t, oj) = b(t, cb)
where a> e Q is determined by x{u, cb) = x{u A [nt]/n, a>), u >0. Then a„ and b„ satisfy the conditions of Lemma 6.1.4, and so we can find a P„ on (Q, M) such that P„(x(0) G r ) =//(F), F e ^ « „ and x ( - ) - > S M - ) ' M O ) with respect to (Q, ^ , , P„). Moreover, by Theorem 1.4.6, {P„: n > 1} is relatively weakly compact on (Q, M). Let P be any limit point of {P„: n > 1} and let {P„.} be any subsequence such that P^ -• P. Clearly f / ( y V(dy) = lim E*-" [/(x(0))] = £^/(x(0))]
144
6. The Martingale Formulation
for all / e Ci,(R^), and so P satisfies the correct initial condition. Moreover, if
f(x(t))-(L^:n4u))du is a P„-martingale, where
Since for each t > 0, the functions
fm(x{u))
du
are equicontinuous at every point of Q, bounded independently of n, and tend to Jo L„/(x(w)) du as n-^ oo, we can apply Corollary 1.1.5 to prove that
j/(x(r))-j^L„/(x(t/))^i/j
CO
for all r > 0 and bounded continuous . In particular, if 0 < fj < tj and <[> is a bounded continuous ,y#,j-measurable function, then we have:
(/W'2))-jJi„/(x(«))d«j« = £•"
^f(x{t,))-j\^f(x{u))duy
since this equahty certainly holds when P and L„ are replaced by P„> and Ljf ^ respectively. It is now easy to see that this equality persists when <E) is simply bounded and ^^^-measurable, and so we have shown that x(') ^ J^(a('), /?(•)) with respect to (Q, ^ , , P). D 6.1.7 Theorem. Let a: [0, oo) x R'^ -^ S^ and b: [0, oo) x R"^ ^ R'^ be bounded measurable functions such that a(r, •) and b(t, •) are continuous for all t > 0. For each 5 > 0 and each probability measure fi on R'^, there is a probability measure P on (Q, ^) such that P(x(t) G r, 0 < r < s) = /x(r),
r
e^j,,,
6.2. Uniqueness: Markov Property
145
and x(') ~ -^d(«(*, ^(*))» H'^ ^(*))) ^^^^ respect to (Q, J^j, P). In particular, the martingale problem for a and b has at least one solution starting from each (s,x)e [0, oo) X R''. Proof. Clearly we need only prove the first assertion. To this end, let s and p, be given and define as and ^5 by as(t,x) = a(t + s,x) and bs{t,x) = b{t -\- s,x). By Theorem 6.1.6, we can find a 2 on (Q,J^) with the properties that Q(x(0) e r) = p(ri
FE^^,,
and x(-) ~ J^dM'^ ^(')\ M% ^(O)) with respect to (Q, ^ , , Q). Therefore, if we now define
6.2. Uniqueness: Markov Property We now have our basic existence resuh. Unfortunately, the corresponding uniqueness theorem is not so easy to prove; in fact, it is not even true. The purpose of the present section is to start laying the groundwork for a reduction procedure with which we will eventually be able to obtain a reasonably satisfactory uniqueness theorem. At the same time, the preparations that we make here will serve us well when it comes time to discuss the consequences of uniqueness. In order to appreciate what is going on in this section, it is helpful to think about how one might try proving uniqueness. Given two probability measures P and Q on (Q, J^\ one knows that they are equal if and only if P(x(t,) G Fi, ..., x(t„) 6 r„) = Q(x(t,) e Ti, ..., x(t„) e r„) for all « > 1, 0 < ^1 < • • < r„, and Fi, ..., r „ e ^^d. That is, if the finite dimensional marginal distributions determined by P and Q agree. If one knows a priori that P and Q are Markovian in the sense that for any 0 < r^ < r2 the conditional distribution of x(r2) given J^^^ is a function of tj, t2, and x(ti) alone and if this function is the same for P and Q, then P = Q if x(0) has the same distribution under P and Q. Of course this line of reasoning cannot be apphed without the a priori knowledge that P and Q are Markovian. Nonetheless, a variant of it is applicable to the study of martingale problems; and it is this variant which we now want to develop. 6.2.1 Theorem. Let a: [0, 00) x R'^ ^ S^ and b: [0, 00) x /?'' -• R'^ be bounded measurable functions and suppose that P on (Q, ^) has the properties that P(X(
146
6. The Martingale Formulation
N E M^ such that Sx(j^^a), ^^^ ®t(a)) Pa solves the martingale problem for a and b starting from (T(CO), X(T(CO), a>)) whenever a>^ N. (Here, and in the similar contexts in the future, Sx is the point-mass onQ concentrated on the path which stays at xfor all time). Proof In view of Theorem 6.1.3, the second part is an immediate consequence of the first part. To prove the first assertion, l e t / e CoiR"^) and set e(t)=f(x(t))-f(x{tAC7))-
(
Lj(x(u))du,
where L„ is defined in terms of a and b by (0.4). Certainly (9(t A a\ Jt^, d^ is a martingale, and, by assumption (9{t) — 6(t A (T), ^^, P) is a martingale. Hence, by Theorem 6.1.2, (6(t), J^t, d^ ®
f(x{t))-j'L,f{x{u))du is a 3^ (x)^P-martingale after time a; and so the proof is complete. D One immediate application of Theorem 6.2.1 is the following. 6.2.2 Theorem. Let a: [0, oo) x R"^ -* S^ and b: [0, oo) x R'^ ^ R'^ be bounded measurable functions for which the martingale problem is well-posed, and assume that the corresponding family of solutions {Ps,x' (•^» x) e [0, oo) x R*^} is measurable (i.e., (s, x) -• Ps^ ^(X) is measurable for all A e J^). Then {P^^ x'. (s^ x) e [0, oo) x R*^} is strong Markov in the sense that if x >s is a finite stopping time then {^co®t(a»^t(ca).x(r(a»,c.)} ^^ « r.C.p.d. of P ,^ ^\J^, .
Proof By Theorem 6.2.1 and uniqueness, we know that if {P^^ is a r.c.p.d. of Ps, XI-^t, then ^^(,(^), „) (x),(„) P„ = P,(^), ^(,(^), ^) for all co outside of P,, ^-null set N e Ji^. Thus P(0
— ^CO ®T(ct>)(<^x(T(c0), to)
®x{oi)Pco)
— 3(a ®x((a)Px((o), x(xi(o), to)
for all (o^ N. D It is worth remarking that the measurability condition on {Ps,x'- (s, x) e [0, oo) X R^} in Theorem 6.2.2 is not really necessary. Indeed, the above proof shows that there is a P,, ^-null set N e ^, such that P„ = 3^ (x),^^) P,^^^^ ^^,^^^^ ^^ off of N; and this is all that is important. Moreover, it is shown in Exercise 6.7.4 below that {Ps,x '- (s,x) € [0,oo) xi?''} is necessarily measurable if the martingale problem is well-posed. However, the proof of this fact requires the application of a rather hard theorem about Polish spaces and there is no circumstance in which we have been able to prove uniqueness without verifying measurability at the same time. We will therefore avoid using Exercise 6.7.4 whenever possible.
6.2. Uniqueness: Markov Property
147
We now turn to the result, alluded to earlier, which will help us to prove uniqueness criteria. Before stating this result, we recall the notion of a. determining class of functions. Namely, <S> ^ B(R^) is said to be determining if whenever two probability measure /Xj and fi2 on R'^ satisfy
\(P dfii = j (p dn2 for all (^ e O the measures fi^ and fij must be equal. 6.2.3 Theorem. Let a: [0, oo) x R'^ -^ S^ andfo:[0, oo) x R^ -^ R^ he bounded measurable functions. Suppose that for every (5, x) G [0, 00) x R^, t > 5, and (p from a given determining class ,
£^'[
E^vixm
whenever P^ and P^ both solve the martingale problem for a and b starting from (s,x). Then starting from each (s,x) there is at most one solution to the martingale problem for a and b. Proof. Let P^ and P^ both be solutions starting from (5, x). Certainly P^ and P^ agree on J^^'^ and therefore, since J/^ is independent of (j(x(t): t > s) under both P^ and P"^, it suffices to check that (2.1)
E'yMh))"fn(4tn))]
= E'VMh)) "fn(x{t„))]
for all n > 1, 5 < ti < • • < r„, a n d / i , ...,/„ e B(R^). Certainly (2.1) is true if n = 1. Next, assume it is true for n. Set ^ = cF{x(tj): 1 <j
(2.2)
F^ = jPU-)^L(^co')
for CO not in a P'-nuU set Ai e Jt. Next, choose a P'-null set N^e Ji^^ so that ;c( •) ~ ^^(«(., x{')),fe(•, x( •))) under P^ for all o^ Ni. Then there is a p"'-null set Bi e J such that PU^i) = 0 for co ^ B,. It follows, from (2.2), that if co ^ iV^ = Ai^Bi, then x(-) - J^^"(«(% x(-)), /?(•, x(-))) under Pj^. But, by induction hypothesis, P^ equals P^ on ^ , and therefore P^(iV2) = P^(N2) = 0 and p2(iVi) = i'M^i) = 0- Thus, if N = Ni u iV2, then F(iV) = 0 and, by Theorem
148
6. The Martingale Formulation
6.2.1, <5^(,„ ,^)(x),„Pj^ solves the martingale problem for a and b starting from (r„, x(r„, o))) for all co ^ TV. We therefore have:
= £*«""'®-'"[/„..Wf„+i))]
for all CO # N; and so there is a bounded ./^-measurable function H such that if = £ ' ^ [ / „ , i W r „ , i ) ) | ^ ]
(a.s.,F)
for / = 1 and 2. Since P^ equals P^ on .J^, this completes the induction.
D
6.2.4 Corollary. Let a:[0, cc) x R"^-^ S^ andb: [0, oo) x R'^ -• R"^ be bounded measurable functions. Suppose that for all (s, x) e [0, oo) x R'' and any pair of solutions P^ and P^ to the martingale problem for a and b starting from (s, x), one of the following holds: (i) £^'[jj (p(t, x(t)) dt] = £''^[JJ (p(t, x(t)) dt] for all T > s and cp in some determining class on [s, T) x R"^, (ii) £^'[Jr e-''(p(x(t)) dt] = £^'[Jr e-''(p(x(t)) dt] for all X>0 and cp in some determining class on R^. Then the solution is unique for each starting point (s, x). Proof From either (i) or (ii) it is easy to see that the conditions of Theorem 6.2.3 are met. D 6.2.5 Corollary. Let a:[0, oo) x R"^ ^ S^ andb: [0, oo) x R*^ -> R'^ be bounded measurable functions and assume that the martingale problem for a and b has at least one solution starting from each (s, x) e [0, oo) x R'^. Suppose that Q> ^ B(R'^) has the property that it generates B(R'^) under bounded pointwise convergence; and assume that for each T > 0 and (p e <^ there is anfj^ ^ e B([0, T] x R^) such that fr,,(s,x)
= E'[cp(x{T))]
"whenever P solves the martingale problem for a and b starting from (s, x) e [0, T) X R*^. Then the martingale problem for a and b is well-posed^ and the associated family of solutions is measurable. Proof Clearly uniqueness follows from Theorem 6.2.3. Let {P^.x* (-s, ^) e [0, oo) X R^] denote the family of solutions determined by a and b. We first note that (s, x)-^£^^--[i/^(x(r))] is measurable on [0,T)xR^ for all T>0 and \jj e B{R^). Indeed, this is true by hypothesis if i/^ e O, and the class oixj/ e B(R^) for which it is true is closed under bounded point-wise convergence. Thus we can, and will, assume that O = B(R%
6.3. Uniqueness: Some Examples
149
The rest of the proof is easy. Let n > 1, 0 = to < • • < ^„ < 5, and (po, ..., (p„ E B(R'') be given. Certainly (s, x) -^ E^'-icpoixito))' • • (p„(x(t„))] = (po(x) • • • (p„(x) is measurable on [t„, 00) x R'^. Next suppose that t„ < s < t^+i for some 0 < m < n. Define •A„(-) = „(•) and
for 0 < /c < «. Then by repeated applications of Theorem 6.2.1, E^''i(Po(x(to))"' (pMt,))] = M^)'"
(PmMft„,,,^„,M^ ^)
which explicitly gives (5, x)-^E''-i(po(x(to))' (p„(x(t„))] as a measurable function on [r^, r^+1) x R"^. D The following variant of Corollary 6.2.5 is the form in which we willfindit to be most useful. 6.2.6 Corollary. Let a: [0, 00) x R'^-^ S^ andb: [0, 00) x JR'^ -• R'^ be bounded measurable functions and assume that the martingale problem for a and b has at least one solution starting at each (s, x) e [0, 00) x R'^. Further, assume that for each 0<s
J (p(t,x(t))dt
whenever P solves the martingale problem for a and b starting from (s, x) e R"^. Then the martingale problem for a and b is well-posed and the associated family of solutions is measurable.
6.3. Uniqueness: Some Examples Before continuing with the preparations for our main uniqueness theorem, it is important to recognize that the contents of Chapters 3 and 5 provide us with some highly non-trivial examples of well-posed martingale problems. Indeed, when dealing with degenerate coefficients (i.e. if the matrix a(s, x) is allowed to be
150
6. The Martingale Formulation
singular at some point (s, x)), then the uniqueness theorems obtained by these methods are not covered by the one toward which we are working. On the other hand, it must be admitted that when Chapters 3 or 5 are appHcable, the martingale formulation does not add greatly to our understanding, although even here it does provide a unifying principle. 6.3.1 Lemma. Let a: [s, oo) x Q -• S^ and b: [s, co) x Q -^ R^ be bounded progressively measurable functions and suppose that P is a probability measure on (Q, J^) such that x{') ^ /d(a(*), &(•)) under P. Let ^ Q [s, oo) x R^ be an open set and define x = inf{r > s: (r, x(t)) i ^}. Then for any f e Cl ^(^) n Q(^) •fAT
f(tAT,x(tAT))-
J
f^-^Lu\fiu,x{u))du
is a P-martingale after time 5, where L„ /5 defined in terms of a(*) and b(') as in Chapter 4. Proof. Let {^„}f be a sequence of bounded open subsets of ^ such that if„ ^ ^n+i,n > 1, and ^ = \Jf J^n- By standard real-variable techniques, we can find {(p„}r ^ Cg^([0, oo) x R') such that Xj,„ <(pn< Xjf„.v Define/„ = cp, • / and T„ = inf{r >s:(t, x(t)) ^ je„}. Then fn{t,x{t))
- / ( ! ; +
Luj
fn(u,x{u))du
is a P-martingale after time s, and so is jAXr,
fn(t A T„ , X(t A T„)) -
J
L„ /„(w, x(u))
du.
But Lit A T„ , X(t A T„)) = f(t A X„ , X(t A T„))
and (J^-^Lu)fniu,x{u))
= ( | A + L , y ( " , x ( « ) ) , S < M < T.
Thus /(rAT„,x(rAT„)) - /
" ( I ; + L„j/(w,^(«))^w
is a P-martingale. Finally, T„ /^T, and so the proof of the lemma is completed by an obvious limit argument. D
6.3. Uniqueness: Some Examples
151
6.3.2 Theorem. Let a: [0, oo) x i?** -• S^ and b\[0, co) x R^ ^ R^ he bounded measurable functions and define L, accordingly. Suppose that one of the following conditions holds: (i) for each T > 0 and cp G Cg^(K'') there exists a function fe B([0, T] x R'^) with f(T, ') = (p(') and a sequence {/„} ^ Ct '([0, T] x R') n C,([0, T] x R') such that f„^f on [0, T] x R'^ and (dfjdt) + LJ„-^Q on [0, T] x R*^ boundedly and point-wise; (a) for each 7 > 0 and cp e Co([0, T) x R"^) there exists a function fe B([0, T] X R'^) with f(T, •) = 0 and a sequence {/„} £ Cl ^([0, 7) x R^) n Q([0, T] X R^) such thatf„^fon [0, T] x R'^ and (dfjdt) -\-LJ„^(p on [0, T) X i?** boundedly and point-wise. Then for each (s, x) e [0, oo) x R*^ there is at most one solution to the martingale problem for a and b starting from (s, x). In fact, if P solves starting from (s, x) and (i) obtains, then (3.1)
f(s,x) = E''[cp(x(T))];
and if (ii) obtains (3.2)
/(s, x) = E^ J (P(t,x{t)) dt
In particular, if(i) or (ii) holds and, in addition, there is a solution starting from each (s, x) e [0, T] X R^, then the martingale problem for a and b is well-posed and the associated family of solutions is measurable. Proof. In view of Theorem 6.2.3 and Corollaries 6.2.5 and 6.2.6, we need only show that (i) implies (3.1) and that (ii) implies (3.2). Since the proofs are so similar, we will only show that (i) implies (3.1). Let r > 0 and (p e CQ(R^) be given. Choose/and {/„} J' accordingly. By Lemma 6.3.1, with ^ = [0, T) X R^, f„(t A T, x(t A T)) - j ^ " 7 £ + L„ J/„(u, x(u)) du is a P-martingale after time s. Thus £''[/„(T,x(T))]-/„(s,x) = £' Upon letting n -+ oo, we get (3.1).
((^^^ufn)(u,x(u))du
D
6.3.3 Corollary. Let a: [0, oo) x R"^ ^ S^ and b:[0, oo) x R'^-* R*^ be bounded continuous functions having two bounded continuous spatial derivatives and define L, accordingly. Then the martingale problem for a and b is well-posed and the asso-
152
6. The Martingale Formulation
dated family of solutions {Ps,x' (^^ ^) ^ [0, oo) x R^} is Feller continuous (i.e., Ps„, x„ -^ Ps, X weakly if(s„, x„) -• (5, x)). Moreover, {P^ ^: (s, x) e [0, 00) x R^} coincides with the family constructed in Theorem 3,2.6. Proof. We can either use Theorem 3.2.6 or Theorem 6.1.7 to produce solutions starting from a given (s, x). To prove uniqueness, we will use (i) of Theorem 6.3.2. Indeed, in the proof of Theorem 3.2.6, we produced of time-inhomogeneous semigroup {r^,,: 0 < s < t} on Ct(R^) with the property that if cp e C?(R'') and /(5, •) = t , r>( •)» 0 < 5 < r, then/(r, •) = <^(•) and there is a sequence {f„}f c C^ 2([0, T) X i?'^) n Q([0, r ] X R*^) such that / „ - / on [0, T] x R'^ and (dfjdt) + L, /„ -• 0 on [0, T) x R'^ uniformly. Thus the proof that the martingale problem for a and b is well-posed is now complete. To prove that P^^ ^^ -* P^ ^ if (s„, x„) -• (s, x), note that {P^^^ x„' " ^ 1} is relatively weakly compact (cf. Theorem 1.4.6) and that any limit point is a solution to the martingale problem for a and b starting from (s, x) (cf. the proof of Theorem 6.1.7). Since there is only one such solution, namely F^, x' the proof is complete. D We could equally well have started with the result of Exercise 3.3.3 and thereby have proved the preceding corollary under the weaker assumptions on a and b given there. Similarly, we could have taken the assumptions on a and b given in Theorem 3.2.1, and, using that theorem, proved the analogue of Corollary 6.3.3 for such coefficients. Indeed, Theorem 6.3.2 is just waiting to be fed results from analysis. The preceding examples should provide ample evidence of its usefulness. We close this section with a discussion of the application of Ito's method to the study of the martingale problem. Actually, all the work has already been done, and all that remains for us to do is make a translation into the martingale formulation of results proved in Chapter 5. 6.3.4 Theorem. Let a: [0, 00) x R'^ ^ S^ and b:[0, 00) x R''-^ R'^ be bounded measurable functions and suppose that (T: [0, 00) x R^ -^ R'^ x R'^ is a bounded measurable function such that a = aa*. Assume that there is an A such that \\(7(t, x) - (T(t, y)\\ + Ib(t, x) - b{t,
y)\
for all X, y e R*^. Then the martingale problem for a and b is well-posed and the corresponding family of solutions {Ps,x'' (^^ x ) e [0, 00) x R'^} is Feller continuous. Proof. Existence follows from Theorem 5.3.1 or Theorem 6.1.7. As for uniqueness, Theorem 5.3.2 clearly suffices. Finally, the Feller continuity is proved in exactly the same way as in Theorem 6.3.3. D
6.4. Cameron-Martin-Girsanov Formula In this section we present the Cameron-Martin-Girsanov formula. Our reason for doing so at this point is that it will play an important role in our study of existence and uniqueness for the martingale problem; particularly when the
153
6.4. Cameron-Martin-Girsanov Formula
coefficient a is non-degenerate. The essence of the matter is contained in the following lemma. 6.4.1 Lemma. Let a : [0,oo) xQ -^Sd,b : [0,oo) xQ -• R"^, and c : [0,oo) xQ -> R'^ be progressively measurable functions such that a, b and (c, ac) are bounded. Let P be a probability measure on (Q, J^) such that x() ^ ^i{a{'),b(-)) and define b( •)) and define _
rvs
R(t) = exp j
2
ic(u), dx(u)y - - J
tvs
a(u)c(u)y du
where x(*) = x(') — ^[""^ b(u) du. Then there is a unique probability measure Q o (Q, ^) such that Q < P on MJor allt>0 and dQ/dP equals R(t) on J^,,t>0. Moreover, x(') ^ -^dW*)' (^ "•" ^^)(*)) tinder Q. Proofi Since (R{t\ Ji^, P) is a non-negative martingale (by Theorem 4.3.7) with mean 1, the measures 2^, r > 0, defined by
QX^)^E\R{t\A\AeM,, are consistently defined probabiUty measures. Thus, by Theorem 1.3.5, there is a unique Q on (Q, M) which coincides with Q, on Ji^ for all t > 0. Finally, if ^ e R^ s
+ c(u\a(u)(0 + c(u))}du , A (6-^ c(ul dx(u)}
1 r'^ 2 J <0 + c(u), a(u)(e + c(u))y du, A =E^ I exp I (0,x(ti) - x{s) - f\b-\- ac)(u)du) -\J\e,aiu)e)du],A].
Hence, by Theorem 4.2.1, x( •) - ^5(«( * M^ + «^)( *)) "^der Q. D
154
6. The Martingale Formulation
6.4.2 Theorem. Let a: [0, oo) x/?''-• 5^, b: [0, oo) x R'^ ^ R'^, and c: [0, oo) X R*^ -• R^ be measurable functions such that a, b, and
exp J
with x(t) = x(t) - jl^' b(u, x(u)) du, t > 0. Proof In view of Lemma 6.4.1, it suffices for us to check that if Q is a solution for a and b -\- ac starting from (5, x), then there is a solution P for a and b starting from (s, x) such that g <^ P on ^ , and dQ/dP on ^f is given by (4.1), f > 0. To this end, let rvs
x(t) = x(t) -
{b + ac)(u, x(u)) du,
t > 0,
and define, relative to Q, m = - \
\c(u, x(u)), dx(u)y,
r>0.
Then, by Lemma 6.4.1, there is a unique P such that JP Y(t) = exp 4(0 - X J
<<^(W' ^(")' «("' x(u))c(u, x(w))> du -
J vs
+ I
1 r'^' - - J
= -(!
6.4. Cameron-Martin-Girsanov Formula
155
and thereby finish the proof. However, this calculation is valid only if we can interpret all the stochastic integrals involved relative to Q. In order to use it to complete our proof, we must therefore show that
is the same no matter whether we interpret the stochastic integral relative to P or to Q. But, as in the proof of Lemma 4.3.2, we can find simple functions 9„: [0, oo) X Q-^/^'^such that sup sup i6„(t, ca), a(t, x{t, co))6„(t, co)} < oo n
t, (a
and (4.2)
[ \0„(t) - c{t, x(t)), a{t, x{t))(0„(t) - c{t, x(t)))y dt^O
in Q-probability for each T > 0. Since P and Q are equivalent on ^^ for all t > 0, this means that (4.2) holds in P-probability as well. Hence, for a given t > 0, there is a set N e ^ , such that P(N) = Q(N) = 0 and (4.3)
/ " ' < C ( M , X(U)\
dx(u)y = lim f^\e„(u\ n-*ao
dx(u)}
s
off N, whether the left hand side is interpreted relative to P or Q. Because the right hand side of (4.3) is defined independent of the measure involved, this completes the proof. D The importance of Theorem 6.4.2 to the study of the martingale problem should be obvious. Namely, it allows us to relate the questions of existence, uniqueness, and measurability for one martingale problem to the same questions about a second martingale problem. Typical of the sort of use to which we will be putting Theorem 6.4.2 is the following. 6.4.3 Theorem. Let a: [0, oo) x R*^ -• S^ andb: [0, oo) x R'^ -^ R^ be bounded measurable functions and assume that there is a X> 0 such that (4.4) <^, a(s, x)ey >X\e\\ (s, x) e [O, oo) x R"^ and 9 e R'^. Then, for each (5, x) e [0, 00) x R'^, the martingale problem for a and b starting from (5, x) has at least one (at most one) solution if and only if the martingale problem for a and 0 starting from (s, x) has at least one (at most one) solution; and if there is exactly one solution P^ ^ for a and 0, then the measure 2s, x given by
(4.5)
^2. dPs,.
tvs
exp
-^f
(b,a-'b}(u,x(u)) du
156
6. The Martingale Formulation
for all t >0 is the only solution for a and b. Finally, if the martingale problem for a and 0 (and therefore also for a and b) is well-posed, then the family of solutions associated with a and b is measurable if the one associated with a and 0 is. Proof Only the final assertion requires comment. Suppose that the martingale problem for a and 0 is well-posed and that the associated family {Ps,x' (^» ^ ) ^ [0, oo) X R'^} is measurable. Then the family [Q^ ^: (5, x) e [0, 00) x R'^] for a and b is given by (4.5). In particular, if r > 0 and A e J^^, then Q,, =
E'^iR%t),A]
where c =^ -al i b and jR^(r) = exp [
dx(u)} - - [
Thus all that we have to show is that
(s,x)-^E'^'iRM^] is measurable for all bounded measurable c: [0, 00) x K** -• R"^. Since the class of c for which this will be true is closed under bounded point-wise convergence, we need only prove it for bounded continuous c's. But if c is bounded and continuous, then for all (s, x): R'Xt) = lim„^^ @ J ( 0 ^^ Pj, ^-probability, where
®MO = exp f
<0„(w), dx(u)y - - f
ie„(ui a(u, x(u))eMy du
and 9n is the simple function given by 6n(t) = c({[nt]/n),x{[nt]/n)). Since (2)j(r,a)) is jointly measurable in s and o),(s,x) -^ £^*>^[@^(t),>l] is measurable for each n>l. Thus, it only remains to check that (g)J(r) -^ R'(f) in L^(P,^ ^) for all (5, x). But E''^'i\(H):,(t)\^]<e^''^'-'^E'
exp f^\2e„(u),dx(u)}
-^f^\2e„(u),ci(u,x(u))2e„(u)}du B(tvs-s)
= e
where B = sup, 3, \\a{t, };)||sup,^|c(r, y)\^; and therefore the £^(^5 j,)-convergence follows from the convergence in P^ j,-probabihty. D One immediate application of Theorem 4.6.3 to the martingale problem is that the martingale problem is well-posed and has a measurable family of solutions when a satisfies the conditions of Theorem 3.2.1 and b is merely bounded and measurable. A second application is the next corollary.
6.5. Uniqueness: Random Time Change
157
6.4.4 Corollary. Let a: [0, oo) x R'^-^S/ and b: [0, oo) x R'^^ R^ be bounded measurable functions such that a satisfies (4.4)/or some A > 0 and a(s, •) is continuousfi?rall s >0. Then the martingale problem for a and b has at least one solution starting from each (s, x) e [0, oo) x jR*^. Proof. Simply use Theorems 6.4.3 in conjunction with Theorem 6.1.7 applied to a and 0. n It should be emphasized that the Cameron-Martin-Girsanov formula (4.5) has many important apphcations to questions other than those coming from the martingale problem. Indeed, it provides the most explicit known expression of the analytic fact that the first order terms of L, can be thought of as a compact perturbation of the second order terms. We will have ample occasion to exploit this explicitness later on.
6.5. Uniqueness: Random Time Change Before proceeding with the main theme, we want to indulge in another brief digression Suppose that a: [0, oo) x R^ -• (0, oo) is a bounded, uniformly positive, measurable function and let P on (Q, ^) solve the martingale problem for a and 0 starting from (s, x). Then x(-) ^ ^\(a{% A')\ 0) under P. Following Exercise 4.6.5, define {T(r): r >0}by (5.1)
s+
a(w, x(u)) du = t,
t >s
and set p(t) = x(T(f)) — x, r > s. Then (j5(r), .^^t)^ P) is a Brownian motion after time s and (5.2)
x(0 =x + p(s+
/ a(w,x(u))duj,
t> s.
What we want to do in this section is explore the possibility of using (5.2) to study the martingale problem. In particular, if we can show that J^ a(u, x(w)) du must be a measurable functional of p( •), then (5.2) gives x( •) as a measurable functional of P{'); and therefore we will know that the distribution of x(-) is uniquely determined; that is, the martingale problem for a and 0 can have at most one solution. But j , a(u, x(u)) du is just the inverse of T(r), as a function off e [s, oo); and so this question of measurability can be transferred to x(t). Moreover,
(5.3)
^^^ = a^Wm''-''
'^'^^'-
158
6. The Martingale Formulation
From (5.3), it is clear that T(r) must be a measurable function of )?(•) if a, as a function of r, satisfies sufficiently strong regularity conditions. (For instance, if a is Lipschitz continuous with respect to t with a Lipschitz constant which is independent of X.) In particular, if a is independent of r, then T(r) = s +
I alBiu iM)
du,
t > s,
and is therefore obvious P( • )-measurable. In order to make the presentation as simple as possible, we will restrict our attention to the case when a is independent off. 6.5.1 Lemma. Let a:R^ -^ Sd and b:R^ -^ R^ he hounded measurahle functions. Then for each (s, x) € [0, oo) x i?^ there is a one-to-one correspondence hetween solutions to the martingale prohlemfor a and h starting from (0, x) and those starting from (s, x). In particular, existence (uniqueness) holds for solutions starting from (s, x) if existence (uniqueness) obtains for solutions starting from (0, x). Proof Given s > 0, define <^^: Q -• Q so that x(t, s(co)) = x((t - s) v 0, co), t > 0. For r > 0, note that <^~^ determines a mapping of J^^+s onto ^f. Moreover, if P solves the martingale problem for a and h starting from (0, x), then
\f{x{t2))-(
\f{x{u)) du, A
= £'' f(x(t^-s))-
C 0
\f{x(u))du,<S>:'A\ J
= £'' f(x(t,-s))-
[" 0
\f(x(u))du,;'A\ J
= £•" o - i f(Au))-
\\f(x(u))du,A
f o r / e C^(R% 5 < fi < ^2, and ^ e J^,^. 1 '^ .. (In this computation L = - ^ d\ -)
d^
^x,ax,^.?,'''<•)&^•>
Also P o ;^(x(t) = X, 0 < r < 5) = P(x(0) = x) = 1. Thus P -• P o ; 1 establishes a mapping from solutions starting from (0, x) into solutions starting from (5, x). It is easy to see, by essentially the same reasoning, that if ^^i Q - ^ Q is defined by x(r, ^^co) = x(t + 5, co), t > 0, then P -• P o ^ 7 ^ establishes the inverse mapping from solutions starting from (s, x) into those starting from (0, x). D Because of Lemma 6.5.1, we restrict our attention to solutions starting at (0, x) when studying the martingale problem for time-independent coefficients.
159
6.5. Uniqueness: Random Time Change
6.5.2 Theorem. Let O be the class of all bounded, measurable, uniformly positive functions q)\ R^ ^ (0, oo). Given q> e<^, define T^: [0, oo) x Q-)- [0, oo) by the relation: X^U, CO)
Jo
J
(p(x(u, (o))
and let S^'.Q^Q be the map determined by x(t, S^o)) = x(T^(r, co), co), r > 0. Then, for each co, T^(% CO) is a continuous, strictly increasing map of[0, co) onto itself; and for each t >0, x^(t, ') is a bounded stopping time. Furthermore, S^ is measurable from (Q, ^,(,)) onto (Q, Jit)for all t > 0. Finally, if a: R"^ -^ S^ and b:R^-^R'^ are bounded measurable functions and if P solves the martingale problem for a and b starting from (0, x), then P o S~^ solves the martingale problem for (p(')a(') and (p(')b(') starting from (0, x). Proof Everything except the last assertion is left to the reader. To prove the last assertion, l e t / e CoiR"^), 0 < tj < ^2, and A e J^^^ be given. Then p « s „ - i f{x{t,))E^'
(\(x(u))Lf(x{u))du,A
f(A%(h))) - j'V(x(T»))L/(x(T»)) du, s; M where L is defined in terms of a and b. Note that, by a change of variables, J
MO
cp(x(x^(u)))Lf{x(x^(u))) du =
Lf(x(u)) du,
t > 0.
Thus, since
[f(A%m -
JJ'''L/(X(U)) du,
ji,^,,„ pj
is a martingale and S~ M e J^^^^tx)^ PoSa,-i
E^'
\f(x(t^-^\{x(u)i)Lf(x{u))du,A ,Tfl>('i)
E^ / W v ( f i ) ) ) - f
Lf(xiu))du,S;'A
= £''-s» ' f{x{t,))-C
160
6. The Martingale Formulation
6.5.3 Lemma. Let and S^, (p e^, be defined as in the preceding. Then
for all cp,\j/e<^. In particular, S^ is invertible and its inverse is S^^. Proof. Clearly all that we have to show is that T^.^(r, oj) = T^(z^(t, S^ co), co). Since, for any ^ e O, T^(% CO) is characterized by the condition that T^(0, CO) = 0 and T^(r, (D) = ri(x(x^(t, co), co)), all that has to be checked is that if (T(t, co) = T^(T
5^C0), co), C0))T;(f, S^O))
= iA(x((T(t, co), (o))(p{x(x^(t, S^co), S^co)) = ll/(x((T(t, co), C0))(p(x(T^(x^(t,
= (p'il/(x((j(t,w),
and so the proof is complete.
S^O)), 0)), co))
0))),
D
6.5.4 Theorem. Let a: R"^ -^ Sj and b: R^ -^ R^ be bounded measurable functions and let (p: R*^ ^(0, co) be a bounded, measurable, uniformly positive function. For each X e R^, there is a one to one correspondence between solutions to the martingale problem for a and b starting from (0, x) and those for (p(')a(') and (p(')b(') starting from the same point. The correspondence sends a solution P for a and b into PoS~^, where S(p is described in Theorem 6.4.5. Theorem 6.5.4 is particularly satisfactory in the one-dimensional case. Indeed, combining Theorems 6.5.4 and 6.4.3 together with Lemma 6.5.1, we arrive at the following corollary. 6.5.5 Corollary. Let a: R^ -* (0, oo) be a bounded, measurable, uniformly positive function and b: R^ -^ R^ a bounded measurable function. Then the martingale problem for a and b is well-posed and the associated family of solutions is measurable. Proof Define S^:Q-^Q as in Theorem 6.5.2 and
6.6. Uniqueness: Localization
161
we are able to do so without having to invoke anything from the theory of partial differential equations. Unfortunately, we will not get off so lightly when we insist on being more ambitious.
6.6. Uniqueness: Localization In this section we will show that the basic questions about the martingale problem reduce to local considerations. To be precise, we will prove the following theorem, which will play an essential part in Chapter 7. 6.6.1 Theorem. Let a: [0, cc) x R'^ -^ S^ and b: [0, oo) x R'^ -^ R'^ be bounded measurable functions. Suppose that for each (s, x) e [0, oo) x R'^ there is an open set ^ 9 (s, x) and bounded measurable coefficients a and B such that: (i) a equals a and B equals b on ^, (a) the martingale problem for a and B is well-posed and the associated family of solutions is measurable. Then the martingale problem for a and b is well-posed and the associated family of solutions is measurable. The proof of this theorem will be broken into several steps, each one being a lemma. Thus in the statement of the following lemmas, we will assume that we are dealing with the situation described in Theorem 6.6.1 and will not mention it each time. 6.6.2 Lemma. Given R>0, set T^ = [0, R] x B(0, R). There exist open subsets Kj, ..., K^ of [0, oo) X R*^ covering T^ and bounded measurable coefficients ai, ..., 5jvf and B^, ..., B^ satisfying (ii) of Theorem 6.6.1 such that af„ = a and Bf„ = b on V„ for 1 < m < M. Moreover, the number p=
inf
max {(r - s) H- \y - x\'.t>
s
and
(r, y) 4 Vm}
(s, X) 6 T/i 1 < m < M
is Strictly positive. Proof. This lemma is an easy application of the Heine-Borel property for T^.
D
Now suppose that R>0 is given. Let Fj = [0, R) x B(0, R) and choose Fj, ..., K^ as in Lemma 6.6.2 for F^^ = F j . For each (s, x) e F^, let m(s, x) be the smallest 1 < m < M for which (r — s ) + |>^ — ^| > P whenever t > s and (r, y) ^ V^. We define the following stopping times on Q. First CR = inf{t > 0: (t, x(t)) ^ r»}.
162
6. The Martingale Formulation
Next, for 0 < s < K, set TQ = s and T^„_l + 1 if TJ_I
>CR
(inf{r > < _ i : (r, x(t)) i K„(.„_ ,s, ,(.„_,.))}) A CK if T^„_i < C R . Finally, if (s, x) ^ T^, let Q^^i be the measure on (Q, Ji) satisfying Qf\(x(t) = x, t > 0, = 1); and if (s, x) e F ^ , let Q^^l be the solution to the martingale problem for ^m(s,x) and 6^(s, ;c) starting from (s, x). Note that {Q^^i: (s, x) e [0, oo) x R^} is measurable. 6.6.3 Lemma. For each (s, x) e [0, oo) x R^ there is a unique P^^i on (Q, J^) with the properties that Pjf^ equals gjf i on Ji^s^; an^, /or each n > 0, Pjf^ coincides on ^ , s ^ ^ wi^/t Ps^i(x),5gifJ^(,s). Moreover, the family {Plfi: (s, x)6[0, oo) x jR'^} is measurable. Proof. Clearly, since Tj(a))/^oo as n -• oo for all co e Q, if such measures P^fi exist, they are unique and {Pl^l: (s, x) G [0, oo) x R'^} is measurable. To prove their existence define P^ ^ = Q^ ^ and, for n > 1, ^ s , X — -f^s, X 09T^ \lzf,, X(T^) •
If we can show that (6.1)
l i m P , % ( T j ^ i < r ) = 0,
T>0
then there is a unique P^^i on (Q, J^) such that P^f^i equals P" ^ on ^ ^ 5 ^ ^, n > 0, and clearly this is the measure we are looking for. The proof of (6.1) reduces to finding a ^ > 0 for which (6.2)
lim s u p P ; , , « ^ i < 5 + ^) = 0. n-»oo (s, x)
Indeed, from (6.2), we can use induction on k to show that (6.3)
lim sup PI ^(T;+ i <
S
+ /C^) = 0.
n-»oo (s, x)
To see this, assume (6.2) and (6.3). Then PIx«+1
< 5 + (/c + 1)^) < P - , ( T ^ , ^ ^ < s + kS) + PI,(TU
1 - T^-M < ^)
6.6. Uniqueness: Localization
163
for any 0 <m
as n — m -• 00; which completes the proof of (6.3) with (k + 1) replacing k. To prove (6.2), let A and B be the bounds of the coefficients a and b, respectively; and choose 0 < ^ < 1 so that BS < p and 2d exp[-(p - BSy/2 dAS] < i Note that for any (s, x) e [0, oo) x R*^ (6.4)
PlM<s
+ S)
Indeed, if (s, x) ^ F^, then x^ = s + I (a.s., Q^ ^), and so (6.4) is trivial. On the other hand, if (5, jc) e F^, then Pi, X(T2 < 5 4- ^) = Pi M <S-^S,T\<
CR)
since T2 = s + 1 if TJ > CK • But
sup | x ( r ) - x | > p l < i
\s
f
by the estimate in Theorem 4.2.1 applied to x{t /\x\) — j^^^^ h(u, x(u)) du. Thus (6.4) is estabhshed. Starting with (6.4), we have:
and so
which certainly proves (6.2). D 6.6.4 Lemma. For each R>0 and (s, x) e [0, 00) x R'^, the measure Pi^l described in Lemma 6.6.3 is uniquely characterized by the facts that Pl^l(x(t) = x, 0 < t < s)=land that x(-) ~ ^3tao,c,)(-W-. xOl Xio.a'M'^ xC))) under P««>.
164
6. The Martingale Formulation
Proof. We first show that x(-) - J^diXic^'H', x(')l X[o,iiA')H'^ 4'))) under 1^,%. To this end, set««(•) = Xio, U' H % 4 •)) and />^( •) = Xw, c«)( % 4 *)) H % x( •)), and let
Given/6 Co(R''), we must show that
9{t)^f(x{t))-(Li'''f{4u))du ''s
is a F^^j^-martingale after time s. This will be done by showing that, for all n > 0, 9{t A TJ) is a P^^i-martingale after time s. There is nothing to check when n = 0. Assume now that 6(t A TJ) is a P^^^-martingale after time 5. Observe that if T^(CO) < CR(CO), then
f(x{t))-(
L\m4u))du
is a eJ|l^),^(,s(,,) ,,)-martingale after time T'„(CO), where^ LJ^^ is the operator corresponding" to the coeflficients a^(xi(o», xixuco),
is a ^o>®r»QJf(L).x(t>),a»-niartingale after time s if <(co) < Ci?(w). On the other hand, if T^(CO) > Ci?(co), then Q,
e((tAzu,)AK)-e(zi) is a ^^^ ®^5(^) (2^^l5i>). x(t*(a>), tormartingale after time 5. Combining these remarks with the induction hypothesis and applying Theorem 6.1.2, we arrive at the conclusion that ^(r AT^+i) is a P^f^i (g),s Q^f^^^.s^-martingale after time s. Since Pi% equals Pi^l^r'Qr^^d") on ^^s^ , it is now easy to see that 9(tAx^„+i) is a P^^i-martingale after time s. We have therefore proved that x( •) ~ > 5 M - ) , M - ) ) under P
6.7. Exercises
165
set m(co) = m(T^(cL>), x(T'„(a)), co)) and P^ = ^o> (8)tco^r^t-) where F;"^ is the unique solution to the martingale problem for a^ and B^ starting from (r, y). Then, if CO ^ N and T^„(CO) < CRH. %(f) = a,(c.)(f, x(r)) and bj,(t) = B^^^^(t, x(r)), T^„(CO) < f < T% and so x(-) ~ ^"/^"^^d^iU'l ^miU')) ^^^^^ ^o- Using the uniqueness of solutions to the martingale problem for a^^^^ and S^(^), we now see that ®TMO)) ^?(w), x(T*(ca), to) for all (o$N satisfying Tj(a)) < CRI^O); and, therefore, for such CO we have that P equals ^a> ®t* 6tf(L), x(Tj(to),«) oi^ -^t'-- On the other hand, if CD ^ N and T;(CO) > (,R((r>\ then PS^i^) = ^('^nl^^)' <^)' t > T^(co)) = 1; and so P^ certainly equals d^ ® ,s(^) gif(L), x(t*„(a,), o)) on ^# in this case. Combining these remarks with the fact that T'^ = T^+ i (a.s., F^) for all co, we conclude that P does indeed equal P ® ^5 Q}^j^(^s) on Ji^s^ . Q It is now a relatively easy task to complete the proof of Theorem 6.6.1. Let us first establish the existence and measurabihty assertions. Let 0 < F i < • • < / ? „ ••• increase to infinity as M -• oo and use P^^"^^ to denote P^^i and („ to denote C^^^. For each /t > 1, {Fi"x- {^^ ^)^ [0,*oo) x R^] is measurable. Moreover, one can easily check that F^"^^^ ®^^(5j^(^^) satisfies the criteria characterizing F^"\. and so F^"J^^^ equals F^"|^ on Ji^^. Finally, by the estimate in Theorem 4.2.1, F?yC„
exp[-(rz - B{T - s)f/2Ad(T - s)]
for T > s such that n > B(T — s). (Here A and B stand for the bounds on a and b, respectively). Thus (cf. Theorem 1.1.10) for each (s,x) there is a unique Ps,x on (Q,J^) such that F^^^c equals F,^ on ^^„ for all n > 1. Clearly {Ps,x'is,x) € [0,oo) X R^} is measurable, and Fs,x solves the martingale problem for a and b starting from (5, x), Finally, uniqueness is proved as follows. Suppose F is a solution starting at (s, x). Given « > 1, note that F (E)^„(5;c(c„) equals P[% for all n > 1; and therefore F equals Pi% on .#^^, n > 1. But this means that F equals the P, ^ just constructed; and so we are done. D
6.7. Exercises 6.7.1. It is interesting to see what the martingale formulation looks like when dealing with discrete time processes. For this purpose consider the following set-up. Let J/' = {0, 1, ..., n, ...}, fi = (RY. and for n e ^r and co e S set x(n, (b) equal to the n^^ coordinate of co (i.e., "the position of co at time n"). Let ^„ = (j{x(k): 0
j(f{y)-f(x))n„(x,dy)
166
6. The Martingale Formulation
for / e Cj,(/?''). We will say that a probability measure P on (Si, ^) solves the martingale problem for A„ starting from (m, x) '\{P(x{n) — x,0
(/W«))-"l^/Wfc)),A,p) m
is a (discrete-parameter) martingale after time (m + 1) for all/ e Ci,(i?'*). Show that this martingale problem is well-posed. In fact, show that P solves it starting from (m, x) if and only if P(x(0)ero,...,x(n)€r„) = Zro(^) • • • Xr^(x) f
• • • f n^(x, ^3;i) ' ' ' n„_ i(>;„_ i, ^yj
for all n > m and Fj, ..., r„ e ,^|jd. 6.7.2. For some technical purposes, the following is sometimes a useful observation. Suppose that a: [0, oo) x R*^-^ S^ and b: [0, oo) x R'^ ^ Sj are bounded measurable functions and L, on CJ'(R'*) is defined accordingly. Next define a:R'^'-^ S,+1 and B: R'^'^R'^' by
|a'^(xo V 0, x) otherwise |o'(xo V 0, x) otherwise for X = (xo, x) G i?^ X R'^ = R'^'^^. Let L be the time independent operator on Co(R'^^^) with coefficients a and E. Given (5, x) G [0, 00) x i^'^, show that there is a one to one correspondence between solutions to the martingale problem for L, starting from x at time s and solutions to the martingale problem for L starting from X = (s, x) at time 0. In fact, show that if ^^i Q^ -• 0^+1 is given by x(r, ^,(ca)) = (r + s, x(t + 5, co)) then P solves the martingale problem for L, starting at (5, x) if and only if P o ^~ ^ solves the martingale problem for L starting from x = (s, x) at time 0. Conversel> check that the mapping O^; Qj+ j -• Q^ given by: x(r,
6.7. Exercises
167
6.7.3. Anyone who is familiar with Meyer's celebrated version of Doob's decomposition theorem might be wondering if the following formulation could not be used to advantage. Let a: [0, oo) X ^'^-•5d and 6: [0, oo) x K^->i?^ be bounded measurable functions and define L^ accordingly. Say that a P on (Q, ^ ) solves the submartingale problem for L^ starting from (s, x) if P(x(t) = x, 0
168
6. The Martingale Formulation
is a martingale for all/ e Co(R^). Check that ^ is a measurable subset of M(Q). This is easily done by writing down a countable number of relations for P to be a member of j / , each of which is the vanishing of a Borel functional on M(Q). Next define a map F\s^ ^R^ by insisting that F[x(0) = F(P)] = 1. Note that F is a one to one map of j ^ onto R'^ by the assumption that the martingale problem is well posed. Check that the mapping F is continuous and therefore measurable from j ^ onto R"^. By a theorem of Kuratowski [1948] a one to one measurable map (F) from any Borel subset (j^) of a Pohsh space (M(Q)) onto another Polish space (R**) has a measurable inverse (P^). That is to say P^ is a Borel map of R*^ into M(Q). 6.7.5. The Cameron-Martin formula has many important apphcations; one of which we now outline. More details can be found in Stroock and Varadhan [1970]. Let a: [0, cc) X Q-^ Sj and ^: [0, oo) x Q-*-P'^ be bounded progressively measurable functions such that (6, a{t, co)^> > A101^ for all f > 0, co e Q, 0 6 P'^, and some A > 0. Assume that P is a probability measure on (Q, J^) such that P(x(t) = X, 0 < r < s) = 1 and x(-) - J'M'l H')) under P. Show that for all T> s and e>0: (7.1)
P sup | x ( r ) - x | > e < 1. \s
The idea behind the proof of (7.1) is the following. First, assume that 5 = 0 and X = 0. Next, note that, because of Theorem 6.4.2, (7.1) holds for the given choice of a and b if and only if it holds for a and any other choice of bounded progressively measurable S: [0, oo) x Q-^P'^. The trick is to make a judicious choice of B. Namely, define
(7.2)
B{t)= -i^^^^^xui2.a\4tm)4t).
We now have to show that (7.1) holds when s = 0, x = 0, P(x(0) = 0), and x(')^J^d(a('l H'))' To this end, choose cpeC^iR'^) so that (p(x)= |xp for |x| <2£, and set r](t) = (p(x(t)). Then, by Ito's formula, rj(') ^ J^i((V(p(x(')% a(• )V<^(x(•)), y(')) under P, where y: [0, oo) x Q-^ P is a bounded progressively measurable function such that y(t, co) = 0 if e/2 < |x(r, co) | < 2e. But this means that so long as |x(•) | e [e/2, 2e], t](')is distributed, under P, like a 1-dimensional Brownian motion run with the clock Jo
6.7. Exercises
169
A > 0. Let Lt be the associated operator. For each (s, x) e [0, oo) x R'^, let P^ ^ be a solution to the martingale problem for a and b starting from (s, x). Using (7.1) and the Cameron-Martin formula, show that if ij/ e C([0, oo), R'^), then for all 0<s
sup \x(t)-x\s
f (A(w)^w| > e "s
< 1. I
In this way, conclude that the support of P^ ^^^^ coincides with the class C{s, x) of continuous (p: [0, oo) -• K** such that
/(s,x)<£^-[/(rAC,x(rAO)]. For more information about these and related matters, consult Stroock and Varadhan [1970] and [1972]. 6.7.6. The following simple example demonstrates that there exist bounded coefficients for which the associated martingale problem possesses no solutions. Consider, for instance, the time independent coefficients on R^ given by a(x) = 0 and b[x) = — 1 for x > 0 and 1 for x < 0. If P solves the martingale problem for a and b, starting from 0 at time 0, then P must be concentrated on paths CD satisfying x(r, o)) = I 6(x(s, co)) ds, •'o
for r > 0.
Show that there are no such paths and conclude thereby that no such P exists. 6.7.7. We know by Theorem 6.1.7 that if a and b are bounded and continuous then there is at least one solution to the martingale problem for any given starting point. It is easy to construct examples to show that uniqueness does not hold in that generahty. For example, on R \ define a(x) = 0 and /?(x) = |x I'' A 1 for some 0 < oc < 1. Show that the ordinary differential equation x(t) = b(x(t)) for r > 0 with x(0) = 0 has more than one solution. Conclude that the martingale problem for a and b starting from 0 at time 0 has more than one solution.
170
6. The Martingale Formulation
A more disturbing example of non-uniqueness is the following. On R^ define b(x) = 0 and let a(x) be any bounded continuous function which is positive for X 7^ 0 and 0 for x = 0 with JL i dx/a(x) < oo. (Take for instance a(x) = 1 A | X I"' for some 0 < a < 1). It is trivial to check that one solution to the martingale problem for a and b starting from 0 at time 0 is the process PQ with Po[x(t) = 0 for f > 0] = 1. We now outline the construction of another solution. Let i^Q be the one-dimensional Wiener measure and define a(r, co), for r > 0, by ait, CO)
I,
j ^
a(x(s, co))
Show that (T(t, co) is well defined and i^Q[(j(t, co) > 0] = 1 for r > 0 and use the technique of Section 6.5 to check that P^ = i^Q c (I)~ Ms a second solution where x(t, (!>(co)) = x(a(t, 0)), co). Even worse show that if /o(f) is the local time at 0 of x(t, co) (cf. Exercise 4.6.12) and 0 < A < oo, then P^ = i^Q o <^~ Ms also a solution where x(t, ^;^(co)) = x((T^(r, co), co) and (TXit, a At, O)) CO)
Jo
1
/
1
\
-\ds + -lo(ds)\ a(x(s, co))\ X I
Each of these P^ is a different solution and each is the member starting at 0 of a different Feller-continuous time homogeneous strong Markov family.
Chapter 7
Uniqueness
7.0. Introduction We are going to show in this chapter that if a: [0, oo) x R^ —• Sd and b: [0, oo) x R^ -^ R^ are bounded measurable functions and if a is uniformly positive definite on compacts and satisfies: (0.1)
lim sup ||a(s, y) - a(s, x)|| = 0 y-*x
0<s
for all T > 0 and x e R^, then the martingale problem for a and b is well-posed and the associated family of solutions {Ps,x- (^» x) G [0, oo) x R'^} is (strongly) Feller continuous. Most of the work involved is devoted to proving this resuh for the situation in which b = 0 and a is very nearly independent of x. Once this case has been thoroughly understood, the general case follows quite easily with the aid of the reduction procedures developed in Chapter 6. Since the details may obscure the idea, we now outline the reasoning behind our analysis. Let a: [0, oo) X R^ -^ S^ he measurable and have the property that (0.2) AI ^ |2 < <^, a(5, x)^>
OeR'
and
(s, x) e [0, oo) x R',
for some 0 < A < A < oo. Assume, in addition, that there is an x° G R*^ such that e = sup sup||a(5, x) — a(s, x^)||
(0.3)
s > 0 xe R
is small. Set c{s) = a(s, x^\ s > 0, and define
(0.4)
gf,(s, x;t,y)=
(2;t)-^/^(det C(s, t))'"" ^""Pf" 2<>^ " ^. C(s, t)' '(y - x)>]
for 0 < 5 < r, where C(s, t) =
c(u) du.
172
1. Uniqueness
For 0 < s < T a n d / e C^([0, T) x R% define (0.5)
Gj/(5, x) = fdt\
f(u y)gXs. x;uy) dy.
's
'R
Then GJ/solves du
+ Ll'^u = - /
in [0, T] with lim,/j^ w(f, •) = 0, where 1
d^
2 ip:i
dxi
dXj
Thus ds
+ LfiJf=-f+K^f,
0<s
where K^f(s,x) =
(L,-Lf)Gjf(s.x).
Note that K''f(s, x) I < -— max 2
d'Gjf
i
(s,x]
and therefore, if 1 < p < oo, then •^ / | | L P ( [ 0 , nxRd) ^^d
Cd(p, A, A)||/||^p([o, T]xRd)'
where Ca(p, A, A) is the constant in equation (0.4) of the appendix. In particular, if 1 < p < 00 is fixed and e > 0 is chosen so that ed^C,(P, K A) < 1, then I - K^ admits a bounded inverse on L^([0, T) x R'^); and, at least formally, we have: (0.6)
ds
+ Lfi^f=
-f
173
7.0. Introduction
in [0, T) with (0.7)
limG7"(r, •) = 0, t/T
where
It is therefore reasonable to hope that if P solves the martingale problem for a and 0 starting from some (s, x) e [0, T) x R"^, then GY(t A T, x(t AT))+
f(u, x{u)) du
will be a P-martingale and, therefore, (0.8)
- FP Gy(5, x) = E
f(u, x(u)) du
We should then apply Corollary 6.2.4 and conclude that there is only one such P. Obviously, theflawin this line of reasoning is that we have not been careful about the sense in which (0.6) holds. In fact, we have not even discussed in what sense G y is defined and (0.7) holds. This latter point is easily dispensed with, since by assuming that p is large enough we can guarantee that for/e Co'([0, T) x R**), G^f is well defined as a bounded continuous function on [0, T] x R*^ with G^f(T, •) = 0. On the other hand, the problem of interpreting (0.6) will not go away so easily. The fact is that the best that can be said about the second spatial derivatives of G^is that they exist as elements of L^([0, T) x R^). In other words, we can find functions (p^ 6 Cl ^([0, Qo)xR^) such that cp^ -^ G'^f uniformly on [0, T] X R^ and dt
+ L,(p, -> - /
in L^([0, T) X R^). Thus, in order to prove (0.8), we must show that, as a function of /, the right-hand side of (0.8) is continuous with respect to L^([0, T) x jR'')-convergence. Put another way, if we expect to extract a uniqueness proof out of this procedure, we must show a priori that for any P solving the martingale problem the measure // on [0, T) x R'^ defined by:
jfd^ = En\ f(u^x(u)) du admits a density which is in LP'{[0, This is the crux of our argument.
T) X R'^)
where p' satisfies (1/p) + (l/p') = 1.
174
7. Uniqueness
7.1. Uniqueness: Local Case In this section we will make a detailed study of martingale problems in which the diffusion coefficients are nearly independent of the space variables. The main focus of our discussion will be the question of uniqueness. However, starting with Lemma 7.1.8, we will be concerned with some regularity properties of the family of solutions to such martingale problems. Although these regularity properties are not needed to establish uniqueness, they will be used to answer some of the more refined questions of the type that will be taken up in the next section following Theorem 7.2.2. Let c:[0, oo)->"5d be a measurable function for which there exist 0 < A < A < 00 satisfying: (1.1)
/ i | ^ | 2 < < ^ , c(s)^>< A 1^1^
s>0
and
0 e R'.
Define g^s, x; r, >') for 0 < 5 < r as in (0.4) and Gj as in (0.5). 7.1.1
Lemma.
//
/ e C?([0, T] x/?'')
and
(d •¥ 2)/2 < p < oo,
then
GjfGC','-([0,T]xR'^)and (1.2)
|Gj/(5, x)| < A,(p, A, A)T^-<^^^>/^^||/||,,,,^).^^,
where AJ(p, A, A) depends only on d, p, A, and A. Moreover, if {E, ^, \x) is a probability space, {J*^,: t >0} a non-deer easing family of sub o-algebras of ^, and^\ [s, Qo)E^R^ a progressively measurable, right-continuous, ^-almost surely continuous function such that ^ ~ -^5(a, 0) for some bounded progressively measurable function a: [0, oo) x E-^S^, then Gjf(t A T, i{t A T)) + l " 7 ( u , i(u)) du - l " V / ( " , iM) du is a martingale after time s relative to (E, ^^, [x), where
Difiu, m) = \ ^ iy^(u) - cHu))0-^~y i(u)y Proof That Gjfe
C?' ""([0, T] x R"^) is obvious from the fact that G,
Moreover, the estimate in (1.2) follows easily from the inequality gXs, x;t,y)\<
^.^-^^—^^^^^exp[(2nX{t~- s)Y
\y - x|V2A(r - s)]
175
7.1. Uniqueness: Local Case
and Holder's inequality. Finally, to prove the last assertion, let {c„}f be a sequence of continuous functions on [0, oo) into S^, each of which satisfies (1.1), such that j
\c„(u)-c(u)\
du-^0.
Given / e C?([0, T] x R'^), note that GlfeCm[0,T]xR^) and that GJ^/-• G.V uniformly in [0, T] x R'^. Hence all spatial derivatives of G,^„/tend uniformly to the corresponding spatial derivatives of Gjf. Thus, since
dt
i,j=l
dxi dXj
we see that
1
dG^ f
1
'^
d^G'^ f
dt
tends uniformly to zero as « -^ oo. It is now clear that
Gjf(tAT,
i(tAT)) + j " /(i/, i(u)) du - f^Dlf(u,
i(u)) du
is a /x-martingale after time s, since for each w > 1,
has the same property.
Q
7.1.2 Lemma. Let a: [0, oo) x Q-^ S^be a progressively measurable function which satisfies J^\e\^ <(e,a{t,o))e)
r>0,
coeQ,
and
6 e R",
for some 0 < A < A < oo. Assume further that for some N > 1, a(t) is Ji[j^tyf^-measurable for all t > 0. Let P be a probability measure on (Q, Jt) and i'. [s, oo) X Q -> R*^ a right continuous, P-almost surely continuous progressively measurable function such that ^ ~ J^5(a('), 0) relative to (Q, ^ , , P). Then for each {d -\r 2)/2 < p < oo there is a constant C < oo depending only on d, p, A, A, T, and N such that
I f(t,i(t))dt\\
andfe
C^([s, T] x R'').
LP([s, t] X Rd)
176
7. Uniqueness
Proof. Set k = [Ns], K = [NT], and f, = (//JV v s) A T for /c < / < X + 1. Clearly:
f fit, m) dt
?p^.
\)(t,m)dt
l=k
where {Pj,} is a r.c.p.d. of P | ^ „ . Note that if co, denotes the element of Q such that x(t, CO,) = x(t A ti,o)),t> 0, then a(t, •) = a(f, w,), 0 < t < t,+,, Pj„-alniost surely. Hence, by Theorem 6.1.3 and Lemma 7.1.1, •(/+!
E'"" [jT f(tM))dt =
£'
[G'liWi'iM
for P-almost all w. Thus, by (1.2): E'""
\"*'f{t,i(t))dt
and so
j /(^ «(0) ^r
<(K-k
+ i K ( p , A, A)ri-<''^^>/^1|/|Up«., Dx^.).
Since K - k ^ I < {T - s + \)N, this completes the proof.
D
7.1.3 Lemma. Let a: [0, oo) x Q -> S^, {: [5, oo) x Q -^ K^ an^ P on (Q, . # ) fee as in Lemma 7.L2. Assume, in addition, that there exists a measurable c: [0, 00) ->• S^ satisfying (1.1) such that sup \\ait,co) - c{t)\\ < (d^Cdip^^lA))-' it,co)
for some (d +2)/2 < p < GO, where Q ( p , i , A) is given in (0.4) of the appendix. Then for any T > s and f € Cg^ils, T] x R^):
\f(t,m)dt
<2AM^,^)T'-^'^'^''nf\U,,ruRdy
Proof Given T > s, define // on ([s, T) x R'^, ^[S,T)^R<') by ;i(A x T) = ^^[JAZr(^(0) ^^] for A G i^[s,,] and T e ^^d. By Lemma 7.1.2, we know that ^ is absolutely continuous and that its density (p is an element of LP'{[S,T) X R'^) (here (1/p) -h (1/pO = !)• We want to estimate \\(p\\Lp'{[s,T]xR'iy To this end, let / € Cf^([s, T) x R^) be given. By Lemma 7.L1,
E^[Gjf(s, as))] = E^
j /(t/,(^(t/))^J-£Hf D^,f(u,i(u))du
177
7.1. Uniqueness: Local Case
Note that E-iGjfis, i{s))]\ < A/p, X, A)r'-«'+^>'^''||/|U^„,„.«., and
'>lfi-^i"))\^2^ch^ii
dXi dxj
(u. au))
Thus if
"^^MiimH) 2\i
then: X,(p, A, A)T'-«''-^»'^''||/II„„,.„.«., > j
dt^J(t,y)(p(t,y)dy T
2d'C,{p, A, A)
j dt ^ H(t, y)(p(t, y) dy
But this means that ^,(p,A,A)Ti-^'^^2)/^^||/||,,,,,).^., + 2 W\LP-ils,T)xRd)\\j
\\LPas,T)xRd) T
> f dt f f(u y)(p(U y) dy
since | | ^ | | L P ( [ . . T)XR4 < ^ ^ Q ( p , A, A ) | | / | U P ( [ S . T)xi?«*).
Because this holds for a l l / e C?([s, T] x R% we conclude that
which is what we wanted to prove. G 7.1.4 Theorem. Let a: [0, oo) x Q->Sj be a progressively measurable function satisfying: ^\e\^<(e,a(t,(o)9}
t>0,
COGQ,
and 6 e R^
178
7. Uniqueness
for some 0 < A < A < oo. Assume that there is a (d + 2)/2 < p < co and a measurable c: [0, oo) ->> 5^ satisfying (1.1) such that
supM.a,)-c(0|U^-^-^, where Q(p, A, A) is the constant in (0.4) of the appendix. Suppose that P is a probability measure on (fl, Ji)for which x(-)~ ^j(a(-), 0). Then for all T > s and fe Co"([s, T) X R"): (1.3)
j f(t, x{t)) dt
<2AM ^ ^)T'-"'^''"'VhHis.n^R^y
Proof Let p e Co(R^)bQ non-negative, supported on the right half-hne and satisfy j p(t) dt^ I. For e > 0, set p^{t) = i/ep(t/sX t e R^ and define ae(t) = f Pe(t - s)a(s) ds, c^t) = j p£(r - s)c(s) ds,
t > 0, t > 0.
For n > 1, define a^^ Jf) = a^([nt]/n), t > 0, and c^^ „(t) == c^([nt]/n). It is easy to check that for each e > 0 and n> 1: A^ „: [0, oo) x Q -• S^ is progressively measurable; a,, „(0 is ./#[„,j/„-measurable for f > 0; Me\'<(e,a,,„(t,(o)e)
and
A|^|^<<^,c,„(r)^>
for all r > 0, CO e Q, and eeR"^; and sup^^, ^) \\a,^ „{t, co) - c^, „(r)|| < l/d^Cd(p, A, A). Moreover, for each T > s: j
\\a,^„(t)-a(t)\\'dt
as « -> 00 and then £ \ 0 . Now define P(t) = j^ a"*(w) ^X(M), t > s. Then (j5(r), J^t,P) is a ^/-dimensional Brownian motion after time s and x(t) = x(s) -f- [ a^(u) dp(u\
t > s,
P-almost surely. Next set 4, „(0 - x(5) + j 4 „(M) ^J5(M),
r > 5.
7.1. Uniqueness: Local Case
179
For each £ > 0 and n> 1, 4,n(*) ~ ^d(^E,n('\ 0)» ^^^ therefore, by Lemma 7.1.3, J
f(t,i,,„(t))dt
< 2 ^ , ( p , A,A)T^-<'^+2>/2^
LP([s, T) X Rd)
for all 7 > 5 a n d / e CJ^([5, T] x R'^y But < 4 £ ' j ||«,.„(t)-a(t)||^
sup |4,„(r)-x(r)p s
and therefore tends to zero as M -» oo and then e\0. In particular, j fit, x{t)) dt
lim lim £\0
and this completes the proof.
j /(t,^,,„(f))dt
«->oo
D
7.1.5 Lemma. Let c\ [0, co)-^ S^ be a measurable function which satisfies (1.1)/or someO < A < A. Suppose that a: [0, oo) x R^ ^ S^ is a measurable function such that (1.4)
?i\e\^ <(e,a(t,y)0)
t>0,
yeR',
and ^ e R'',
and define
^'•/(s, ^)=^ _ iy'(s, X) - c''(s))^^(s. X)
(1.5)
for fe C2'([0, T) x i?''). Then for each 1 < p < oo, K'^ admits a unique extension (again denoted by K^) as a continuous operator on L^([0, T) x jR**) into itself and the extensions corresponding to different p's in (1, oo) are consistent. Finally, if 1 < Pi < P2 < oo and (1.6)
sup \\a(t, y) - c(t)\\ < l/d^C,(p, A, A),
p^
(t,y)
then I - K^ admits a bounded inverse (I - K^)" ^ on n([s, T) x R**) for Pi < p < P2^the bound on (I — K^)~ ^ is less than or equal to two, and (I — K^)~ ^ is consistently defined for different p's in [pi, P2]. Proof Clearly: KY(s,x)
i
d'Gjf dxi dxj
< i s u p \\a(t, y) - c(t)\\ X; (t,y)
i,j=
(s, x)
2\i
d^GJ [s,x) dx; dx,
180
7. Uniqueness
for fe C5'([0, T] x R^), This proves that K^ admits a unique extension to L^([0, T) X R^) and that the norm of the extension is no greater than d^l2Ci{p,X,\)sw^^^y)\\a(Uy)-c{t)\\. Moreover, since if / is in L^([0, T) x R^) n L«([0, T) X R'^), where 1 < p, ^ < oo, we can find {/„}f c cy([0, 7] x R'^) such that/„ -•/in both L^([0, T] x R'^) and L«([0, T] x R*^), we see that the extensions of K^ are consistently defined for different p's. Finally, if (1.6) holds, then the norm of K^ on lf([0, T) x R**), pj < p
0
It follows that for p e [pj, P2], (I — K^)~ ^ has norm no greater than 2 and that (/ — K^)~ ^ is consistently defined for p's in this range. D 7.1.6 Theorem. Let a: [0, 00) x R^ -^ S^be a hounded measurable function satisfying (1.4) for some 0 < A < A < 00. Suppose that there exists a measurable c: [0, 00) -• 5^ satisfying (1.1) and (d + 2)/2 < p < 00 such that (1.7)
sup||a(t,y)-c(0||
define K^ and (I - K^)~ ^ on L^([0, T) x R^) as in Lemma 7A.5 for each T > 0. Then the martingale problem for a and 0 is well-posed; and if{Ps,x'' (^» x) 6 [0, 00) x R*^} is the associated family of solutions, then for 0 < s < T, x e R"*. and fe C?([s, T] X R"): (1.8)
j /(t, x{t)) dt =
Gjo(i-K^)-^f(s,x).
In particular, {Ps,x'' (^» x) e [0, 00) x R**} is measurable. Proof We will first prove the existence of solutions. To this end, let p e Co(R^) be a non-negative function with ^^d p(x) dx = 1, and define p„(x) = n'^p(nx) for n> i and x e R**. Set a„(t, •) = Pn * ^(^' *)• Clearly a„(t, •) is continuous for all t > 0, (1.4) and (1.7) are satisfied with a„ replacing a, and for each compact A:
^ [0, 00) X R'* and 1 < ^ < 00
lim f [ \\a„{t, y) - a(t, y)\\^ dt dy = 0. Given (s, x) e [0, 00) x R*^, let P^"^ be a solution to the martingale problem for a„ and 0 starting from (s, x). Because the a„'s are bounded independent of n, {P^"^: n > 1} is weakly compact on Q. Let {P^"'^} be a convergent subsequence of {P^"^} and let P be its limit. We want to show that P solves for a and 0 starting from (5, x). This is easily seen to reduce to showing that if s < tj < ^2 and O: Q -• RMS
181
7.1. Uniqueness: Local Case
a bounded ^j^-measurable function, then for/e CoiR'^)'(1.9)
lim E
Pin)
O firnxit))
dt] = E'
dt
n'-*ao
where 1 **
d^
2i,j^i
dXidxj
and 1 **
a^
2f,i^i
'dXidxj
To prove (1.9), let £ > 0 be given and choose N so that (1.10)
||a„ (r, >^) - %(r, y)p rfy)''' < a
(f'' dt f
for all n' > N. Note that: hm sup
£P(«')
o I Vy(^(0) ^^
Q>\
LJ(x(t))dt
< lim sup
\E^ O f "(L, - Lr>)/(x(t)) dt since rP(''')
O f Vif(x(r)) dt\-^E'
as n' -• 00. But there is a constant M depending only on the bounds on O and the second derivatives of/such that ?p(-')
>Csupp(/)
(x(f))|ia„.(t, x(t)) - a^(t, x(t))\\ dt L ' 1
xiCdtl Vs
\\aAt.y)-as(Uy)rdyX'\ •'suppr/)
/
182
7. Uniqueness
where we have used Lemma 7.1.4 in the derivation of the last line. Thus, by (1.10), if n' > iV, then 'P(n')
(Lr-Lnf{4t))dt <2MA,(p,X,A)t\-^'^^^/^PB.
(D
Since it is obvious that i\' dt\ Vs
\\a(t,y)-ar,(Uy)\\Pdy]
''suppif)
' < s, I
we can derive exactly the same estimate for r^2
(DJ [L,-LT^)f(x(t))dt once we have proved that (1.3) obtains for our P. But (1.3) holds when the P there is replaced by F^" ^, and therefore it must also hold for the weak limit of {P^"'^}; that is, for our P. We have therefore derived the following estimate: lim sup £F(n', O j''li«y(x(r)) dA-E"" O \''Lf(x{t)) dt <4MAd{p,^,A)t\-^^^^^^^h. Since e > 0 was arbitrary, our proof of existence is now complete. In view of Corollary 6.2.6, uniqueness of solutions as well as the measurability of {Pj ^: (s, x) 6 [0, 00) X R"^} will follow as soon as we prove that (1.8) holds when Pj ^ is an arbitrary solution starting from (s, x). Given a solution P starting from (s, x) and T > s, we know, from Lemma 7.1.4, that there is a <^ e IF ([s, t) x K*^) such that T
T «
f f(t, x(t)) dt\=\ ''s
\
dt \ f(u y)(p(t, y) dy •'s
•'Rd
for a l l / e CS'([s, T] x P**). Given/e CS([s, T] x R% we now have from Lemma 7.1.1:
-Gjf(s,x)^-E^ -\
J /(t,x(r))t/t U E H J K^f(Ux(t))dt dti
[(I-K^)my)cp(t,y)dy,
7.1. Uniqueness: Local Case
183
Since this equation holds for all/e Co([s, T] x R'^) and both sides are continuous with respect to n([s, T] x R'^)-convergence, we conclude that Gjf(s. x)=fdt\[(I-
K^)f](t, yMt^ y) dy
for all fe n([s, T] x R% But (/ - K^) is invertible on B([s, T] x R% and therefore f /(r, x{t)) dt\ = \ dt\
f(u y)(p{U y) dy
for a l l / e C^([s, T] x R'*). The proof is now complete. Q
7.1.7 Corollary. Let a: [0, co) x R'^ ^ Sj be a measurable function satisfying the conditions of Theorem 7.1.6 and let b:[0, co) x R*^ -* R*^ be bounded and measurable Then the martingale problem for a and b is well-posed and the associated family {Ps,x'' (s^ x) e [0, oo) X R*^} is measurable and has the strong Markov property. Proof. This result is an immediate consequence of Theorem 7.1.6, Theorem 6.4.3, and Theorem 6.2.2. D We have not as yet taken full advantage of (1.8). Obviously (1.8) implies regularity properties for the dependence of F^^ ^ on (s, x). To be precise, suppose that a: [0, oo) X R^ -^ Sa satisfies the conditions of Theorem 7.1.6, let {Ps,x'- (^» ^) ^ [0, 00) X R''} denote the associated family of solutions to the martingale problem for a and 0, and denote by P(s, x; r, •) the corresponding transition probability function. Given T>0 and a bounded measurable (p: R*^-• C, set (1.11)
f(s,x) = ^(p(y)P(s,x;T,dy),
0<s
and x e R'^.
By the Chapman-Kobnogorov equation, one has: (1-12)
f(s,x)=jf(t,y)P(s,x;t,dy)
for all 5 < r < 7. Thus, if 0 < /i< T - s, then (1.13)
/(s, x)=^lf
dt\ f(t, y)P(5, x; t, dy).
We are now ready to prove the next result.
184
7. Uniqueness
7.1.8 Lemma. Let P(s, x\U ')be the transition probability function described in the preceding paragraph. Given h> 0 and e > 0, there is a Sf,(s) > 0 such that for any R'^ satisfying \G - s\ v \i - x\ < Sh(e): T > hand (s, x), (a, i) e[0,T-h]x j (p(y)P(s, x; T, dy) - j (p(y)P(cT, i; 7, dy)
j (p(y)P(s, x;s-hKdy)-
j (p(y)P((T, i; s + h, dy)
I
(p^(y)P(^, (^; 5 + /i, ^y)
-hh)-x\>R]
+ P,,,[\x(s + h)-x\>
R]).
Given £ > 0, use (2.1) of Chapter 4, to choose R> 0 such that P.,,[\x(s-^h)-x\>R]<s/2 for all (T, rj)e[(s — 1) vO, s] x B(x, 1). Having chosen this R, define fit.
•) = X[5,.+/.](0 j (pR(y)P(t. % 5 + /i, rf>;).
Note that by estimate (2.1) of Chapter 4: (1-15)
||/||LP([O.S+/.IXK-)<^||(P||,
where A depends only on d, p, A, R and h. Thus by (1.13) and (1.8), j >R(y)P(5, x; s -\- K dy) - j (pR(y)P((7, i; s + h, dy) =
-^\Gl^'F(s,x)-Gr'F(a,i)\
where F = (I - X*'^'')" y. But an easy computation shows that lim lim J
dt \ dy\g^s, x, t, y) - g,((j, ^, r, y)\P = 0
185
7.1. Uniqueness: Local Case
at a rate depending only on d, p, X, A and h (we have extended g^s, x, t, y) to be zero for t < s). Thus we can find 1 > ^ > 0 so that 1.16)
G^'Fis, x) - Gr'Fic, a i < ^ i|f llMio,../.,x«fl
so long as (a,i) € [(s-^) VO,s] x B(x,3). Combining (1.15) with (1.16) and substituting them into (1.14), we arrive at: (1.17)
j (p(y)P(s, x;s-\-Kdy)-
^ (p(y)P((^. ^; s-\-K dy) <e Up
whenever (cr, (^) e [(s - (5) v 0, s] x B(x, S). Finally suppose that 0
\(p{y)P{s^K",T.dy).
Then if s — tr and |^ — x\ are less than d, we have from (1.17) that f (f,(y)P{s, x;T,dy)-j
(p{y)Pia, ^; T, dy)
= J il/{y)P(s, x;s + h,dy)-j
>l/{y)P(
<eUH4
- ^
J\b,a-'b)(t,x(t))dt\
186
7. Uniqueness
relative to P^^ ^. By Theorem 6.4.3, we know that if r > s, then Q^ x< Ps,x on ^t and that R^{t) is the corresponding Radon-Nikodym derivative. Now let /i > 0 be given, and suppose that (cr, {), (5, x) e [0, T - h] x R'^ with (7 < s. Define (p^(y) = E^^^'i^l
0 < a < /i.
By the Markov property: £G'..[0] = E<^''i(p,(x(s + a))] for all f < s and 0 < a < /i. Thus: |£;Qs.x[(D]_£;Q.,|-o]| = \E'-iR%s + a)9«(x(5 H- a))] - £^-fK^(s + 0i)cpMs + «))] I < \E'^'icpMs + a))] - £^-^b,(x(5 + a))] I + Icp^E'^'ilR^s
+ a) - 11] + E'-i\R^(s
+ a) - 11]).
Clearly ||(p, II <||a)||, (£''-[ IR^(s + a) - 11])2 < £ ' ' - p ^ ( s + a) - i f ] = £^-[(R^(s + a))^] - 1 <exp{a||/.||VA}-l, and similarly: (E^-ilR'is
+ a) - l]f < exp{(5 + a - c)\\b\\^/X} - 1 = exp{a||^pM} - 1 + exp{a||/.p/A}(exp{(5 - a)\\b\m
" !)•
Given e > 0, first choose 0 < oc^ < h so that exp{a,||6PM}-l<(6/3)^ Then, choose ^^ > 0 so that exp{a, \\b\m{exp{p, \\b\m - 1) < (e/3)l Finally, use Lemma 7.1.8 to choose 0 < dhi^) < fi,, so that I £"'• '[
7.2. Uniqueness: Global Case
187
7.2. Uniqueness: Global Case Most of the work is now done. In order to obtain the result announced in Section 7.0, all that remains to do is feed the results of the preceding section into the machinery developed in Chapter 6. 7.2.1 Theorem. Let a: [0, oo) x R"^ -^^ S^ andb: [0, oo) x l^*^ -^ R^ be bounded measurable functions. Assume that for each T > 0 and x e R^: (2.1)
inf 0<s
(2.2)
inf <^, a(s, jc)^>/1 ^ 1^ > 0, deR
lim sup ||a(s, y) - a(s, x)\\ = 0 , y-^x
x e R"^.
0<s
Then the martingale problem for a and b is well posed and the corresponding family {Ps,x'- (sy ^) ^ [0, oo) X R'^} is measurable. In particular, for any s > 0 and any stopping time T > s, {(5c.®t(a,)^t(ca),x(t.a»} is a r.c.p.d. ofP,^^\J^,. Proof We will show that the hypotheses of Theorem 6.6.1 are met by a and b. Let (s, x) E [0, oo) X R"^ be given. Then we can find S > 0 and 0 < >1 < A < oo such that
x\e\'<(e,a(t,y)e)' - x | < (5, and ^ e jR'^. Let ^ + 2/2 < p < oo and denote by Cj(p, X, A) the constant in (0.4) of the appendix. Then we can find an 0 < £ < <3 so that '\{^ = [((s - e) V 0, s + e) X B(x, e)], then sup \a{t, y) - a(t, x)\\ < l/d^Cd(p, K A). Now let c: [0, oo) -• 5^ be a measurable function satisfying A |^ |^ < (^, c(r)^) < AI ^ 1^ for all r > 0 and B e R^ and equaling a(t, x) for r e [0, s + b\ Next, let a: [0, oo) X i?*^ -• Sj be the measurable function given by a{t. y) = lAu y)a[t, y) + (1 - i^{t, y))c{t). Clearly X\Q^ <{Q, a{t, y)d) < A |^|' for all (r, y) e [0, oo) x R^ and Q e R\ and sup ||a(r, y) - c{t)\\ < i/d'C,(p, A, A). (^y)
Finally, define E = b. Then Corollary 7.1.7 applies to a and 5, and so the martingale problem for a and b is well-posed and the associated family of solutions is measurable. We have therefore shown that the hypotheses of Theorem 6.6.1 are satisfied by a and b. D
188
7. Uniqueness
7.2.2 Theorem. Let a: [0, oo) x R^ -^S^bea bounded measurable function satisfying (2.1) and (2.2) and suppose that b\\0, co) x R^ ^R^ is a bounded measurable function such that sup sup (b(t, x), a~^(t, x)b(t, x)> < oo for all T > 0. 0
xeRd
Given (s, x) e [0, oo) x i?*^, denote by P^, ^ ^^ Qs,x^ respectively, the solution to the martingale problem for a and 0 and a and b starting from (s, x). Then, for each T > s, Qs ^ is absolutely continuous with respect to Pg ^on Jtj and exp \\a-'b(t,
x(t)\ dx(t)) -lf(b,
a-'b}(t, x(t)) dt
is its Radon-Nikodym derivative. Proof In view of Theorem 6.4.2, there is nothing to prove. D 7.2.3 Lemma. Let 5: [0, oo) x R** -• S^ and h: [0, co) x R*^-^ R"^ be bounded measurable functions for which the martingale problem is well-posed and the associated family of solutions {Ps,x' (s^ x) e [0, oo) x R*^} is measurable. Suppose that a: [0, oo) X R** -• Sd and b: [0, oo) x R*^ -* R'^ are bounded measurable functions such that a = a andb = hon[s, s -\- S) x B(x, S)for some(5, x)G [0, 00) x K"andS > 0, and let P solve the martingale problem for a and b starting from (5, x).IfO
where
A=
sup {t,y)e[0,
sup (e,a(t,y)e)/\e
oo)xi?«l d e R
Proof Let T = (inf{r > 5: |x(t) - x | > ^}) A (s + ^) and let Q — P (x)^(.)P^(.) j^(^(.) . ) .
Then Q solves for a and 5 starting from (s, x), and so Q = P^ j,. On the other hand, P = Q on ^ ^ , and therefore P equals P^ ^on Ji^. This means that
i£''[/(x(o)] - £'"[/(x(t))]i < [P(T < 0+p,At < mm = 2||/||P.,.(T<0
and the desired estimate follows now from (2.1) of Chapter 4. D 7.2.4 Theorem. Let a: [0, 00) x R'^ ^ S^ and b: [0, 00) x R*^ ^ R'^ be measurable functions which satisfy the hypotheses of Theorem 7.2.1, and denote by {P^, x' (5> x) e
7.2. Uniqueness: Global Case
189
[0, oo) X R*^} the associated family of solutions to the martingale problem for a and b. Given T > 0 and x e R^ let
A= inf
inf
<^4ii)^,
0 < r < r 9eR''\{0}
.
\"\
(e, a(t, x)e)
A=
sup
sup
•' |2
0 < r < r eei?''\{0}
I" I
and B = sup
sup I b(t, y) I.
0
Also, for d>0, define p(S)=
sup 0
sup \\a(t, y) - a(t, x)l \y-x\<8
Then for B> 0 and 0 < s < T, there is a S > 0, depending only on d, X, A, B, p(*), and T — s as well as e, such that \E''^'iO]-E'''''i
<e||0||
whenever (1>:Q -^ C is a bounded J^^(^ G(x(t)\t > T)ymeasurable function and ((T,{) € [0,r) X R^ satisfies |(T - s| V |J - x| < d. In particular, {Psy.{s,x) € [0,oo) y.R^} is strongly Feller continuous. Proof. Choose and fix {d -\-2)/2
^<«)M2')M
^\e\'<{e,a{t,y)e)
eeR\
and \a{t, y) - a(t, x)\\ < for all 0 < r < r and \y-x\
^^^'^^
1 d'C,(p, X/2,2A)
< a. Define
\a(t, y) if 0
|y - x | < a
190
7. Uniqueness
and b(t, y) if 0 < r < 7 and b(t.y) = 0 otherwise,
\y- x\
and let {Ps,x'(s,x) e [0,oo) x R^} be the family of solutions to the martingale problem for a and b (cf. Theorem 7.1.6 or Corollary 7.1.7). Now suppose that e > 0 is given. Choose t e (s, T) so that (t - s) < ^a/(l + B) and 8^ exp
{l-B^.-,)
/2dA(t - s) <8/3.
Next choose y > 0, in accordance with Theorem 7.1.9, so that
|£:^-[a)]-£^-^[o]| <^||a>|| for all bounded ^ : Q -• C which are ^'-measurable and (a, (^) e (0, t) x R'^ satisfying |cr — s| V |i — x|
(^ - B(f - s + S)\/2d A(r - 5 + S) <£/2.
Given a bounded O: Q -* C which is ./#^-measurable, set f(y) = E^'i^l
y e R'.
Then for (cr, i) e [0, t] x R'^ satisfying | ( T - S | V | { - X | < ( 5 : I £P..[o] - £^-[
- E'^'ifixm
I
We have used here Lemma 7.2.3 and the inequality ||/|| < 11 Oil.
7.3. Exercises 7.3.1. Let a: [0, oo) X R^ ^ Sa and b: [0, oo) X R"^ -^ R'^ be measurable functions satisfying the hypotheses of Theorem 7.2.1. Denote by {Ps,x'' (s, >^) e [0, oo) x R"^} and {Qs ^: (5, x) e [0, 00) x R'^} the family of solutions to the martingale problem for a and 0 and a and b, respectively. Show that for each (s, x) e [0, 00) x R'^ and
191
7.3. Exercises
T > s, Qsx o^ ^ T is absolutely continuous with respect to P^ ^. The idea is the following. Define b„: [0, oo) x R^-^ R' for « > 1 by
10 otherwise and let {P\"\: (t, y) e [0, oo) x R^} be the family associated with a and b„. By Moreover, Q^ ^ equals P^^\ onJij^^^, where Theorem 7.2.2, P^^^^ ^P^^onJij. T„ = inf{f > s: Ix{t) - x\
>n}.
From these observations, it is easy to conclude that Q^ ^ itself is absolutely continuous with respect to P^^ on Jij. With a Httle more work, one can identify the Radon-Nikodym derivative of Qs ^ with respect to P^ ,, on Jij. To this end, denote this derivative by R and check that TAt„
£^-[K|^^,J=exp j
TAln
{a-'b(u, x(u)\ dx{u))--\
(b,a-'by(u(x(u)) du
Therefore, T A tr
R = lim exp j
(a- 'b{u, x(u)\ dx(u)) - - 1
?, a~ 'b){u, x{u)) du
If one makes the extension of the stochastic integral calculus suggested in 4.6.10, one can rewrite this as R = exp j {a-'b{u,x(u)\dx{u))--\
1 '^
(b,a-'bXu,x(u)) du
7.3.2. Using an estimate of Alexandroff [1963] (see also Pucci [1966] and Krylov [1974]), one can show that if a: R*^ -• S*^ is a bounded continuous function and if A\e\^ > {9,a(x)e) > X\e\^ for some 0 < i < A and all 0 € i?'^ and x e R\ then for all p > ^, T > 0, R > 0 and / G CQ(R^) with supp(/) ^ B(0, R): (3.1)
^Px
\ f(At)) dt
|LP(i?<0'
where C depends only on A, A, p, T and R (not on the modulus of continuity of a). This estimate enables us to prove existence of solutions to the martingale problem for all bounded measurable a: R^ -^S*^ and b\\0,co)xR^-* R^ such that a(-) is uniformly positive definite on compact subsets of R^. By the reasoning we have already developed, we need only handle the case when 6 = 0 and A|^|^ > (0,
192
7. Uniqueness
a(x)9) > A10 p for all 6 e R' and x e R'^ for some 0 < A < A < oo. Given such an a("), choose {a„(*): n > 1} to be a sequence of continuous maps of R*^ -• S** so that M0\^< (e, a„(x)e) < A\d\^ for all x e R^ 6 e R^ and n > 1, and j
\\a„(x)-a(x)rdx-^0
B(0, R)
for all 1 < ^ < 00 and R < oo. Given x e R**, let P^"\ for n > 1, be the solutions to the martingale problem for
^"' = lU'(^)eiT. I
^ ^ j
Starting from (0, x). Note that {Pj*^; n > 1} is weakly conditionally compact and we can assume without loss of generality that it has a weak limit F as « - • oo. Using (3.1), show that i f / e CS'(R'^) and supp(/) cz B(0, R), then
(3.2)
LP(R<')»
where C is the same constant as in (3.1). Extend (3.1) and (3.2) to bounded measurable / having support in B(0, R). Finally use these estimates to show that for all 0 < ?! < r2, bounded continuous O: Q -• T a n d / e C5^(R''):
(3.3)
Q>f\Lf)(x(t))dt
lim E^'^"'O f "(Zi">/)(x(r)) dt
where L = i^a'-'(x) ^V^x^ ^Xj. Conclude that P is a solution for L starting from X. 7.3.3. When rf = 1, one can show that the martingale problem is well posed for any bounded measurable coefficients a: [0, oo) x R^-•(0, oo) and b: [0, 00) X R^ ^ R^ such that a is uniformly positive on compact sets. As usual, we need only look at the case in which b = 0 and a satisfies X < a(u y) < A for all (r, y) 6 [0, 00) X R**, for some 0 < A < A < oo. Given such an a, the idea is to show that Theorem 7.1.6 can be applied. Indeed, since 2> (d + 2)1d when ^ = 1, we can take p = 2. Next observe that Cj^l, A, A) in (0.4) of the Appendix is precisely 2/A when d = 1 and that sup(, y^ \a(r, y) - c(t) \ < A - A if c(-) = A. That is
sup \a(u y) - c(t)\ < ill - ^ ) A ' Q ( 2 , A, A).
7.3. Exercises
193
If 2(1 — (A/A)) < 1, then Theorem 7.1.6 appHes directly as it is stated. In general, note that the proof of Theorem 7.1.6 goes through essentially unaltered if one replaces (1.6) by sup \\a(t, y) - c(t)\\ < oi/d^C,(p, K A) («,y)
for some a < 2. The above method establishes the uniqueness; and existence can be seen from Exercise 7.3.2. Alternatively, consider a sequence ani^^x) of continuous coefficients satisfying uniformly the bounds 0 < >1 < a„(t,x) < A < oo for all n > 1, and converging to a(r,x) in the sense: lim \ dt \
I a„(r, x) - a(r, x) \'^dx = 0
for every l < ^ < o o , T > 0 and M > 0. Use the procedure described in the earlier part to derive the following bounds:
(3.4)
j f(t, x(t)) dt
where C depends only on A, A and T[here P"^ is the solution corresponding to a„(-,) starting from (s,x)]. Now use (3.4) instead of (3.1), and proceed as in Exercise 7.3.2. 7.3.4. It turns out that there is another case in which the type of reasoning used in 7.3.3 yields an interesting result. Namely, when d = 2 and a: R^ -• S^ is bounded and measurable and is uniformly positive definite on compact subsets of R^, the martingale problem for a and any bounded measurable b: [0, oo) x R^ -* R^ is well posed. The existence of course is covered by Exercise 7.3.2. The proof of uniqueness rests on the following procedure initially carried out by Krylov [1969]. Again we need only look at the case when b = 0 and for some 0 < A < A < oo, ^e\^ <(e, a(x)e) < A |^|^ for all OeR^Sindxe R'^. In addition, the results of Section 6.4 allow us to assume that Trace a(x) = 2 for all x e R^. Given such an a, set L = i YI, j= 1 ^'•'(^) ^V^^i ^Xj and define 00 ^ - M t
(^^•^^(^^ = 1 k ? ! '~''"''''"'fiy)dy for // > 0, X e R^ and fe Co(R^). The first important remark is that for each // > 0, there is an extension of G^ which is a continuous map from l}(R'^) into Cfc(R'^). Next, define
194
7. Uniqueness
for /i > 0 a n d / e C^{R'^). Note that {K,f){x)\=i
(a"W-l) + 2a'^x)
dxl
S'G,f{x) d'GJ(x) + (a"(x)-l) dxl dxi dx2
< i [ ( a " ( x ) - l ) ^ + (a'^(x))^]*
By elementary Fourier analysis: ''d'Gj,,
d'Gj
Li(W«-w«)-(a^j
dx
{AlJll±^lf(^)l2Ji
Moreover, since a^^(x) + a^^(x) •= 2, (a^H^) - 1)' + ( « ' ' W ) ' = detl/ - a(x)| < (1 - X(x)f where A(x) is the smaller of the two eigenvalues of a(x). Since .^ < i(x) < 1, we conclude that (3.5)
iX,/||,.,^.,<(l-A)^||/||,.,^.)
Starting with (3.5) and the fact that G^ maps l3(R^) boundedly onto Cb(R^l one can repeat the arguments leading up to Theorem 7.1.6, only now one should use the functional jg" e~^Y(x{t)) dt instead of j j /(^(O) ^^- The reader should arrive at the conclusion that the martingale problem is well posed for such an a, and the solution Px starting from (0, x) satisfies
fe''-mt))dt
Go(i-K)-'f(x).
7.3.5. Show that the families of solutions constructed in 7.3.3-7.3.4 are strongly Feller continuous in the sense described in Theorem 7.1.9.
Chapter 8
Itd's Uniqueness and Uniqueness to the Martingale Problem
8.0. Introduction The contents of this chapter are based on the work of Watanabe and Yamada [1971], What we will be trying to do is give a careful comparison of the notion of uniqueness natural to the martingale problem as opposed to the notion of uniqueness inherent in Ito's method. We have already seen indications that there is a distinction between these two notions (cf. Exercise 5.4.2) and that the distinction is intimately connected with questions of measurability of the path x(') with respect to j5( •) in the stochastic integral equation (0.1)
x(t) = x+\
G(U, x(u)) dp{u)-^ \
b(u,x(u))du.
Also, we have reason to suspect that the origin of any difference which exists has to do with the fact that Ito's notion of uniqueness is tied to the choice of c in (0.1) satisfying a = era*, whereas a does not appear in the martingale formulation. All in all, there are several unresolved questions originating from such considerations, and it is our purpose in this chapter to deal with some of them.
8.1. Results of Watanabe and Yamada The first thing that we have to do is give a precise definition of " Ito uniqueness." Unfortunately, the definition is somewhat awkward. Let a: [0, oo) x R'^-^R'^ X R"^ and b: [0, oo) x R^ -^ R'^ be bounded measurable functions and (s, x) G [0, oo) X R''. We will say that a and b satisfy Ito's uniqueness condition starting from (s, x) if and only if for every probability space (£, J^, /i), every non-decreasing family {J^,: t >0} of sub cr-algebras of J*^, and every triple of right-continuous, /i-almost surely continuous progressively measurable functions P:[0,oo) X E^R'^, c^: [0, oo) X £:->K^ and rj: [0, oo) x E-^R'^ such that (P(t), ^t, //) is a ^-dimensional Brownian motion and the equations at) = X + f^'(T(u, i(u)) dp(u) + fb(u,
^u)) du,
t > 0,
196
8. Ito's Uniqueness and Uniqueness to the Martingale Problem
and ri(t) = X -f I
(7(M, rj(u)) dp(u) + I
b(u, rj(u)) du,
t > 0,
hold fi-ahnost surely, one has ^( *) = ^( *) /^-almost surely. Although the preceding is entirely natural, it has an inherent deficiency. Namely, it is not clear how one can use it to compare solutions to Ito's equations involving different Brownian motions. In particular, how does one show that if cr and b satisfy Ito's uniqueness condition starting from (s, x) then there is at most one solution to the martingale problem for aa* and b starting from (s, x)? Moreover, if Ito uniqueness obtains, must it be true that the solution is measurable with respect to the underlying Brownian motion? The answers to these questions comprise the content of this section. As usual, it is enough to treat the case in which s = 0; and so we will restrict our attention to this case. Our first step is to give a characterization in terms of distributions of what it means for a process to be the solution of a stochastic differential equation. 8.1.1 Theorem. Let a: [0, oo) x R'^-^ R^ ® R'' andb: [0, oo) x R'^--^ R"^ be bounded measurable functions and define 5; [0, oo) x R^** -> Sjd cmd B: [0, oo) x R^'^ -• jR^** by
<•"
^••<-»=(-(.;';)' '"i")
and
(1.2)
Ht,iy,^))-('%').
Let (E, ^ , fi) be a probability space and suppose that ^: [0, oo) x E-^ R*^ and are ^^o, «> ^ ^-measurable, right-continuous, fi-almost surely continuous fiinctions. Set ^^ = G{{t,(s\ P(s)): 0<s
i(t) = y + (CT(U, ^U)) dp(u) + (b(u, ^
i(u)) du,
t>0
•'o
H-almost surely if and only if the distribution P on Q2d of the pair (i('),P{)) solves the martingale problem for a and b starting from {0,(y,0)). Proof We first prove the "only if" assertion. To this end, define a: [0, oo) x E-^R^'^^R'^ by
^(0=(''<''f ))
8.1. Results of Watanabe and Yamada
197
Since (p{t), ^ , , //) is a ^-dimensional Brownian motion, we know, from Theorem 4.3.8, that joa(u) dp(u) - /2d(^a*('l 0) with respect to (£, J^,, fi). But dd*(t) =
im) ^ (o) ^ {^^"^ "^^^"^ ^ {^"' ^^^"^' ^^"r^>^^0 ^ '"' /z-almost surely. Thus, (^'1 Pi')) - /UH%i{-)J(% bi;i{')Ji'))) with respect to (E,^t,fj), and /i[((^(0),i?(0)) = (>',0)] = 1. It is therefore obvious that P solves the martingale problem for a and b starting from (0,(y,0)). Now suppose that P solves the martingale problem for a and 5 starting from (0, (y, 0)). Then //[(^(O), m) = (y, 0)] = 1 and
(inp('))-/Ua('Ainp(^m h(% (i('l P(')))) with respect to (£, J^^,, //). In particular, (j5(-), J^^,, P) is a fidimensional Brownian motion. Moreover, if 9 e R^ is given and R^^ is defined by: e:[0,oo)xE-^
then, by Theorem 4.3.7, exp W i(t) - y - f'«.(«, i{M)) du - f'<7(u, i(u)) d^(u) [ *'o *'o ZJQ
is a martingale relative to (£, J^,, /^). But a simple computation yields
(•), a(-, (a-),/J(-)))0(-)>=O, and so
Jo |i-almost surely. This completes the proof
Jo D
8.1.2 Corollary. Let o", b, 5, and B be as in Theorem 8,1.1 and set a = era*. Given a solution P to the martingale problem for a and b starting from (0, y), there is a solution P to the martingale problem for a and h starting from (0, (y, 0)) such that P = P o n " \ where 11: Qid -^ ^d '^ defined as the natural projection of0.2d onto its first component when Q2d ^^ represented as Qj x Q^.
198
8. Ito's Uniqueness and Uniqueness to the Martingale Problem
Proof. This result is an immediate consequence of Theorems 4.5.2 and 8.1.1.
D
We next want to give a distributional characterization of Ito's uniqueness condition. 8.1.3 Lemma. Given bounded measurable functions a: [0, oo) x R^ -^R^ ®R^ and b: [0, oo) X i?'^ -> R"", define a: [0, oo) x R^' -> S2d and £ [0, oo) x R^"^ -^ R^^ by:
^^A^
;,r. v^-K'^')^*(''>^')
CT(U/)o*(u y')\
(1.4)
a[Ux)-y^^^^^,^^^^^^,^
a{u y^)c-{u
y'))
and (..5)
«">=(:!;:;:!)
for t >0 and x = {y^, y^) e R"^ x R'^ (= R^"^). Then a and b satisfy Ito's uniqueness condition startingfrom (0, y) ifand only ifany solution P to the martingale problem for a and B starting from (0, (y. y)) has the property that P(y^(t) = y^(t), r > 0) = 1. (Here/{•)={x,C),---,x,(-)) and y'(-) = (x,,,(-l...,x,A-)))Proof First suppose that a and b satisfy Ito's uniqueness condition starting from (0, y). Let P solve the martingale problem for a and b starting from (0, (y, y)). Define ^: [0, oo) x R^'^ -> R^^ (g) R^'^ by
Define a: [0, oo) x R""^ -^ S^^ and 6: [0, oo) x R^*^ ^ R"^^ by ,1 .,2\\
^/i-
{.A
,,2
and
Hui(y^,yM^K^m=(^^''%'''^). By Corollary 8.1.2, there is a solution P on Q^^ ( = ^id ^ ^id) to the martingale problem for a and 6 starting from (0, ((y, y\ (0, 0))) such that P = P o 11" \ where n projects Q2d x Qjd o^ito its first component. Thus, we need only show that Piy^i^) = y^it),t > 0) = L But (z^(),z2()) is a 2J-dimensional Brownian motion with respect to (Q4d,./#t,P) and, by Theorem 8.1.1,
+J '0
S(u, (y'(ul y^u))) du,
t>0
8.1. Results of Watanabe and Yamada
199
P~almost surely. That is, y^(t) = y+ (a(u, y'(u)) dz'(u) + (b(u, y'(u)) du, ^0
t>0
•'o
and y^t) = y-h\ G(U, y^'iu)) dz'(u) + [ b(u, y^u)) du, -'o •'o
t > 0,
F-almost surely. Thus, by Ito's uniqueness condition, P(y^(t) = y^(t), f > 0) = 1. Next suppose that for any P, P{y^(t) = y^(t), r > 0) = 1. Given a probability space (£, J*^, /z), a non-decreasing family {J*^^: t > 0}, and right-continuous, ^-almost surely continuous progressively measurable functions ^^: [0, oo) x E -> R^ ^^: [0, oo) X £ -• R'^ and j5: [0, oo) x E-^ R'^ such that (P(tl J^,, /z) is a f^-dimensional Brownian motion and ^'(0 = y+ (^(u, i\u)) dp(u) + (b(u, e(w)) du,
t > 0,
/i-almost surely for i = 1, 2, note that
Jo\%, r(w))/
f >0,
/i-almost surely. Thus, ^((^^0), i'(0)) = (>', y)) = 1 and (^H'). «'(')) ^ /2d(^(% ((^^(•), (^^(•))), f(-, ({^(•), (^^0)))) with respect to (£, J^,, /z). But this means that the distribution of ((^^(•), ^^(O) on Q2d solves the martingale problem for a and 5^ starting from (0, (y, y)), and therefore n(i\t) = {^(r), t>0)= 1. D We now have all the preliminaries. What we still have to do is learn how to compare solutions coming from different underlying Brownian motions. It is the technique for accomplishing this that is at the heart of Watanabe and Yamada's work. We begin with the following simple observation. 8.1.4 Lemma. Let a: [0, oo) x R"^ -^ R'^ 0 R"^ and b: [0, oo) x R'^ -^ R'^ be bounded measurable functions and define a and h by (1.1) and (7.2), respectively. Given a solution P to the martingale problem for a and b starting from (0, (y, 0)), let {P^,} be a r.c.pJ. of P\(T(Z{U): U > 0), where z(-) = (Xd+i(-), ..., X2d('))- Then for all t>0 and AEG(y{u): 0 < w < r), where >;(•)= (xi(-), ..., Xa(')l P.(A) = P(A\a(z(u): 0
200
8. Ito's Uniqueness and Uniqueness to the Martingale Problem
Proof. Let A e (T(y{u): 0 < w < r), Be a(z(u): 0 t). Since P(C\^t) = P(Q (a-s., P), we have: E^[P.(A), B nC] = P(An
B nC) = P(An
and C e (7(z(u) -
B)P(C)
= E^[P(A\a(z(u): 0
r)), B]P(C)
= E^[P(A\a(z(u): 0
r)), B n C].
Thus, since P. (A) and P(A \ (7(z{u): 0
O-(Z(M): M
> 0)-measurable,
Before stating the next theorem, we need to introduce some notation. Given CO G Q j d , let
3;^(-,co) = (xi(-,co), ...,Xd(-,co)), 3;2(-, cy) = (Xd+i(% co), ..., X2d(-, co)), z(%co) = (x2d+i(-,co), ...,X3d(-, ca)). Define the a-algebras J^l, ./#', ^ ^ , and J^ over Qj^ by ^ ; = (T(y(M): 0 < M < r), J^' =
Given a> e Q2d •> define y(%co) = (xi(-, co),..., Xd(-, co)) z(-,cy)= (xd+i(%co), ...,X2d(-,co)). Finally, define 11;: Qj^ -• Q2d» ^ = 1» 2, so that 3;(-,n,co) = y(-,co) z(-, niCo) = z(-,co). 8.1.5 Theorem. Let a: [0, oo) x R*^ -• R^ (g) R^ andb: [0, oo) x K*^ -* R'^ be bounded measurable functions^ and define a and b accordingly as in (1.1) and (1.2). Let P and Q on il2d ^<^h solve the martingale problem for a and i startingfrom (0, (y, 0)). Then there is a solution R on Q34 to the martingale problem for <7
a{t, y')a*{t, y')
o{t, y > * ( t , y')
<7c*{t, y')
a(t, y'Y <7(t, y') /
8.1. Results of Watanabe and Yamada
201
and
starting from (0, (y, y, 0)) such that P = R o rij ^ and ^ = R o JIj ^ Moreover, if and A e (T(y(u): u > 0), then {Pj is a r.c.pJ. of P\a(z(u): u>0) R ( n r U | ( 7 ( ^ 2 ^ yr)) = Pn,(.)M)(a.s.,R). Proof Define O,-: Qjd ^^3d,7 = 1. 2, by M-,co) if ^'(•'^^")=| 0 if
i=7 i^j
z(% (Djw) = z(% (o). Set Q' = Q°
C]
= £^[p.(a>r M), or ^c] = p(or H>I n c)) = PoOr^(^nC); and so P equals P o
(7(M, /(M))
•'o
dz{u) + f ^(M, •'o
/(M))
du,
t > 0,
P-almost surely, i = 1, 2. Thus, if we can show that z( •) is a ^/-dimensional Brownian motion with respect to (Qj^, Jt^, R) (JM^ = a{{y^{u\ y^(u), z(u)): 0 < w < t)\ then
y(0\ M
,M«,/(«))\
,/M«,/H)\
202
8. Ito's Uniqueness and Uniqueness to the Martingale Problem
K-almost surely relative to (^3^, M^, R)\ and the desired result is immediate from this plus Theorem 4.3.8. To prove that z(*) is a ^-dimensional Brownian motion with respect to (^3^, Jt,, R\ let Q eR\ ^
exp '•<e, Z(f2)> + 2
ti
^ A r\ B r\ C\
= £« exp i{Q,Ati))^^^h
¥,Br\c\
= £" exp i<^.z(tO> + ^ t i F,Br\c\ = £« exp and so the proof is complete.
'<e,z(r,)>4-4^ V A r\ B r\ C\
D
8.1.6 Corollary (Watanabe and Yamada). Given hounded measurable functions a: [0, 00) X R^ -* R'^ (^ R"^ and b: [0, 00) x R^ -^ R^ satisfying Ito's uniqueness condition starting from (5, y), set a = era*. Then there is at most one solution to the martingale problem for a and b starting from (5, y). In fact, for any z e R'^, there is at most one solution to the martingale problem for a and B starting from (5, (y, z)), where a and Bare given by (LI) and (1-2).
Proof Without loss of generality, we may and will assume that 5 = 0. Moreover, in view of Corollary 8.1.2, the first assertion follows from the second. Finally, we need only prove the second assertion when z = 0. Suppose that P and Q are two solutions to the martingale problem for a and b starting from (0, (>', 0)), and let R be the associated measure on Q^^ given in Theorem 8.1.5. Note that (in the notation introduced just before Theorem 8.1.5) the distribution of the pair (y^('), y^(')) under R solves the martingale problem for a and 6(cf. 1.4) and (1.5)) starting from (0, (y, y)). Thus, by Lemma 8.1.3, ,2i f > 0) = 1. But this means that (>^^( z(-)) and (i'^^.)^^^.)) have ^()^H0 = y^(^\ the same distribution under K, and therefore P = R
8.L7 Corollary. Let a: [0, 00) x R'-^ R'(g) R' and b: [0, 00) x R'-^ R' be bounded measurable functions and define a and b accordingly as in (L4) and (L5). Then 0 and b satisfy Ito's uniqueness condition starting from (s, y) if and only if there is at most one solution to the martingale problem for a and h starting from (s, (y, y)).
8.1. Results of Watanabe and Yamada
203
Proof. Again we take s = 0. Assume that Ito's uniqueness condition holds and let P and Q be solutions to the martingale problem for a and b starting from (0, (>', >^)). Then (in the notation of Lemma 8.1.3), P{y^(t) = y^(tl t>0) = Q(y^(t) = y^(t), r > 0) = 1. Moreover, the distribution of >;^(-) under both P and Q solves the martingale problem for aa* and b starting at (0, y). Thus, by Corollary 8.1.6, };^(-) has the same distribution under P as it does under Q. Combining these remarks, we see that P = Q. Next suppose that there is at most one P. Because of Lemma 8.L3, it suffices for us to show that if P exists, then P{y^{t) = y^{t\ r > 0) = 1. Given F, let Q be defined on Q^, by Q(y'(t,) e F j , ..., y\t,) e r„; y^'it,) e Aj, ..., ^ ^ ( g e A„) = P(y\t^)er^ n Ai, ..., y(t„) er„ n A„) for 0
i(t) = y + f
'^(w, «")) dP(u) + | ' " % , i(u)) du,
t>0
fi-almost surely, then, for each t >0, ^(t) differs from a <j(P(u): 0 0) = 1. Moreover, the construction of R in Theorem 8.1.5 is such that R(Uj^A\G(J^^VJjr)) = PTI,{)(A) (a.s., R) for A G (y{y{u):u > 0), where {Fa,} is a r.c.p.d. of P\o{z{u):u> 0). Thus, if F € ^^d, then
E'{(xMt))-P\y(t)eT)f] ^P{y{t)eT)-E\P.{y(t)ET),y(t)eY] =0
204
8. Ito's Uniqueness and Uniqueness to the Martingale Problem
Since
E''[P.(y{t)sr),y{t)er] =
E%P.(y{t)er),y(t)er]
= £''[Pn,()0''(0er),J'^(0en = R{y'{t)er,yHt)€r) = R{y'(t)er)
= PMt)er). Thus, by Lemma 8.1.4, Xr(y(t)) = P.(y(t) ^ r ) = P{y(t) e T \ a(z(u):
0
P-almost surely, and so y(t) = E^[y(t)IG(Z(U): 0
(a.s., P).
Q
8.2. More on Ito Uniqueness We are now going to develop a criterion for Ito uniqueness which, in certain special situations, turns out to provide the easiest proof of martingale uniqueness. The criterion which we have in mind as well as its proof is again due to Watanabe and Yamada [1971]. 8.2.1 Theorem. Let p: (0, oo)-^ (0, oo) be an increasing function satisfying: J. , du = 00. lim . n •'. P iu) Let a: [0, oo) x R^ -^ R^ (^ R^ and h: [0, oo) x i?^ -> R^ he hounded measurahle functions for which there is an M < oo such that \(j(s, x) - a(s, y)\ < M p ( | x - y | ) and \b(s,x)-b(s,y)\
<M|x->;|
for all s>0 and x, y e R^. Then o and b satisfy Ito's uniqueness condition starting from each (s, x) G [0, oo) x R^ In particular, if a = ff(7*, then the martingale problem for a and b is welUposed and the associated family {P^, x- (^^ x) e [0, oo) x R''} is Feller continuous.
205
8.2. More on Ito Uniqueness
Proof. Suppose that thefirstpart has been proved and let us check that the second part follows easily. Existence of solutions is obvious from the continuity in x of (T(t, x) and b(t, x) (cf. Theorem 6.1.7). Moreover, uniqueness is a consequence of Corollary 8.1.6. Finally, Feller continuity follows from uniqueness and the continuity of a(t, •) and b(t, •). We now turn to the proof of Ito uniqueness. As usual, we assume that s = 0. Let (£, ^, fi) be a probabiHty space, {^^: r > 0} a non-decreasing family of sub aalgebras, and i^: [0, oo) x E -^ R\i^: [0, oo) x E -^ R\ and p: [0, oo) x E-^ R^ right-continuous, /i-abnost surely continuous progressively measurable functions such that (P{t\ ^t, fi) is a 1-dimensional Brownian motion and ^'(r) = X + f(7{u, ^%u)) dp(u) + fb(u, ^\u)) du,
t > 0,
//-almost surely, i= 1, 2. Set
»;(•) = !'(•)-^^(OThen „it) =
\'{a{u,^^u))-a(u,e(u)))dp(u) *'o +
({b{u,i'(u))-b{u,e{u)))du,
r>0
/i-almost surely. In particular, if/e Cl(R)y then
E[f(ri(t))]-m = iE \(a{u,i'(u))-a(u,mum"(ri{u))du 4-£ ({b(u,4^u))-b(u,^^u)))f'(n('*)du ''o
(p'{\ri(u)\)\f"{r,(u))\
du
+ ME (\r,(u)\\n„(u))\du We now choose the sequence
{OLJ^Q
^ (0, 1) so that a^ 10 and
p\u)
du = k
206
8. Ito's Uniqueness and Uniqueness to the Martingale Problem
for /c > 1. For this choice of a^'s, define {(Pk}f ^ Co([0, oo)) so that (pk(x) = 0 for 0 < X < a^, 0 < (Pk(x) < 2/kp^{x) for a^ < x < a^. j , (Pk(x) = 0 for x > a^. i, and j?(p;;(x)^x=l.Set (Pk(x) = \
dt \ (p';(s) ds.
Clearly cp^ e C^(R\ On the other hand, (p^ is unbounded. Nonetheless, | q>iJ(x) \ < IXI, and so an easy limiting procedure can be used to check that
EWMt))]<^E
f pH\i{u)\)W'Mum du
+ ME ( \r,(u)\\<(,Uu))\
du
Thus
EWM))] < ME (HU]
du
Mt
for all k > I. Letting k-^ co and noting that (Pk(x)-• |x | as ^ -• oo, we conclude that E[\rj(t)\]<M
(E[\rj(u)\]du,
r>0.
Thus, by Gromwall's inequality, £[ |^/(f)|] = 0 for all f > 0, and so (^H*) = {^(O (a.s., /i). D It is interesting to see exactly how good this result is. First, let a: /^^ -• [0, oo) be a bounded, uniformly positive, smooth function away from 0 such that a(x) = IX f"'^ for IXI < 1 and some a > 0 and | (T'(X) | < M for | X | > 1 and some M < o o . I f a > l , then it is clear from the preceding that a and 0 satisfy Ito's uniqueness condition starting from any point, and therefore the martingale problem for (T^ and 0 is well-posed. On the other hand, if 0 < a < 1, then Theorem 8.2.1 does not apply. In fact, for a's in this range, as we saw in Exercise 6.7.7, the martingale problem for a^ and 0 is ill-posed. In Exercise 8.3.2 below, the reader is asked to demonstrate the invalidity of the natural analogue of Theorem 8.2.1 in higher dimensions. The idea is to exploit the same reasoning as that used in Exercise 6.7.7 and thereby show that in two or more dimensions the martingale problem for ^(T^( | X | )A starting from (0, 0) has at least two solutions as soon as a is less than 2. Combining these remarks, we see that the situations amenable to the techniques in Theorem 8.2.1 are quite special; but when they apply, such techniques can yield essentially optimal results.
8.3. Exercises
207
8.3. Exercises 8.3.1. Let (T: [0, oo) X K^ -• /?^ and ^: [0, oo) X Ri -• i?i be bounded measurable functions which satisfy Ito's uniqueness condition starting at any point. Let (£, J^, ft) be a probabihty space, {J^,: t >0} a. non-decreasing family of sub aalgebras of ^ , and p: [0, oo) x E^ R^ a 1-dimensional Brownian motion with respect to (£, J^,, /x). Suppose that x
r > 0,
and rj(t) = y + [ (T(U, rj(u)) dpu 4- [ b(u, i(u)) du,
t > 0,
^-almost surely. Show that ii(i(t) < rj{t), r > 0) = L A version of this result was proved using a very clever argument by A. V. Skorohod [1961]. The method we have in mind consists of the following steps: (i) let P on Q2 be the distribution of (^(•), rj(')) and show that P solves the martingale problem for a and 6 (cf. (1.4) and (1.5)) starting from (0, (x, y)); (a) let T = mf{t>0: y^(t)=-f(t)} and, for T > 0, let {PJ be a r.c.p.d. of PI^,^7-. Argue that for P-almost all co satisfying T(CO) < T, Poy(y^(t) < y^(t\ T(CO) < r < T ) = 1. 8.3.2. Let d>2 and suppose that a: R**-• [0, 00) is a bounded function which is smooth for x away from 0 and equal to | x I'' for | x | < 1 and some a > 0. Show that the martingale problem for a and 0 starting at (0, 0) has exactly one solution if and only if a > 2.
Chapter 9
Some Estimates on the Transition Probability Functions
9.0. Introduction In the derivation of our basic uniqueness theorem, Theorem 7.2.1, we made essential use of an analytic representation for the transition probability function associated with diffusion coefficients which are nearly independent of the spatial variables. We have already seen that this representation not only allows us to prove uniqueness, but also leads to regularity properties of the transition probability function as a function of the "backwards'' variables. The purpose of the present chapter is to exploit this same representation to conclude properties about the transition probability function as a function of its " forward *' variables. More specifically, we will show that the transition probabihty function has a density with respect to Lebesgue measure and that this density satisfies certain L'*-estimates. The procedure which we will use should be becoming familiar. That is, we start with the analytic representation obtained by perturbation theory. This gives us local estimates. We then use a probabilistic procedure to convert the local results into global ones. The estimates at which we eventually arrive in this way are not new to analysts. Fabes [1966] and Fabes and Riviere [1966] proved essentially the same estimates for parabolic equations of any order (not just second order); and their estimates together with our uniqueness results can be combined to yield most of the estimates which we will derive below. Thus our approach not only isn't the only one, it may not even be the most efficient one. Nonetheless, it is our belief that the approach that we give below is the most readily accessible one for probabilists and is of interest in and of itself. A word of warning is in order. Because we want to keep rather careful track of constants and their dependence on various parameters, we will put the equation number in which a constant first appears as a subscript. Thus Cj 5 means the constant which appears for the first time in Equation 1.5. If an explanation of the manner in which a constant depends on various parameters is required, such an explanation will be found somewhere in the paragraph containing the equation in which that constant is introduced. We hope that this procedure helps to clarify more than it obscures.
209
9.1. The Inhomogeneous Case
9.1. The Inhomogeneous Case Let c: [0, oo) -• Sj" be a measurable function for which there exists numbers A > 0 and A < 00 such that (1.1)
^e\^<(0,c{t)e)
t>0
and
OeR'.
Define C(s, t)= \ c(u) du,
(1.2)
0 < s < r,
and (1.3)
g'^\s,x;uy) =
(2nr\C(s,t)\^''^
for 0 < s < r and x, y e R^. It is easy to check that: g''\s, x; r, y) < (A/Xf'g'^\t - s, y - x)
(1.4) where (1.5)
g^'^Ht -s,y-x)
=
exp
\y-x
Moreover, a simple computation shows that: (1.6)
f dt\ •'O
^3;|^<^>(r,3;)r = Ci6T<''/2)[(d + 2)/d-r] ^Rd
for 1
Next, define (1.7)
Gjf(s, x) = f "^ ^r f dy g^^\s, x; t, y)f(u y), (s, x) € [0, T] x R^
for/e L*([0, T] x R*^). We now have the following simple estimate. 9.1.1 Lemma. Ifr^, r2 e [1, oo] and (1.8)
^ 1 0<
1
2 < d + 2'
210
9. Some Estimates on the Transition Probability Functions
then for T>Oandfe (1-9)
£^[0, T] x R") n L*([0, 7] x R^): ll<Jc/llL^2([0,r]x/?'') :^ Ci.9ll/llLM[(0,T]xi?'').
where
vvit/i Ci.6 computed for r = ri and rs is determined from
1 = 1 - 1 + 1. ''2
^3
''l
Proof Given/, note that for (s, x) e [0, T]x R'^: \Gjf(s,x)\
< (A/Af V * iA(s, x)
where
and Ht^y) =
X[o,T](t)\f(t,y)\-
Thus
Now define rj by 1 ^3
and note that I
1 r2
1 ri
1
+ 2)1 d. Hence, by (1.6),
and clearly \w\u\{RxK<^) = II/IIL^I([0,T]X/?'')
Finally, by Young's inequality, \^*W\u2{ByxR<^)
and so the proof is complete.
< \\(P\\uHRixRd)\\y^\\LniRxR
D
211
9.1. The Inhomogeneous Case
Next note that, by estimate (0.4) in the appendix, for each 1 < r < oo there is a Cj(r, A, A) < 00 such that (1.10)
S'G-f dxi dxj
< C d ( r , / I , A)||/||^.(o,r]XI?.) L'-([0,
T]xRd
for all/e C^([0, T] x R^) and 1 < ij < d. Remembering that the IT-bound of a linear operator is a log-convex function of 1/r, we see that for all 1 < r^ < r2 < oo there is an (1.11)
0 < e = ed(ri, r2. A, A)1
such that (1.12)
e^^Q(r, A, A) < 1,
Thus, if a: [0, oo) x R^ ^S^
r^ < r < r2.
is a measurable function satisfying sup ||a(5, x) - c(5)|| < e {s,x)
and (1.13)
Dl J{s,x)^\
I (a'^(5, x) - c%)) 1 ^ ^ ( s , x), ^x,- dXj
then \^a,cf\\Lr{[0,T]xRd)
^ 2 ^
^ 2
^'^^'*' ^' ' ^ ) I I / I U ' ' ( [ 0 ,
|L'-([0,
T]xRd)
T]xRd)
for Ti < r < r2. It follows that for fs in this range, (I — Dl^)~^ exists as a continuous operator on E([0, T] x R^) into itself having bound less than or equal to 2. Moreover, as pointed out in the proof of Lemma 7.1.5, the fact that (/ — Dj J~ Ms given by a Neumann series shows that it is consistently defined for these r's. Given 0 < A < A and 1 < r < oo, let j^(r, A, A) stand for the class of measurable a: [0, 00) x K'^-> S/ with the properties that:
Me\^<(e,a(s,x)e)
212
9. Some Estimates on the Transition Probability Functions
satisfies (1.14)
e^Ci(p, k,^)
r
m
V r,
where a(*) = a(*, x°). For a e j^{r, A, A), (/ — D^, a)~ ^ is consistently defined as a bounded operator on L^([0, T] x R'^) into itself for r < p < ((d + 4)/2) v r. Thus, for r < p < {(d + 4)/2) v r, X j / ^ G j o ( / - D i ; T, ) \ - l
(1.15)
is consistently defined as a bounded operator on If ([0, T] x R*^) into L^([0, T] X R% where 0< p
G
<:i—-. a -\-2
We are now ready to prove the next lemma. 9.1.2 Lemma. Let a e s/(r, X, A) and set (1.16)
N
Then (Klf^^ maps E([0,
(1.17)
\d-\-2 1
T] x R"^) into Q([0, T] x i?^• and, in fact,
\(K:r-'f(s,x)\
where C^^-j depends only on d, T, r, A, and A. Proof. First observe that for p > (d + 2)/2, Gj maps L^([0, oo) x i?**) continuously into Cj,([0, r ] X R"^) with a bound depending on d, T, p, X, and A. Thus if p > r and (d + 2)/2 < p < ((d + 4)/2) v r, so does Kj. In particular, there is nothing to prove if r > (^ + 2)/2. Assume that r <(d + 2)/2, and choose (d + 2)/2 < r^ < (d + 4)/2 so that ^H-2/1__n r
rsl
f^i I 2 r
Then we can find r = r© < • • • < r^ so that l/r^. i — l/rj^ = l/iV(l/r — 1/r^). By Lemma 9.1.1, Xj maps L"*" ^([0, T] x /?'*) continuously into r*([0, T] x R'^) with a bound depending only on d, T, l/N(l/r — 1/r^), >l, and A. Therefore, (Klf maps E([0, T] X R'^) continuously into £""([0, T] x R*^) with a bound depending only on d, T, 1/r - 1/r^, A, and A. Combining this with the first part of the proof, one arrives at the desired result. D
9.1. The Inhomogeneous Case
213
In order to translate these results into estimates on transition probability functions, we need the following simple observation. 9.1.3 Lemma. Let P(s, x;t, ')be a transition probability function with the property that (s, x; t)-^P(s, x;t,r) is measurable on {(s, x; t) e [0, oo) x R'' x [0, oo): s < t}for all r e^R,. Given T > 0, define P^ on B([0, oo) x R^) by T
P'^fis, x) = I dt dtj P(s, x; r, dy)f(t, y). 'R4
Then (P^ffis,
(1.18)
X) = - 1 j V - sf dt I P{s, x; t, dy)f{t, y).
Proof. There is nothing to prove if N = 0. Assuming (1.18) for iV — 1, we have: (P^r^y(5, x) = j
^
^ j\t -sf-'dt
jj(s, x; t, dy)P^fit, y)
=(jv~i)! \y - ^)~" * (''' ij^'' ^''' ^^wi^' ^) iV ! Jg
JRd
We are at last ready to give our preliminary estimates. Let a e j^{r, X, A) for some 1 < r < 00. Then, by Theorem 7.1.6, the martingale problem for 1
'^
2 i^j^i
d^ dxi dx
is well-posed. Denote by {Ps,x* (s^ x) e [0, oo) x R'^} the corresponding family of solutions. Then {P^, ^: (s, x) e [0, oo) x R'^} forms a strong Markov family which is strongly Feller continuous. In fact, if P(s, x; t, •) is the associated transition probability function, then (1.19)
Py(5, x) = j dtj P(s, x; r, dy)f(t, y) dt = Kj/(s, x)
if/e L^([0, T] X R**) n B([0, T] x R**) for some r
( P T " y ( ^ , x ) = (Kj)^^y(s,x)
4)/2) v r. Hence
214
9. Some Estimates on the Transition Probability Functions
for any A^ > 0 and any fe Co([0, T) x R% In particular, if N = [(d + 2)/2r], then (1.21)
\(P'f*'f{s,x)\
9.1.4 Lemma. If a e s/(r, A, A) and P(s, x; t, ')is the transition probability function determined by 1
^
d^ dXi dXj'
«.j=i
then for r < p < oo (1.22)
I {t - sf dt j P{s, x; t, dy)f(t, y) ^
^1.22
11/ ||LP([0,
oo)xRd),
where (1.23)
N=
d + 21 2 r
and Ci.22 = (A^JCi.17) v(T^"'V(N + 1)). Moreoverjoreachd > Oandr < p < oo T
(1.24)
P{5, x; r, dy)f(t, y) dy
[ dt [
2A\\J
||LP([0,
T]xR
'Rd\B{x, d)
where Cj 24 depends only on d, T, 3, p, r, A, and A. Proof Combining (1.21), (1.18), and the fact that Jj^^ P(s, x;t,dy)= 1 for r > 5, we arrive at (1.22) by an easy interpolation argument. To prove (1.24), we assume that 7 > 5 + 1 and note that [ dt [
P(s, x; t, dy)f(t, y)\ 00
s + (l/«)
< Z
dt
«=1
*'s+l/(n+l)
+ f
(t-sf
•'s+ 1
P(s,x;t,dy)\f(t^y)\ ''Rd\B(x,d)
dt f
P(s,x;t,dy}\f{t,y)\.
•'R'
The final term is estimated by (1.22). To handle the terms in the sum, define Al:W=['
"
*'s+l/(n+l)
dt\
P(s, x; t, dyMt^ y)''R<'\B(x,d)
9.1. The Inhomogeneous Case
215
As a linear functional on E([0, T] x R% A^;i\ is bounded by iV! (w + I f Ci 17. On the other hand, if 9 e B([0, T] x R% then sup
\A':\(P\
\s
\x(t)-x\>s)
+ (l/n)
I
<'^\\(p\\e-"''i^''\
n
Thus, by interpolation, if 1/p = (1 - ^)/r, then |A?W|
<(N\(n+\rc,.,,Y-'i^-^e-'"''^^'^\W\Uo.n
Rd)
and clearly this proves (1.24). D In order to get away from coefficients which are nearly independent of x, we will use the following localization procedure. The results which we are going to prove in this connection are slightly more refined than are absolutely needed in the present context; however, the generality in which we prove them is useful when dealing with unbounded coefficients of the sort often encountered when dealing with diffusions on a manifold. 9.1.5 Lemma. Leta:[Q, co) x R^ ^S^ and B: [0, 00) x K*^ -• R'^ be bounded measurable functions with the property that the martingale problem for
is well-posed and determines a measurable strong Markov family {Ps,x'' (^^ ^) ^ [0, 00) X R^]. Let a:[0, oo) X R^ -^ Sdandb:[0, 00) X R'^ ^ R** be bounded measurable functions such that a = a and b = Bon [s, T] x G, where 0 <s
Starting at (5, x) and x = inf{r > s: x(t) 4 G\ then P(x(t) 6 r, T > f) < P,, J(x(t) 6 r n G) for s
andT e ^^^.
216
9. Some Estimates on the Transition Probability Functions
Proof. Put R = P(S)^^Tpz^T,x(t.T)' Then R = P,,, and R equals P on ^.^jThus, if s < r < r and r e ^^,, then P(x(t) e r, T > r) = R(x(r) E T, T > r) < R(x(t) eV n G)
=^P,,Mt)^^r^G).
D
9.1.6 Lemma. Let a: [s^, oo) x Q -• S^ antf fc: [s^, oo) x Q -• jR** 6e progressively measurable functions after time s^ and suppose that P is a probability measure on (Q, ^) such that x( •) - / f («, b). Given 0 < R^ < R2andx^ e i?^ define T_ i = s® T2„ = inf{t > T2„_ 1: I x(t) - x^ I = R2I
n>0
T2„+i = inf{r > T2„: |x(r) - x° | = /^i},
n>0.
IfT>s'^ and A = sup{||a(f, co)\\; (r, G)): 5^ < r < r B = sup{|% (y)|; (r, w): s^ < r < r
and |x(r, co) - x^ | < /?2} an^i |x(f, co) - x^| < R2},
then (1.25)
where Cj 25 c^« ^^ chosen to depend only on d, T, R2 — Ri, A, and B. Proof. Let (7 be a bounded stopping time after time 5° and define C = inf{r>(T: |x(r)-x((T)| = R2}. We will show that if ^ > 0 is chosen so that 2d exp and BS
(K2 - ^1 - Bd)'
1
IdAd
R^ then for any H e J^„ satisfying H ^{\x(a) - x^\ < R^ P({C„-G<S}nH)
To do this, let {PJ be a r.c.p.d. of P | ci^^ and define xM = ^(^ A C,(,,)) - x(^(w), co) -
b(u) du.
9.1. The Inhomogeneous Case
217
Then for P-almost all a, x^{-) ~ /J<'"'(Zi„«,,?„„,)(•)«(•). 0) under P„. Thus, for P-almost all o) 6 H: PS. -c
= P„(C,,„, -
sup \(T{lO)
\xS)\^R2-Ri-Bd) +d
J
<2d exp[- (R2 - ^1 - BSy/l dAS] < i and clearly (1.26) follows from this. We now use (1.26) to prove that (1.27)
P(r2„<s' +
S)<(i)\n>0.
There is nothing to prove if n = 0. Assuming (1.27) for n, we have: P{r2n^2 < 5« + (5) = P(T2n^2 < ^^ + <5, T^^^i < 5^ + ^)
< P({C -cj<S}n{a<s''
+ S})
where a = T2„+I A (s^ + d). Since J^a B W < s^-^ S} ^ {\X((T) - x^\ = Ri}, (1.26) implies that P(^2n^2 <S'^S)<
iP((7 < S« + ^) < iP(T2„ < S« + S).
and so (1.27) follows by induction. Next set ko = [T - s^/S] + 1 and choose WQ > 1 so that /co(i)"^ < i- We will show that (1.28)
P(r2Uno<s' + k3)
k>L
When /c = 1, (1.28) is the same as (1.27) with n = HQ. Assume (1.28) for k. Then P(^2ik^ l)no < 5^ + (/C + 1)^) < P ( T 2 , „ O ^ ^^ +
^^)
+ P(T2(fe+ l)no < S^ + (^ + 1)^, ^2kno > ^'^ + ^^)-
The first term on the right is less than or equal to k{j)"^. As for the second, let {P^} be a r.c.p.d. of P\J^„ with a = T2fcno^ (-^^ + (^ + 1)^) and for each co define <^ n > — 1, the same way as we did T„, M > — 1, only now with 5^ replaced by (T(CO). Then 1(2*+ i)„o = C (^-s., PJ if T2fc„o(<^) < 5^ + (/c + 1)^; and for P-ahnost all co, x(.)^yj(^) (a, 6) under P^. Thus P^(x"^, < (T((O) + S) < (\p (a.s., P) and therefore P(T2(fc+l)«o < 5^ + (fc + 1)^, T2fc;^ > S^ + /C<5) = £^[P.(T„,
218
9. Some Estimates on the Transition Probability Functions
This completes the proof of (1.28). In particular, if /c = /c© in (1.28), we arrive at (1-29)
P(T2,
where iV = /CQ WQ • Proceeding in exactly the same way as we did in the derivation of (1.26) from the estimate P(x2 < s^ -\- 3) <j, we conclude from (1.29) that (1.30)
P(T2„N <
7^) <
(i)",
n >
0.
Since Z XisO, T]('^2n)
2^ X[so, ny^inN)
= ^ I pi^ins < n (1.25) is an immediate consequence of (1.30).
•
9.1.7 Lemma. L^r a: [0, oo) x R"^ -• S^ and b: [0, oo) x R'^ ^ R'^ be bounded measurable functions and let P solve the martingale problem for I d
^2
d
Q
dX;
Starting from (s^, x^). Suppose that a: [0, oo) x R*^-^ S^ and B: [0, oo) x R"^-^ R' are bounded measurable functions such that the martingale problem for i
d
^2 dXidxj
i,j=l
d
3
i=i
dX:
is well-posed and determines the measurable strong Markov family {Ps,x' (^^ ^) ^ [^^ 00) X R"^}. Finally, assume that for some T > s^ and K > 0, a(t, y) = d(t, y) and B(x^, R). Then for each 0 < 6 < R, b(t, y) = b(t, y)ifs^
I f(t. x(0) dt
<E
PsO, x l
j
+ Ci.3i
\f(t,x{t))\dt sup
£^-
I
\f{t,x(t))\dt
sO<s
for allfe Co([s^, T] x B(x^, R — 3)). The constant Ci^i can be chosen to depend only on d, T — s^, 3, sup
\\a(t,y)l
and
sup
sO
sO
\y-xO\
\y-xO\
\b(t,y)\.
219
9.1. The Inhomogeneous Case
Proof. Let Rj = ^» ^^^ ^i = R - ^^ and define T„ for M > — 1 accordingly as in Lemma 9.1.6. Then ff(t,x(t))dt\\<E'\\""'^\f(t,x(t))\dt J""
\f(t,x{t))\dt
Note that by Lemma 9.1.5 TOAT
f
\f(t,x{t))\dt\^\
E'-[\f(t,x{t))\,ro>t]dt T
pPsO,
xl
f\f(t,x(t))\dt
Also, if {PS^} is a r.c.p.d. of Pj^.^n-iAT' then by Lemma 9.1.5
r^'
\f(ux(t))\dt]
T2n-1AT
1 ,p(«)
<£'
£'-'[|/(t,x(t))|,T2„>f]dt,T2„-,
L-'T2n-l <
£ ' ' l£J*T^2n-l(-), ^ ( t 2 n - l ( - ) )
j
l/(f.^(0)Mf
t2„-i(-)<7^
t2„-l(-)
since S^(,^^_ ^^j ^^ (8)?2n- i(co) ^S^^ solves the martingale problem for L, starting from (T2„- i((o), X(T2„- i(a)), co)) for P-almost all co satisfying T2„- i((o) < T. Hence f f(t. x(t)) dt
<;
+
£Ps0.xl
sup
I
|/(t,^(t))|dt E Ps.x j
\f(t,4t))\dt lP(r2n-i
sO <s
and so Lemma 9.1.6 can be used to complete the proof. G 9.L8 Lemma. Let a: [0, oo) x R^ -^S^be a hounded measurable function, (5^, x^) e [0, 00) X R'^, T > 0, and R>0 such that the function ja(s,x) if s ^ < 5 < r ''^''''^"la(5^x^) otherwise
and
|x-x«|
220
9. Some Estimates on the Transition Probability Functions
is an element of s/(r, 2, A) for some 1 < r < oo and 0 < >i < A. Let P solve the martingale problem for
starting at (5^, x^). Then for each 0 <<x < R and r < p < 00
(1.32)
f
{t-smt,x(t))dt
32IM ||LP([SO, r]xi?d)
for allfe Co([s^, T] x B(x°, a)), where N = [(d + 2)/2r] and C1.32 depends only on d, r, r, p, R — cc, A, and A. Also, if 0 < a < R, r < p < 00, and \x^ — x^\ > a, then
I /(r, x(t)) dt
(1.33)
33 IM ||LP([so,r]xi?d)
for all fe Co([s°, T] x B(x^, a)), vv/iere Ci 33 depends only on d, T, r, p, R — a, Ix^ — x° I — a, A, anti A. Proof Let {Ps,^' (^^ x) e [0, 00) x R^] be the measurable strong Markov process determined by 1
^ i,j=l
d^ dxi dxj
Taking S = (R - a)/2 in Lemma 9.L7, we see that:
I (^(f, x(r)) ^r
< £^^0.^
f k(^x(r))Mr , 1
+ C1.31
sup sO<s
E^
J |(/)(r, x(f))| ^r
R-d
for all (p E Co([s°, T] x B(x°, a)). We can therefore apply Lemma 9.L4 to complete the proof. D We have at last arrived at the main resuh of this section. 9.L9 Theorem. Let a: [0, 00) x R"^ ^ S^ and b: [0, 00) x R'^-^ R^ be bounded measurable functions. For each T > 0, assume that there exist numbers
221
9.1. The Inhomogeneous Case
0 < AT- < AT- < 00 and Bj < co and a non-decreasing function Sj-: (0, oo) -• (0, oo) such that:
Xr\e\'<(e,a(s,x)e)
6 e R^
r]xR^
and sup
||a(s, x^) - a(s, x^)\ < e, 8 > 0.
0<s
\x^-xH
Set
and let P(s, x; t, •) be the transition probability function determined by L,. Then P(s, x; f, •) admits a density in the sense that there exists a non-negative measurable function p(s, x; t, y) on {(s, x; t, y) e ([0, oo) x R^)^: s < t} such that for each (s, x) € [0, oo) X R'^ (1.34)
P(s, x; r, r ) = I p(s, x; r, y) dy,
for almost every t > s. Moreover, ifO<s< (1.35)
i\\t-sfdt\
T
G
^j^a,
T and 1 < q < co, then
\p(s,x;t,y)\Uy]
where (x = (d -\- 2)(q — l)/2 and Cj 35 depends only on d, T, q, Aj, A^, Bj, and dj('). Finally, ifO<s< T and 1 < q < 00, then for each d> 0 (1.36)
if Vs
dt\
\p(s,x;t,y)\'^dy)" ''R'i\B{x,d)
where Cj 35 depend only on d, T, q, kj, Kj, Bj, 6j('), and d. Proof We will first assume that fe = 0. Given 1 < ^ < 00, let (j' be the conjugate of q and choose 1 < r < ^' so that N =
^+21] d + ll = [ 2 r
222
9. Some Estimates on the Transition Probability Functions
Let £ be defined by 1
C,{r, kj, A r ) v C J | - y
v r j + 1, Xj, A^j
Then e satisfies (1.14). Thus, by Lemma 9.1.8, for any (s^ x^) e [0, T] x R'^ and any x e R^. (1.37)
I (t - sf dt I
P{s, x; r, dy)f{t, y)
Lf>([s, T) X Rd)
B(xO, , )
where 7 = iSj(8), p = (r + q')/!, and Ci 37 depends only on d, T, r, p, A^^, and A^^ In particular, if
for /c = (/ci, ...,/Cd)GZ^ then (1.38)
f (f-5)^^r j s
P(s,x;r,^>;)/(f,y)
^
^ 1 . 3 7 | | / | | L P ( [ 0 , r]x/?d)
Qk
On the other hand, by (2.1) in Chapter 4, we know that P(s, x; f, Qk) <2d exp
\k\y Id^AjT
for s < f < 7; and therefore 2„.2
(1.39)
I (f-s)^^r j
P(5,x;f,^y)/(r,>;) < C i . 3 9 e x p
^l^y Id^'KjT
where Cj 39 depends only on f/, T, and N. Combining (1.38) and (1.39), we know by interpolation that 1.40)
f (t-sf
dt f
P(s,x;udy)f{uy)\ 2„2
<(Ci.38r-'(Ci.39rexp where 0 < 0 < 1 is defined by the equation
\
_\-e
2^2A^r
L9'([0. T] X Rd)
223
9.1. The Inhomogeneous Case Summing over k e Z^ in (1.40), we arrive at 1.41)
[ (t-sfdt
I
P(s,x',udy)f(uy)
^ ^ 1 . 4 1 ||/|iL'J'([s, r]x/?d)
where Cj 41 is determined by d, 6, Ci 38, Cj 39, y and A^ in the obvious way. We now remove the assumption that b = 0. Let
be the second order part of L, and define {P^^/. (^^ x) e [0, 00) x R"^} and P^'is, x;t, •) accordingly. By (1.41), if 1 < ^ < 00 and N = [(J + 2)/2^'], then
(1.42)
\ (t-sf
dt I
P''(s,x;t,dy)f{t,y) ^ ^1.41 11/II L4'([s, r]xRd)
for all 0 < s < 7 and X e R^. Note that P(s,x;r, r ) = £^-1X^(7), ^ ( O e r ] for 0 < s < r < T and T e M^i, where
X\T)
= exp
^{b{u x(0) ^^(0> - ^ r<M^ ^(^)).«" H^ A^mu At))) dt
As we have seen before:
1.43)
EPs.^\(X'(T)f]
< exp A j^
c..4 3
Thus, if 1 < q < 00 and N = [(i/ + 2)/2g')], then
ll (t - 5)"'^ dt jj(s, X- t, dy)f(t, y) = E ' ' - ' \\t-sr'f(t,x(t))dtX%T)
<
Ty -- 5)*C!.,3( ^''^-^ \\\t
<{T—S)
- sf \f(U x(t)) 1^ dtV^
C1.43C1.41 ll 1/ I ||/>/2([s, r]xRd):
224
9. Some Estimates on the Transition Probability Functions
where the Cj 41 here is the one in (1.42) when ci is replaced by ((II. We have therefore shown that if 1 < ^ < 00 and iV = [(c/ + 2)/2^')], then (1.44)
f (f - sfi^ dt [ P{s, x; t, dy)f(t, y)
4 4 IM l|L«'([s, T]xRd)
•'R'
where Cj 44 depends only on d, T, q, kj, \j, Bj, and ^7(-)We next show that p(s, x\ U y) exists. For n > 1, define /i" ^ on ([0, 00) x K**, ^ ,[ 0 , oo)xi?dJ by < . ( [ ^ 1 , ^2] X r ) = j
;cui,*2](0^(s, x; t, r ) ^r.
's + ( l / n )
Clearly (s, x) -• //" ^ is measurable; and, by (1.44), /x" ^ <^ dt x dy. Thus, by standard results, there is a measurable non-negative function (p„ on {(5, x; r, y) e ([0, 00) X R'^f: s
t/r X fiy
(t,y) = (p„(s,x; t,y).
We define /?(s, x; r, y) =
h\t
- sf^" dt \jp(s,
x; r, y)|^ dyV' < C, .44-
Since Nq_l
T"~2
^ + 22 2
J + 2/^
1\
d-\-2.
^,
^'
(1.35) follows. The proof of (1.36) can be accomplished in either one of two ways. One can proceed in exactly the same manner as we just have in the derivation of (1.35) from (1.32), only this time using (1.33); or one can derive (1.36) directly from (1.35) in the same way as we got (1.24) from (1.22) in the proof of Lemma 9.1.4. The details are left to the reader. D 9.1.10 Corollary. Let a: [0, 00) x R'^-^ S^ and b: [0, 00) x R'^-^ R'^ be bounded measurable functions. For each T > 0 and R> 0, assume that there exist numbers ^ < ^T,R^ ^T,R < 00 and Bj R < 00 and a non-decreasing function Sj^:
9.1. The Inhomogeneous Case
225
(0, oo) ^ (0, 00) such that
(s, x) e [0, T] X B(0, R) Ib(s, x)\
and
6 e R'^,
(5, x) 6 [0, T] X B(0, JR),
and IIa(s, x^) — a(s, x^)\\ < £, s > 0.
sup
sup
0<s
xi,x2eBi0, R) \xi-x2\
Set
and let P(s, x; r, ') be the transition probability function determined by L^. Then P(s, x;t,') admits a density p(s, x;t,y) in the sense described in Theorem 9.1.9. Moreover, ifO
( [ (t-sYdt
f
\p(s,x;t,y)\'^dy]
'
where a = ((^ + 2)l2)(q — 1) and Cj 45 depends only on d, T,q, R — r, X^, R ^ ^T, R » and ST,R('). Finally, ifO
BT,R,
(1.46)
Ifdt
f
\p(s,x;t,y)\'^dy]
where CI 46 depends only on d, T,q,R — r,XjR, Aj,
'
R.BTR,
and ST,R(')as well as 3.
Proof Given T > 0 and 0 < r < R, choose measurable a: [0, 00) x /?'*-» 5 / and B: [0, 00) X i?** -• R'^ so that a = a and 5 = 6 on [0, T] x 3(0, r) and a and b satisfy the conditions of Theorem 9.1.9 with A^ = ^T,R^ ^T = ^T,R^ ^T = ^T,R and 6j(') = ^J |j(•). Set
and define P{s, x; t, •) accordingly. By Lemma 9.1.7, IT
\\
I
dtl
T
P(s, x; t, dy)f(t, y)\ < { dt i P(s, x; t, dy)\ f(t, y)\ 'B(0, r) T
Ci.31 sup I s
dt
jP(u,z;t,dy)\f(t,y)\
226
9. Some Estimates on the Transition Probability Functions
where C1.31 depends only on d, T, R - r, A^^, ^ and Bj, R . Thus (1.45) and (1.46) follow easily from Theorem 9.1.9. G One of the applications of Corollary 9.1.10 is to the connection between the diffusion process associated with L, and the known analytic facts about solutions to the backwards equation. To be precise, we give the next theorem. 9.1.11 Theorem. Let a: [0, 00) x R'^-^ S^ and b: [0, 00) x R'^ ^ R'^ satisfy the conditions of Corollary 9.1.10 and let {Ps,x'- (s, x) e [0, 00) x R'^} and P{s, x; f, •) be defined accordingly. Suppose that {f„}f ^ C^' ^([5, T] x R*^) has the properties that fn-* (p e Ci,([s, T] X R'^) boundedly and pointwise in [s, T] x R'^ while
dt
+ LJ„-^iAeB([5, 7] xR')
in Lfoc,((s, T) X R'^) for some i < p < 00. Then, for each x e R"^, (p(tAT, x(t A T)) - jl^^ \l/(u, x(u))du is a P^^^-martingale after time s. In particular,
(1.47)
(p(s, x) = I P(5, x; r , dy)q>{T, y)-\
dt [ P(s, x; t, dy)il/(t, y).
Proof. Clearly (1.47) will follow if we prove the first assertion. To do this, let s < t^ < t2 < T be given and define XR = mf{t> t^: \x(t)\ >K},
R>0.
Then for any A e M^^'. E^'iUh
A TR , X(t2 A T«)) -f„(ti
A XR , X(ti A T^)), A]
^^\[
, [j^+L,y^(t^x(t))dt,A\.
As « - • 00, E''ifn(t
A T« , X{t A T^)), A] -> £^-[(/>(r A T«)), A].
and by (1.45)
< I
^^ J
I ^ + LAf„(t, y)\ - il/(t, y)p(s,
x;t,y)dy^O
9.1. The Inhomogeneous Case
227
where p(s, x; r, y) is the density of P(s, x; r, •). Thus E^'i(p{t2
A T^ , X(f 2 A T;^)) - (/)(f 1 A T^ , x(t^ A T^)), >l] .'2
= E' Letting R/co, Fj^-martingale.
we D
see
that
J
ATR
iA(t, x(f)) dt, A
(p(t, x{t)) — J^ i/^(w, X(M)) ^W
is
indeed
a
Before one can make use of Theorem 9.1.11 it is necessary to have existence theorems from analysis. Fortunately, such existence theorems are often known. For instance, if a is uniformly positive definite and uniformly continuous, then Fabes and Riviere [1966] have shown that each ip e Co([0, T] x R^) is the limit in every L^([0, T] X R^), for 1 < ^ < oo, of a sequence /„ given by (//„ = (dfjdt) + L,/„, where/„(•, T) = 0,/„ e Cl' ^([0, T] x K'^) and the functions/„ converge uniformly to a hmit/. With a little more effort one can extend their results to mesh with the more general set-up considered in Theorem 9.1.11. There is one more type of estimate about p(s, x; r, y) which we want. The origin of our interest in this estimate will not become apparent until Chapter 11. Given 0 < A < A, recall that s^(d + 3, A, A) stands for the set of measurable functions a: [0, oo) x R'^ -^ S^ such that A\e\^ <(e,a(s,x)0)
(S,X)E[0,OO)XR'
and
9 e R',
and, for some x^ e R"^,
|a(s,x) - a(s)\\ < -^Cdid + 3,A,A),
(5,x) G [0,oo) x R'
where a(-) = a(', x^). Given a e ^(d -f- 3, X, A) and T > 0, define D j ^ as in (1.13), and recall that (/ - Z)Ja)~ ^ exists as an operator with norm at most 2 on L^^^([0, T] X /?'') into itself 9.1.12 Lemma. Given a e ^(d + 3, i. A), define Kj on L^^^([0, T] x R'^) by (7.25). Then for fe I^^'([0, T] x R'), Klfe C,([0, T] x R% and, in fact,
(1-48)
| | ^ J / | | < C 1.48
| | / | | L ' ' + 3([0,
T]xRd)
and (1.49)
\Klf{s,x')-Klf{s,x')\
228
9. Some Estimates on the Transition Probability Functions
where the constants C1.48 and Cj 49 depend only on d, 7, A, and A. Moreover, if h E R^ and a'(s, x) = a(s, x + h), (s, x) e [0, 00) x R*^, r/ien a' e s/(d -I- 3, A, A) aw^ (1.50) II KU-
Klf II < Ci.50
sup
||a(5, x + h)- a(s, x)\\ ||/|U..3(«^
(S, X ) 6 [ 0 , QO)x/?d
w/i^re Ci 50 depends only on d, T, A, an^ A. Proo/. Clearly (1.48) is an immediate consequence of (1.9) (with r^ = t/ + 3 and r2 = 00) and the boundedness of (/ — Dj^^)"^ Moreover, by the same sort of reasoning as led to Lemma 9.1.1, it is easy to see that (dGjf)/dXi e Ci,([0, T] x R*^), l
dGjf
(1.51)
dXi
Ld + 3([0,
T]xRd)i
where Cj 51 depends only on d, T, A, and A. Again, because (/ —Z)J^^)~^ is bounded on I?"'^([0, T] x R**), this proves (1.49). Finally, to prove (1.50), note that we can use the same a in defining Kj, as we use for Kj-. Since \\Dl,f- /)J../|k.^3ao,nxK.)
+ 3, A, A)||/||,..3«o,nxi,^
where rj = sup(s^)g[o, oo)xRd ||«(5, x H- /i) — a(s, x)|| and because l i ^ f l , f l / | | L ' ' + 3([0, r]x/?d) V l l ^ a - . a / l k d + sao, 7] x/?rf) ^ l H / | | L < ' + 3 ( [ 0 , T] x/?«*)»
we see that \\(Dl,rf- (/)J,.)y|U..3ao.nxi,.)
K A)||/||L.^3([o,nxKa).
Thus (1.52)
11(7 - Dl-XH-
{I - ^J'..ry||L-3«o, nx,,.) <
CI.52^||/||L-3«O,
nxi,-),
where Cj 52 depends only on d, A, and A. Using (1.9) once more, we now get (1.5) from (1.52). Q 9.1.13 Lemma. Let ae s/(d -\- 3, A, A) and denote by P(s, x; f, *) the transition probability function determined by the martingale problem for 1
'^
2 ij^i
d^ '
dXi dXj'
229
9.1. The Inhomogeneous Case
Given T>0 andfe C?([0, T) x R% defineff,(s, x) =/(s, x - h). Then
f dt\
(1.52)
(f(t,y)~Mt,y))P(s,x;t,dy)
sup {t,y)e[0,
\\a(t, y) - a(u y +
Ld + 3([0,
T)xRd)
oo)x/ld
Proof. Set a'(s, x) = a(s, x + h), (s, x) e [0, oo) x R^, and let F(s, x; t, •) be the associated transition probability function. It is easy to check that f (PH(y)P(s, x-\-h;t,dy)=\ -'Rd
(p(y)F(s, x; r, dy)
JRd
for (p e Ct(R'^). Thus: \C dt\
(Ut.y)-f(t,y))P(s,x;t,dy) < f dt \ f,(t, y){P(s, x; t, dy) - P(s, x-\-h; t, dy)) + f dt\ f(u y)(F(s, x; r, dy) - P(s, x; t, dy)) = I KjMs, x + h)- KlU(s, x) I + \Klf(s, x) - Klf(s, x)\
where fy = sup(,,,),[o, oo).Rd \\a{U y't-h)-
a(u y)\\. D
9.1.14 Lemma. Let a: [0, co) x R'^ -^ S^ be a hounded measurablefiinction,and suppose that for some (s^, x^) € [0, oo) x R*^, T > s^, and R > 0, r/iere exists anae j^(d + 3, A, A) 5Mc/i r/iar a = a on [s^, T) x 5(x°, R). Let P be a solution to the martingale problem for
starting from (s°, x°). Then for any 0
\E'-\f(f,(t,x{t))-f{t,x{t)))dt < C,J
\h\ + \
sup
\\a(t, y + h)- a{t, y)\\ ]\\f\\,
(t,y)e[0,ao)xRd
where Cj 54 depends only on d, T, A, A, R, and r.
J
230
9. Some Estimates on the Transition Probability Functions
Proof. Define T„, « > — 1, as in Lemma 9.1.6 with R^ = r and R2 = R. By the famihar reasoning, if g e CQ([S^, T] X B(X^, r)), then: f g{h x(t)) (it <
t\E'\r^\{t,x{t))dt,X2„-,
rT2„-i
J
i / Mr
< >
/
g{t,x(t))dt P(d(o)
\E'^\ I
T2n-1 (C0)
where P^ = Px2„-i{co),xir2n-i{co),(o) and Ps,x solves the martingale problem for
2 , ; ^ ! ^' ^ax,^x,. starting from (s, x) and T?„ = inf{f>T2„_i(co): | x ( r ) - x « | > R}. But if T = inf{f > 5: |x(f) - x^ | > R}, then P(s, x; r, r ) = P,, ^(x(r) e r, T > r)
and so
+£^- U
dtj^^ \g(t,y)\P(s,x{T);tJy), T < T
Taking g =fh —/and applying Lemma 9.1.13, we arrive at \\UUx(t))-f(t,x(t))dt
I
X 2C1.53 |^| + sup||a(r,}; + /i)-a(r,>;)|| 11/11 \
(f. y)
We can now use Lemma 9.1.6 and the obvious estimate of terms of r and ||/|| to get (1.54). D
'
L''+3([0, T]^R
m
231
9.1. The Inhomogeneous Case
9.1.15 Theorem. Let a: [0, oo) x R^-* S^ and b: [0, oo) x R'^-^ R'^ be bounded measurable functions. For each T > 0, assume that there exist numbers 0
<\T\e\\{s.x)e[0,T\xR'
\b(s,x)\
and 6 e R\
T] x R\
and sup
||a(s, x^) — a(s, x^)\\ < £, £ > 0.
o<s
Ut
L. =
lia%.)^^^ib'it,.)±
and define P(s, x; t, •) and p(s, x; r, y) accordingly as in Theorem 9.1.9. Then for each T > 0 there is a non-decreasing function O: (0, oo) -• (0, co) depending only on d, T, Aj, A J and dj(*) such that linig^o ^(^) = ^ ^^d T
(1.55)
j
dtj\p(s,x;t,y
+
h)-p(s,x;t,y)\dy<mh\).
Proof. Assume that we have proved (1.55) in the case when b = 0. Set
and let {P^,x' (^^ ^) e [0, oo) x R''}, be the associated family of solutions to the martingale problem. Then our assumption implies that E^'^ j
(f,(t,x(t))-f(t.x(t))dt
<0(|/i
for/e C?([0, T) x K'^). By Theorem 7.2.2, ^Ps.,
j(Mt,x{t))-f{t,xmdt ^P9
f{Mt,x{t))-f(t,x{t)))dtX%T)\\
P9. \^\\f,(t, x{t)) -f(t. < E^^l
x(t)))
dtJYEni(X%T)n
232
9. Some Estimates on the Transition Probability Functions
where P^ ^ solves for L, starting from (s, x) and X'(T) = exp j (a-'b(x(u)ldx(u))--\
1 "^
(b,a-'b}(u,x(u))du
Since, as we have already seen, E^^'^[(X^(T))^] is bounded by exp[Bj{T — s)/2XT]y it remains only to estimate
En. {^j\Ut,x(t))-f(t,x(t)))dt But (\\Ut,x(t))-f(Ux(t)))dt^ (f^(u,x{u))-f(u,x(u)))du\ ''s
(fh(v,x(v))-f(v,x(v)))dv •'M
= 2En. j V.(«, x{u)) -f(u, x{u))) du E'S-JfiMv,
<2\\fmh\)En
j
x{v)) -fiv, x{v))) dv
\Mu,x(u))-f(u,x{u))\du
<4||/||^(r-5)(|;.|) Thus. £•''•'
J (Mt,x(t))-f{t,x{t)))dt <2exp|-^(r-s)
ll/ll(r-s)5((l'i|))'.
It is easy to see from the preceding that (1.54) now holds with 273 exp[B|.T/2Ar]*(|/i|)5 replacing
9.2. The Homogeneous Case
233
where C1.54 depends only on d, T, Xj, Aj, and p. By the estimate in (2.1) of Chapter 4, we can now find A depending on d, p, and Cj 54 and a > 0 depending on d, AT^ and p such that for a n y / e ^([0, T] x R'^) and \h\ < p:
\
(Mux(t))-f(t,x(t)))dt
for all n > 1. Since there is a C depending on d, A, and a such that infA(n'(\h\ H-M/i)) + ^-"")
and (1.55) follows.
D
9.2. The Homogeneous Case In this section we will show how to refine the results of the preceding section in the time homogeneous case. To be precise, let a: i^'' -• Sj" be a bounded continuous function, b: R^ -^ R^ 2i bounded measurable function, and I d
32
d
^
The martingale problem for L is well-posed. Let {P^ ^\ (s, x) e [0, 00) x R"^} be the associated family of solutions. At the heart of the aforementioned refinements is the observation made in the next lemma. 9.2.1 Lemma. For s > 0 define O^: Q -• Q 50 that x(t, O^(co)) = x((t - s) vO, co), t > 0. Then P^^ x = Po,x° ^s^-1^ particular, ifP(s, x; t, -) is the transition probabilityfi^nctionassociated with L, then P(s, x; r, •) = P(0, x; t — s, • )for 0 < s < t. Proof. The first assertion is immediate from the fact that FQ, x ° ^7 ^ solves the martingale problem for L starting from (5, x). To prove the second part, note that from P^^ = P Q ^ o 07 ^ we have P(s, x; r, r ) = P,,,(x(r) e F) = Po,;(x(r - s) G T) = P(0, x; r - s, T). D In view of the preceding lemma, it is natural to adopt the following conventions when dealing with time-independent L's. In the first place, we don't really need Pj ^ for any s other than s = 0; and so we use P^ to denote PQ, ^ and abandon the special notation for solutions starting after time 0. Secondly, we define P(r,x,) = P(0,x;t,)
234
9. Some Estimates on the Transition Probability Functions
and call P(t, x, •) the {time homogeneous) transition probability function associated with L. Observe that the Chapman-Kohnogorov equation now reads: (2.1)
P(t + s, X, r ) = j P(r, y, r)P(5, X, dy).
The fact that the transition probability function possesses a special form in the time homogeneous case enables us to refine the results of the preceding section. The next lemma is typical of the sort of improvement that we can expect to make. 9.2.2 Lemma. For each T e ^^d, (s, x) -• P(s, x, F) is continuous on (0, oo) x R^. In particular, for all (s, x) G (0, oo) x R*^, P(s, x, •) possesses a density p(s, x, y) with respect to Lebesgue measure. Proof Let s > 0 be given and let (s„, x„) -• (s, x) with 0 < s„<2s for all n. Then P(5„,x„,r) = P ( 2 s - ( 2 5 - 5 „ ) , x „ , r ) = P ( 2 s - 5 „ , x„;2s, r ) ^ F ( s , x; 25, r ) = P ( 5 , x , r ) , where we have used the strong Feller property. Finally, we know from Corollary 9.1.10 that P(s, X, •) = P(0, x; s, ) has a density for almost every s > 0. In particular, if r has Lebesgue measure zero, then P{s,x,r) = 0 for almost every 5 > 0. Thus, by the preceding, P(s, x, F) = 0 for all s > 0, and so the proof is complete. D
We now want to obtain L^-estimates on p(s, x, •) for 1 < ^ < oo. Unfortunately, this will require our proceeding, as in Section 9.1, from the case when ^ = 0 and a is a small perturbation. However, much of the procedure is essentially identical to the one in Section 9.1, and therefore at times we will not bother again with every detail but, instead, will simply refer to the appropriate part of Section 9.1. Since we are now dealing with time independent coefficients, nothing will be lost by restricting our attention to perturbations off of constant c e 5 / . Thus, we introduce the notation: (2.2)
g^^^s, x) = ^^2nsr^c\)^exp[-Kx,
c"^x>/25]
for c E Sj, s> 0, and x e P**. Obviously, if c(-) = c, then g^'^(t - s, y - x) = g^%, x; t, y)
235
9.2. The Homogeneous Case
where the right-hand side is given by (1.3). In particular, i f O < A < A < o o are the lower and upper bounds on the eigenvalues of c, and if G, /(5, x) =\
dt\
g''\t -s,y-
x)f(t, y) dy,
0 < s < 1,
then for 1 < ri < r2 < 00 satisfying l/r^ - 1/^2 < 2/(d + 2): P-^)
\\Gcf\\u2{[0,l]xR'i) ^ C2M\f\\ \\ui{[0,l]xRd)^
where C2.3 depends only on d, r^, Vj, A, and A. Next set (c)
9\Ht,y)=^^-f^(t,y),
i
Then an easy computation yields: i/q
((d'ij0\1(t,y)\'dy^
is finite for 1 < ^ < (^ -I- 2)/{d 4- 1) and is bounded by a constant depending only on d, q. A, and A. Hence, by Young's inequality,
dGJ (s,x)
(2.4)
dX;
< C 2.4
IM llM[0,
l]xRd)
for d + 2
xeR'
and 0 e R'
and, for some x^ e R^, sup ||a(x) -a||^^Q(p,>l,A) < 1,
r < p < (i/ + 3) Vr
where a = a(x^) and Cd(p, A, A) is the constant given in (0.4) of the appendix. For a e s^oir, A, A) we define
and K, = G , o ( / - D „ . , )
236
9. Some Estimates on the Transition Probability Functions
on If([0, 1] X R^) for r < p <(d -\- 3)vr. The comments about Kj as defined in (1.15) apply equally well to K^ defined above. With these preliminaries, we are ready to begin. 9.2.3 Lemma. Let 0 < X < A and 1 < q < co be given. If a e J^O(^^ r — c(, then the associated density function p(s, x, y) satisfies: (2.6)
K
A), y^here
ij s^ ds j Jp(s, X, y)\^ dy\ '
and (2.7)
{[ s'ds\
\p(s,x\y)-p(s,x\y)\^dy\
' < C2.7IX'- x^ |
for some a > 0 depending only on d and q and some C2.6 cind C2.7 depending only on d, q, X, and A. Proof Set N^[(d + 2)/2r]. By Lemma 9.1.1, X^"" ^ e Q([0, 1] x R^) with (2.8)
\KV'f{s.x)\
where C2.8 depends on d, r, I and A. Also, since K^ is continuous on ir([0, l\x R^) into itself with a bound of 2, II ^fl "^
/||L'-([O,
i]x/?d) < 2
ll/llLrao, i]x/?d)
Thus if p = (^ -H 3) V r, then (2.9)
IIXa
/ | | L P ( [ 0 , l]x/Jd) < C 2 . 9 | | / | | L r ( [ 0 , IJxK'')'
where C2.9 depends only on d, r, A, and A. Since Ka has a bound of 2 on L^([0,1] X R^) into itself, we conclude that
^Xf
^2C2.4||-K^a
/ | | L P ( [ 0 , l]x/?<0
< 2C2.4 ^2.9 |i/||Lr([0, 1] x/?d),
and so (2.10)
1^^2/(0, x') - x r y(0, x^)| < C,.,o|x' - x^ I ||/|U,ao, ILR^
237
9.2. The Homogeneous Case
Using Lemma 9.1.2, we have from (2.8): \\\'^'ds\
p(s.x,y)f(s,y)dy\ i?
fs'^'dsj 0
=
p(s,x.y)\f{s,y)\dy "Rd
N\(Kr'\fmx)
and from (2.10): [ s^^^ ds [ (p(s, x\ y) - p(s, x^ y))f(s, y) dy
< (jv +1)! |x' - s' I | x r m x')- x r y(o, x^)i <(N+1)!|X'-X^|C,..O||/||L^[O,.,X«.,.
From these it is easy to get (2.6) and (2.7) with oi = (N -\- l)q. [J We are now going to see how estimates like those in Lemma 9.2.3 can be used to yield /^-estimates on p(s, x, •). The idea is familiar to experts in the theory of partial differential equations and comes under the general heading of " Sobolev inequalities "; although what we need is, more precisely, a " Morrey inequahty." In any case, the general principle is that uniform estimates can be derived from mean estimates plus sufficiently good estimates on the modulus of continuity. Our next theorem makes these comments precise. 9.2.4 Theorem. Let (£, J^, P) be a probability space and F: R^ x R" x E-^ R^ a measurable function. Assume that (i) for each q e E and (p e Co(R% x -• L^(x, q) is continuous, where L
< B\x'-
x''\'-'\
Then for each P > (d -h a -\- l)/q there is an M = M(d, q, cc, p) such that
for all x^ e R'.
238
9. Some Estimates on the Transition Probability Functions
Proof. It suffices to handle the case when x^ = 0. Let rj = 2d -\- a/2. Then H=\
\x' -x^l'^''-'^ dx,dx2
f •'B(O, 1) V o ,
1)
Hence, for each R>0:
[f
1 (
1''B(0,R)
*'B(0,
M _ F^Y^^II \\nx')-F(x
\Q
J dx' dx'
< BHR'^^''^\
i?)\
where i/«
\\F(x') - F(x')l = (f^^ \F(x\ y) - F(x\ y)\^ dy^ From this we know that
p(3
f
*^B(0, R)
*'B(0,
R)
X -^^^^^0—h^^dx' dx' > A||
Applying Exercise 2.4.1, we can now assert that pUp e Co(R")):
sup
x l . A : 2 e B ( 0 , B)
L^[x')-L,(x') \x'-x^f
> AX^11^11.)
A where ^ = ajlq and
and y = 7d is explained in Exercise 2.4.1. In particular, (2.11)
P\\
sup
\ x e B ( 0 , R)
for all K > 0 and /I > 0.
F(x)-F(0)|L
, , , , , \ ^ B H K•d + a/2
>Ak'i''\<
9.2. The Homogeneous Case
239
Taking R = N > i and choosing A so that
we see that (2.11) becomes: W sup
»:i?^^LN-')<(^rBi/N-'-^
\xeB{0,N)
1^1
/
\-^/
and so pl
sup
l | f W - m < ^ ) < ^ ^ . . a-qp
\ x e B ( 0 , N)\B(0, N - 1 )
( 1 + |^|)
/
^
Since ^ f iV'^^""^^ < oo, the theorem is immediate from this.
D
9.2.5 Lemma. Let d < q < co and set r = q'. Given a e s/o{r, A, A), let p(s, x, y) be the associated density. Then for all 1 < p < q, p(s, x, •) e L{R'^) for all (s, x) e [0, oo) X R'^. In fact, (2.12)
||p(5,x, • ) I I L . ( / . ^ ) < C 2 . I 2 ( 5 A 1 ) - \
where C2.12 depends only on d, q, p, X, and A and v > 0 depends only on d, p, and q. Proof In Theorem 9.2.4, take E = (0, 1], P(ds) = ^7(1 + a) ^5 with the a in Lemma 9.2.3, and F(x, y, s) — p(s, x, y). By (2.7), we have:
- - ^ j ' s " ^5 j |p(5, x \ y) - p(5, x^ y)r ^y <j^CU^'
- x^l-^
and therefore, by Theorem 9.2.4, for any p> I -\- \/q:
where C2.13 depends only on ^, €2.7/(1 + a), g, and j5. On the other hand, from (2.6) we also know that: (2.14)
P(s:||p(s,x», •)||,>L)
where C j . ^ = C^ g/(l + a). Thus, if 0 < ff < 1 is given and we take L = yla\ where y = 4(a + l)C2.i3 C2.14 and v = a/q, then p | s e (0, c;): ||p(s, x«
> yo ' or
sup J^%^4;^'^:^V^>7'^-)^w,''))-
240
9. Some Estimates on the Transition Probability Functions
Thus there is an s e (0, G) such that
Using the continuous version of Minkowski's inequahty, we can write: «\l/4
||p(a, x°, • )||^ = I f dy\\ p((T - s, x^, x)p(s, x, y) dx\ < I p((7 - s, x^ x)\\p(s, X, •)||^ dx
is bounded by a constant depending only on d, P and A. We have therefore shown that (2.15)
\\p(s, X,
1] xR^
where C2.15 depends only on d, q, A, and A. If 1 < p < ^, define 9 by 1/p = (1 - 0) + (O/q). Then, since ||p(s, x, ')\\mRd) = 1, (2.16)
\\p(s, X, Olkpd..) < (C2.l5r5-^^ (5, x) G (0, 1] X R'.
Finally, if s > 1, then \\p{s, A '^Lm") ^ f Pi^- 1' ^''' ^)IIP(1' ^' ')L dx < (C2.27)', and so the proof is complete.
Q
9.2.6 Theorem. Let a: R^ ^ S/ he a bounded, uniformly continuous function with 3: (0, 00) -• (0, co) as a modulus of continuity and assume that X\e\^<(e,a(x)e)
XER'
and
OeR',
for some 0 < A < A. Let h\ R^ -^ R^ he a hounded measurable function such that \h(x)\
241
9.2. The Homogeneous Case
and let p(s, x, y) denote the density of the transition probability function determined by L. Then for each 1 < q < oo (2.17)
(f^^ \p{s, X, y)!' dyj"
< C2.„(sA 1)-^
where v depends on d and q alone while C2.ii depends on d, q, <5(*), A, A, and B. Moreover, 1/ a > 0, then (2.18)
([
\p(s,x,y)\^dy\
"
where C2.18 depends only on d, q, a, ^(•), A, A, and B. Proof Because sup (f |p(5,x,y)M
\Te^ Rd,
is a non-increasing function of s > 0, we need only consider 0 < 5 < 1. Moreover, the Cameron-Martin formula can be used in the same way as we did in the proof of Theorem 9.1.9 to get the general resuh from the case in which fe = 0. Finally, the partitioning technique used in the proof of Theorem 9.1.9 can be employed again here to reduce the problem to that of showing that there is a 7 > 0 depending on d, (5( •), q, A, and A such that (2.19)
(f
\p(s.x,y)Ydy\"
for all (s, x) e (0, 1] x R^ and x^ e R^, where v depends only on d and q while C2.19 depends on d, q, X, and A. But (2.19) can be proved from Lemma 9.2.5 by the same localization technique as was used to prove Lemma 9.1.8 from Lemmas 9.1.4, 9.1.5, and 9.1.6. All that is needed is the observation that if 5 satisfies the hypothesis of Lemma 9.2.5 and p(s, x, y) is the associated density, then for any a > 0 and 1
i f
UP
\p(s, X, y)\'^ dy] [
'^R<'\B(x, a)
and so
M \p(s, x, y)\^ dy
p(s,x,y)dy\ 1
[''R''
< (2d exp[-aV2^sA])^(C2.i2S~')^"^
(2.20) (f
\p(s, X, y)\'dy)
VR''\B(x,a)
' I
(l-9)lq
242
9. Some Estimates on the Transition Probability Functions
where C2.20 depends on a as well as d, q, p, X, and A. With these remarks, the reader can complete the proof of (2.17). As for (2.18), it can be easily derived from (2.17) by the same argument as we just used to get (2.20) from (2.12). D 9.2.7 Corollary. Let a: R^ ^ S^ be a bounded continuous function and b: R^ -^ R^ a bounded measurable function. For each R> 0, let the numbers 0 < A/j < A^ and BR satisfy XR\9\
<(e, a(x)e)
XE B(0, R)
and
0 e R\
and \b{x)\
and let SR(*) be a modulus of continuity for a restricted to B(0, R). Then for each 0 < r < R and I
(2.21)
(f |p(5, X, y)|^ dy\" VB(0,r) J
<
C2.2,(sAi)-\
and for a > 0
(2.22)
(f
\p{s.x,y)\'dyY
In (2.21) and (2.22), v depends only on d and q, C2.21 depends on d, q, R — r, ^ji(*), XR, A|J, and B^, and C2.22 depends on these as well as a. Proof The derivation of this result from Theorem 9.2.6 is the same as that of Corollary 9.1.10 from Theorem 9.1.9. We leave the details to the reader. D We will close this section with the analogue for the present context of Theorem 9.1.15. 9.2.8 ^af^
Lemma. Let a e s^^id + 3, >l. A). Q([0, 1] X R^) and, in fact:
Then
for
fe L^"'^([0, 1) x R%
and
(2.24)
\KJ{s,x')-KJ(s,x^)\
< C , . , , | x ' - x^| ||/||,,.,„o, x.x«.,
9.2. The Homogeneous Case
243
where C2.23 ^^^ ^2.24 depend only on d. A, and A. Moreover, ifheR^ a{x + /i), X G R^, then a' e ,^Q{d + 3, A, A) awcf (2.25)
||X„/-X.,/|| < <^2.25 sup \\a{x + /z) - a(x)\\
(2.26)
and a'{x) =
\(KJ(s,
x') - K^J(s,
x')) - (KJ(s,
||/||La+3([o. DX/J^)
x') - K^J(s,
x'))\
< C2.26 sup \\a(x + /i) - a(x)\\ \x^ - x^ \ ||/||Ld+3ao, i].m xeRd
where C2.25 ^^d C2.26 depend only on d, A, and A. Proof. Clearly (2.23) is a special case of (2.8) and (2.24) follows easily from (2.4) plus the boundedness of (/ - D^^^y ^ on L^"'^([0, 1) x R'^) into itself. Also, (2.25) can be viewed as a special case of (1.50). Thus, we can restrict our attention to the derivation of (2.26). Just as in the proof of Lemma 9.1.12, we see that if t] = sup^ \\a(x + h) - a(x)\\ then (2.27)
||(/ - D , , ) - y - (/ - D„,,)-y||,..3«o, i)xK-) ^C2,27*l\\f\\u*3(lO,l)^R^)'
where Cj.a? depends only on d. Thus, by (2.4): \(K,f(s,
x ' ) - K,.f{s,
x')) - (K^{s, x') - KAs,
x'))\
= |G, o [(/ - D „ , , ) - ' - (/ - D„.,)-']/(s, x ' ) - G, o [(/ _ i)„ ,)-1 _ (/ - D„, , ) - i]/(s, x^) I
and so (2.26) is proved.
~ ^
!'/.II/||L''+3([O, i)xi?d)5
Q
9.2.9 Lemma. Let a e s/o(r, A, A) for some 1
d^ dxi dxj'
244
9. Some Estimates on the Transition Probability Functions
Given r
(2.28)
+ 3, define 6 by \/p = e/(d H- 3) + (1 - 0)/r. Then
(f^ s^ jjp(s,
X, y + /i) - p(5, X, y)\^' dy\ '
a(y)\\\
and ( [ s" f (2.29)
\(p(s,x\y-\-h)-p(s,x\y)) - (p(s, x ^ y + /i) - p(5, x ^ y)) p' ^>; j ' ' ' < C^J
|/i| 4- sup ||a(>; + /i) - a(y)|| j |x^ - x^ |,
w/i^r^ a > 0 is the same as in Lemma 923 and the constants C2.28 ^^^ ^2.29 depenc only on d, r, p, A, am/ A. Proof. Suppose we can prove that: (2.30)
{[
ds\ ^\p{s,x,y
+ h)-p{s,x,y)\^dy\
'
ds^
\p(s, x\ y •¥ h) - p(s, x\
y)\
- (p(s, x\ y + h)- p(s, x\ y)) |^ dyf'^ < C2.3i( I/i I + sup \\a(y + h)- a(y)\\ j |x^ - x^ |,
where l/
9.2. The Homogeneous Case
245
function €>: (0, oo) -• (0, oo), depending on d, r, A, and A, such that lim^ ^Q 0(a) = 0 and (2.32)
f I p(s, x,y + h)- p(s, x, y) \ dy < ( 5 A l ) - ^ ( D | | ; i | + s u p | | a ( y + fi)-a(>;)||j.
Moreover, for each S>0 there is a non-decreasing function ^ : (0, oo)-* (0, oo) depending on S as well as d, r, A, and A such that Xvoa^^Q ^(e) = 0 and (2.33)
[
I p{s, X, y 4- /i) - p(5, X, y) | dy
^\y-x\>8
< 4 ' ( | / i | + s u p | | a ( y + ;i)-a(y)||J. Proo/. Choose r < p < (d — l)/d. Then p' > d. We can therefore apply Theorem 9.2.4 (in the same way as we did in the derivation of Lemma 9.2.5) to (2.28) and (2.29) and thereby obtain (2.34)
l^jjp(s,x,y
+
h)-p(s,x,y)\^Y < C2.3^(sA\)-{^\h\
+ sup \\a(y + /i) - a(y)\\ J,
where v > 0 and y > 0 depend on d and p while C2.34 depends on d, r, p, A, and A. Since J
|p(s, X, y 4- /i) - p(s, X, y) I rfy <
so long as \h\ < <5/2, we can use Holder's inequality to show that if \h\ < 3/2 then (2.35)
(f
|p(5,x,y + ; i ) - p ( 5 , x , ) ; ) | ^ P < €2.35(1/I I +snp\\a(y
where G = 2p7(p' 4-1). Because I l>'-^l>i?
p(s, X, y) ^y ^ 0
+ h)-a(y)\^
,
246
9. Some Estimates on the Transition Probability Functions
as R /^ 00 at a rate depending only on d and A, it is easy to obtain (2.32) and (2.33) from (2.34) and (2.35), respectively. D 9.2.11 Lemma. Let a\ R!^ -^S^he a bounded measurable function and suppose that there is an R> 0, an x^ e R'^, and ana e s/o(p^ ^^ ^) ^i^h p <{d — l)/d such that a = a on B(x^, R). Let P be a solution to the martingale problem for
starting from (0, x^). Then for 0
\E'[f,(x(s))-f(x(s))]\
<5-^
+sup \\a(y + h) - a(y)\\^
forfe C?(B(x^ r)). (Recall that fj(x) = / ( x - h).) Proof Define T„, M > - 1 , as in Lemma 9.1.6 with 5^ = 0, Ri = (R + r)/2, and R^^RAige Co(B(0, {R + r)/2)), then
\E'{g{xm\ <\E'[g{x{s)\T,>s]\ CO
+ 1 1
<
\E''[g(x(s)lx2n-l<S
|£^^-[^(x(s)),To>5]| 00
+ ^
/
|£:''^^^2-^"'>->[g(x(5-T2„-l(C0))), TO
>S-r2n-l(C0)]\P{d0j)
where {P^: x e R'^} is the family associated with
Noting that P,(x(r) e r, To > 0 = Pit, X, r ) - E^iP(t - To, X(TO), r), TO < r], where F(r, x, •) is the transition probability function determined by L, we now see that the desired estimate follows by taking g =fh —f with \h\ < (R — r)/2, and then applying (2.32), (2.33), and Lemma 9.L6. D Starting from Lemma 9.2.11, we can derive the next result in exactly the same way as Theorem 9.1.15 was obtained from Lemma 9.1.14.
9.2. The Homogeneous Case
247
9.2.12 Theorem. Let a: R^ -^ S^ he a uniformly continuous function with S(*) as a modulus of continuity and assume that /l|^|'<<^, a(x)^>
xeR'
and
0 e R'.
Let p(s, X, y) be the density associated with 1
'^
d^
Then there is a non-decreasing function O: (0, oo) -• (0, oo), depending only on d, A, A, and d{'\ such that \mi^\Q 0(e) = 0 and (2.37)
f \v(s,x,y
+
h)-p(s,x.y)\dy<s-^
where v > 0 depends only on d, A, and A.
Chapter 10
Explosion
10.0. Introduction Up to this point we have discussed the martingale problem only in connection with bounded coefficients. However it should be evident that many of the more difficult aspects of the martingale problem are really local and do not depend on the global properties of the coefficients. Thus, it should be, and indeed it is, a simple exercise to extend most of the results of Chapters 6, 7 and 9 to martingale problems associated with locally bounded coefficients. The one place at which difficulties can arise is the place where our methods were not local, i.e., in the proof of the existence of solutions to the martingale problem. Although the local existence of solutions is determined completely by the local properties of the coefficients, it is impossible to predict on the basis of local considerations whether a diffusion " runs out to infinity " in a finite time. Perhaps the best way to see this is to look at the simple ordinary differential equation
(0.1)
dx
It
= x^(t\
x(0) = 1.
Clearly any solution to the martingale problem for a{x) = 0 and b{x) = x^ starting from X = 1 at time 0, is concentrated on paths which satisfy (0.1). But this means that for 0 < r < 1 x{t)
1 1-f
= 1
and so lim x(t) = 00
= 1.
b T 1
In other words the martingale problem has no solution for times after 1. In Section 10.1 we will reformulate the resuhs of Chapters 6, 7 and 9 when the boundedness assumption on the coefficients is replaced by the assumption that the
10.1. Locally Bounded Coefficients
249
associated diffusion does not explode. In Section 10.2 we will find some conditions on the coefficients that ensure that the process does not explode, i.e., does not run away to infinity in a finite time.
10.1. Locally Bounded Coefficients Let a: [0,oo) x R^ -^ Sd and b: [0,oo) x R^ ^ R^ be locally bounded measurable functions. Define, as usual,
We say that the probability measure P on (Q, J^) solves the martingale problem for L,, starting from (s, x) if P[x(t) = x, 0 < f < s] = 1 and
f(x(t))-j{L,f)(x{u))du is (Q, ^f, P) martingale for times t > 5, for all functions/in CJ(/?'*). The essential fact on which our analysis of such martingale problems rests is contained in the following theorem. 10.1.1 Theorem. Let a: [0, oo) x R'^-^ S^ and b: [0, oo) x R'^-^ R'^ be locally bounded measurable coefficients and let L, be defined by (LI). Given a solution P to the martingale problem for L, starting from (s, x), a stopping timeriQ. -• (5, 00), and a r.c.p.d. {P^^} ofP given Ji^, there is a P-null set N e M^, such that for all (o ^ N, ^x(r{co), o>) ® x{(o) Poi solves thc martingale problem for L^ starting from (T(CO), x(x(o}), oj)). Moreover, if the martingale problemfor L^ is well-posed and {Pg, x- (^^ ^) ^ [0, 00) X R'^} is the associated family of solutions then the map (s, x)-^ P^^x is measurable and forms a strong Markov family. Finally, if the martingale problem for L, is well-posed and a: [0, 00) x R'^ ^ Sj and b: [0, 00) x K*^ -> R*^ are a second set of locally bounded measurable coefficients such that a = a and b = b on some bounded open set G c= [0, 00) x R'^, then for (s, x) e G and any solution P to the martingale problem for a and b starting from (s, x): P equals P^^^ on M^, where T = inf{r > 5: (r, x(t)) i G]. Proof. The first assertion is derived from Theorem 1.2.10 in exactly the same way as we proved Theorem 6.1.3. To prove the second part, once the first part has been established, it is obviously enough to check the measurability of {Ps,x'- [s^ ^) ^ [0, 00) X R^]. But the argument given in Exercise 6.7.4 depends in no way on the coefficients being bounded and therefore applies in the present context as well. Finally to prove the last assertion, take P = P and Q^ = F,^^) ^^^^^^ ^) in Theorem 6.1.2. Then by that theorem P®t()G- solves the martingale problem for L, starting from (5,x) and is therefore equal to Ps,x' At the same time, by construction, P ®t(.) Q. equals P on Ji-^. Thus P equals Ps,x on Jt^. D
250
10. Explosion
10.1.2 Corollary. Let a: [0, oo) x R^-^S^ and b: [0, oo) x / ? ' ' ^ i?" be locally bounded measurable functions. Assume that there exists an increasing sequence of bounded open sets G„ cz [0, oo) x R^ with [j„ G„ = [0, oo) x R^ and bounded measurable coefficients a„: [0, oo) x R^ ^ S^ and b„: [0, oo) x R'^-^ R^ such that [0, oo) x R'^ — \J^ G„, an = a and bn ^ b on Gn, and for each n the martingale problem for an, bn is well posed. Then for each {s,x) there is at most one solution to the martingale problem for a, b starting from (s,x). Moreover if P^^^ is the solution for an, bn starting from {s,x) and if Xn = inf{r > s:{t,x(t)) ^ G„}, then a solution for a and b starting from (s, x) exists if and only if (1.2)
lim P<"yT„ < T] = 0 for each
T > s.
Finally if (1.2) obtains, then the unique solution P for a and b starting from (5, x) equals P^"^^ on Jf^^ for every n. Proof In view of Theorem 10.1.1 there is really very little to be done. Indeed, if a solution P for a and b starting from (5, x) exists, then for every n, P must agree with Pl% on J^,^. Since P[x„ < T] -• 0 as n -»• 00 for every T > s, this proves the uniqueness of P as well as the necessity of condition (1.2) for the existence of P. Finally, to prove the sufficiency of (1.2), observe that P^"l^^ equals Pl"\ on Jf^^. Thus, by Theorem 1.3.5, if (1.2) holds there is a P which equals P^^\ on ^ , ^ for every n. Clearly such a P is a solution to the martingale problem for a, b starting from (s, x). D We now specialize these results to the case in which we can use the results of Chapter 7. Namely, let a: [0, 00) x P'^ -> 5^ and fe: [0, 00) x P^ -• R'^ be locally bounded measurable functions and assume that for every T > 0 and x e R'^: (1.3)
inf
inf <0, a(s, x)^> > 0
o < s < r \d\ = i
and (1.4)
lim sup \\a(s, y) - a(s, x)\\ = 0. y^x0<s
Given « > 1, let G„ = [0, n) x P(0, n) and choose il/„ e C^{[0, 00) x P'^) so that 0 < iA„ < 1, i//„ = 1 on G„ and iA„ = 0 outside G„+ j . Set (1-5)
a„ =
i^„a^(l-ikn)I
and (1.6)
b„ = ,l^„b.
10.1. Locally Bounded Coefficients
251
Then it is clear that a„ and b„ satisfy the hypothesis of Theorem 7.2.1, and therefore the associated martingale problem is well-posed. Denote by {P^"\: (s, x) e [0, oo) X R**} the corresponding family of solutions. (We note that the measurability of {P^"y is established independently of Exercise 6.7.4.) Using Corollary 7.1.2, we have now proved the next theorem. 10.1.3 Theorem. Let a: [0, oo) x R'^-^S^ and b: [0, oo) x R'^-^R'^ be locally bounded measurable functions and assume that a satisfies (1.3) and (1.4) for each T > 0 and x e R^. Then for each (5, x) G [0, 00) x R** there is at most one solution to the martingale problem for a and b starting from (5, x). Moreover, ifa„ and b„ are defined by (1.5) and (1.6), respectively, and if{P^"\: (s, x) G [0, 00) x R**} denotes the family of solutions to the corresponding martingale problem, then a solution for a and b starting from (s, x) exists if and only if 1.7)
lim P<") n->ao
sup \x(t)\ > n
0.
s
Finally, if (1.7) holds, then the unique solution F^.x satisfies (1.8)
Ps,,=^Pt\
on
M,^
for every n> s, where T„ = (inf{r > s: \x(t)\ > n}) A n. 10.1.4 Corollary. Let a and b be as in Theorem 10.1.3 and assume that (1.7) holds for each (s, x) G [0, 00) x R^. Then the martingale problem for a and b is well-posed and the associated set {Ps,x' (s^ ^) ^ P^ ^) ^ P''} of solutions form a measurable strong Markov family. Moreover, given T > 0 and x e R"^ let A=
inf o
A = sup
inf (9, a(t, x)e) \e\ = i
sup {6, a(t, x)6)
0
\d\=l
B = sup
sup
I b(t, y) I
0 < r < r |>'-x|< 1
and for 0 < 3 < i, define p(S)=
sup 0
sup y:
\\a(t, y) - a(t, x)\\.
\y-x\
Then for each s e [0, T] and s> 0, there is a 3^> 0, depending only on d, k. A, B, p('), T — s and z, such that (1.9)
\E^-i^]-?''<'•
i
<8\m
252
10. Explosion
where : Q-^ R is a bounded J^^( = (T(x(t)): t > T) measurable map and (or, y) e [0, T] X R^ satisfies \(7 — s\< 3^, and \y — x\
i\\t-srdt
f
Vs
*'B(0,r)
where (x = (d -\- 2)(q — l)/2 and ^T,R=
\p(s,x,t,y)\^dyY\c,,,o. /
CIIQ
depends only on d, T, q, R — r,
ii^f o<s
inf (9, a(s, x)6} \e\ = i
X e B(0, R)
AT,R=
sup o<s
sup {6, a(s, x)9) \e\ = i
xeB(0,R)
BT,R=
sup | % x ) | 0<s
and the function PT,R'- (0, 1] -• (0, oo) given by PT,R(S)=
sup
\\a(s,Xi)-a(s,X2)l
0<s< T xi,x2eB(0,R) \xl-x2\
and ifO
then l\
dt [
\p(s,x,t,y\'^dy)
where Cj n depends on d, T, q, ^T,R^ ^T,R^ PT,R(')
'
Proof. The first assertion follows immediately from Theorem 10.1.3 and Theorem 10.1.1. (Avoiding Exercise 6.7.4 if desired.) The proof of (1.9) follows immediately from (1.8) and Theorem 7.2.4 applied to a„ and b„. Similarly, (1.10) and (1.11) follow at once from Corollary 9.1.10. D 10.1.5 Corollary. Let a: R'^-* 5 / be continuous and b: R'^-* R'^ be locally bounded and measurable. Let {ij/„: n > 1} be a sequence of non-negative functions in CQ(R^) such that XB(O, n)<^n< XB(O, n+1) and set a„ = ij/^a-\- (1 - iA„)/ and b„ = \l/„ b. For each n> \, let {PJ: x e R'^) be the family of diffusions associated with a„ and b„ (as in Lemma 9.2.1). Then the martingale problem for a and b is well posed if and only if lim„^^ P?^(supo<,
10.1. Locally Bounded Coefficients
253
Px = PQ, X ^^ s: Q -^ Q 15 as described in Lemma 9.2.1. Thus {P^: x e R^]forms a measurable, time homogeneous, strong Markov family with the property that for each X e R^, T > 0 and £ > 0, there exists a S^> 0, depending only on d, T, X= inf<^, a(x)^> A = sup<^, a(x)0> B = sup-|b(x)| y:\y-x\<\
and the function p: (0, 1] -* (0, oo) given by p(3)=
sup
\\a(y) -
a(x)l
y:\y-x\
such that |£''.[(D]_£^^[
<s\m
whenever \y — x\ < S^ and O: Q^R is a bounded M'^ measurable function. Finally, if P(s, x, •) is the time homogeneous transition probability function associated with {P^: X e R'^}, then P(s, x, •) has a density p{s, x, y) and for all l0 and 0
(1.12)
f[
\p(s,x,y)\^dyy''
where v depends only on d and q and C1.12 depends only on d, q, R — r, Af^= inf \x\
mf(e,a(x)e} 191=1
AR = sup sup {0, a(x)6} \x\
BR=
\d\=l
sup \b(x)\ \x\
and the function p: (0, 00) ->• (0, 00) given by p{S)=
sup
||a(x')-a(x^)||.
ixl|
Proof. Only the last part is not an obvious consequence of Corollary 10.1.4. On the other hand, the last part follows easily from Corollary 9.2.7. G
254
10. Explosion
10.2. Conditions for Explosion and Non-Explosion In this section we discuss a few simple criteria which guarantee explosion or non-explosion. We begin with the following condition which relates the problem to a purely analytical one. 10.2.1 Theorem. Let a, b, a„ and h„ he as in Corollary 10.1.2 with G„ = [0, n) x B(0, n). A sufficient condition that the martingale problem fi)r a, b to be well-posed is that for each 7 > 0, there exists a number /I = /ly^ > 0 and a non-negative function (p = (PTe C^ \[0, T] X R'^) such that lim inf (p(t, x) = oo |jc|->oo o
(p(T, x)
xeR''
and
(i'-'^-y
> Xq)
on [5^, T] X R^ then there is no solution to the martingale problem starting from (s^, x^) corresponding to a and b. Proof To prove the first part, let 0 < s < T and x e R*^ be given. Choose nQ> Tso that (5, x) G G„ for n>nQ. Let (p = (pjhQ the function in the hypothesis. Since {dcpldt) ^ L,(p - Xcp <{) on [0, T] x R\ we observe that e - X s , X) > f^i^-^^^^^-ViT^ AT„, X(r AT„))] >.-^^£n"».[^(T„,x(T„));T„
10.2. Conditions for Explosion and Non-Explosion
255
and |X(T„) | = W if T„ < T, it follows that lim„_.^ Ps"x(^n < T) = 0. Thus the first part has been proved. We now turn to the second part. Choose HQ SO that (5^, x^) e G„o and define T„ for n>no as before. Since we are now assuming that (d(p/dt) -h LtCp — ?.(p >0 on [s^, 00) x R'^, we have for our T > s^ and « > MQ AsO
+ ( sup
f
sup
If lim„^^ P^so,xo{^n ^ T\ were zero we would then have ^-^^V(5^ x^) < ^"'^ sup (p(r, x) contradicting our assumption. The proof of the second part is therefore complete. D A rather crude but nevertheless useful application of Theorem 10.2.1 is the following result. 10.2.2 Theorem. Let a, 6, a„, h„ and G„ be as in Theorem 10.2.1. If for each T > 0, there is a Cj < 00 such that (2.1)
sup ||a(r, x)\\ < CT{1 + \x\^),
for x e R'
0
and (2.2)
sup <x, b(t, x)> < CT(1 + |x |2),
for x e R"^
0
then the martingale problem for a, b is well posed. Proof Let (p(x) = 1 + \x\^ ioi x e R*^. Then L^(p{x) = Trace(a(r, x)) + 2(x, b(t, x)) < ( J + 2)Cr(l+ \x\') = XM^) where Xj = [d + ^)CT- Thus we can apply the first part of Theorem 10.2.1 with (py,= 1 -f | x | 2 and kj = (d + 2)CT.
D
The reason that Theorem 10.2.2 is crude is that it does not take into consideration the dimension d of the space or the effect of any balancing between a and b. A more refined result is the test of Hasminskii which we will present below. The treatment is adapted from H. P. McKean [1969].
256
10. Explosion
10.2.3 Theorem. Let a, h, a„, h„ and G„ be as in Theorem 10.2.1. Suppose for each T > 0 there exists an r > 0 and continuous functions A-pi [r, oo) -> (0, oo) and Bj'. [r, oo) -• (0, oo) such that for p > (2r)^ 0 < t < T and \x\ = p,
<x, a(t, x)x>B7^jy j > (Trace a(t, x) + 2<x, b(t, x ) » and f[Cr{p)r
' dp flCr(a)][Ar(cT)]-'
d
where CT(P)
= exp fBr(c7] da
Then the martingale problem for a, b is well-posed.
Proof. Let T > 0 be given. For p > r define, by induction, UQ = I and u„{p) = 2 j\c(
d<j \' ^ ^ «„_ ,(T) dx,
where A = Aj, B == Bj and C = Cj. Using induction, one sees that for all n > 0, u„>0,u'„>0 and (2.3)
u„(p) < -^ \^j\c{o)]-'
da [ ° 1 ^ dr j " .
Finally (2.4)
K = j^n-i-Bu'„
for
p>r.
From these observations, it is easy to see that the series ^^=o "n converges uniformly on compact subsets of [r, oo) and can be differentiated twice term by term. The sum u is twice continuously differentiable on [r, oo) and is such that w > 1, u' > 0 and (2.5)
u = jA(u" + Bu'),
for p >r.
10.2. Conditions for Explosion and Non-Explosion
257
Finally, since u>u^, our hypothesis implies that (2.6)
limM(p) = 00. P t 00
Now choose cp e C^(R'^) so that (p > I and (p{x) = u(\x\^/2) for | x | > (2rf. Certainly we can find a A > 1 so that L, (p(x) < A for all 0 < r < T and | x | < (2rf. Also if 0 < r < T and |x | > (2r)^, then
+ (i Trace a(t, x) + <x, b(u x)))u'l^^
-m
j
(p(x).
We have used (2.5) to obtain the last line and also to show that u" -\- Bu' > 0, thereby justifying the second to the last line. It follows that LfCp < Xcp on [0, T] X R^, and the first part of Theorem 10.2.1 can now be applied to finish the proof, n We end this section with a partial converse to Theorem 10.2.3. The idea again goes back to Hasminskii.
10.2.4 Theorem. Let a, b, an, /?„ and G„ be as in Theorem 10.2.1. We now assume that for each R>0 inf inf inf {6, a{t, x)6} > 0 \e\ = i t>o \x\
and sup sup \b(t, x)\ < oo. t>o \x\
In addition, assume that there are continuous functions A: [^, oo)--^ (0, oo) and B: [i, oo) -• (0, oo) such that for p > I, t >0 and \x\ = p, ^(yj<<x,a(r,x)x>, <x, a(t,
X)X>B|^|
< (Trace a(t, x) + 2<x, b(t, x)>),
10. Explosion
258 and C{a)
l^^^^^^^'^lW)'"
< 00
where C(p) = exp f B((T) do Then there is no (s, x) e [0, oo) x R^ starting from which the martingale problem for a and b has a solution. In fact, for each (s, x) G [0, oo) x R'^, lim lim P[%[x„ < T] = 1, where x„ =
mf[t>s:(t,x(t))^G„l
Proof The first step in the proof is to show that for each e > 0 there is an R^ > 1 and a Te > 0 such that (2.7)
n'.(r„ < 5 + T,) > 1 - £
for all n > 2, s > 0 and |x | > R^.To prove (2.7),define{u„,n> 0} and uas in the proof of Theorem 10.2.3 using the functions A and B given in the statement of this theorem. Once again u e C^([^, oo)), w > 1, w' > 0 and u = iA(u" + Bu'). Moreover, by (2.3) u(p) < exp
'L^'^'^^^-'''^lw)'^
and so 1 = w(^) < u(p) and u{p) increases to a finite limit M(OO) as p -• OO. Define (p(x) = ^ ( | x p / 2 ) for xeR'^\B(0,l) and observe that L,(p>(p on [0, oo) x R'*\B(0, 1). Thus if n > 2, and (5, x) e G„ with |x | > 1, we have En"H^--Xx(aAT„))]>e->(x) where a = inf{r > s: |x(f)| < 1}. In other words .2
+ e-"E^^(p(x(n)l
n
Because En"'ie-\
x„<(7An]-\- £ ^ ^ % - ^ d < T J < 1
^"'"(-^l
10.3. Exercises
259
we obtain from the above
t/(^J-t/(i)-e-<"-Moo) (2.8)
£n">.[e-Un-s)] >
u(co)-u(i)
for all « > 2 and (s, x) e G„ with I x I > 1. Now choose R^ > 1 so that
-(JiE)-m: w(oo)-w(i)
>1-^ 4
for all
\x\>R.
and take 7, > 0 so that 2(l-e--)>l ^ ^
and
-f^^^.' M(oo)-w(i) 2
We claim that (2.7) holds with the choice of R^ and T^. Indeed, let n > 2, s > 0 and IXI > Rg be given. If either |x| >n or n <s -\- T^, (2.7) is trivial. But if n> |x I V (5 + r,), then by (2.8) and the choice of R, and T,, L
J _
2'
and so
KA^. >s+Z)< ^ ^ 3 ^ E^m - e-"'-'^]
^W---e^)^'Having estabUshed (2.7), we will be done once we show that for each R > 0 and (5, x) 6 [0, 00) X R^
(2.9)
lim
sup
r-^oo fi>i?v(s+r)
P?U
sup
\s
| x(t) | < R) = 0. }
However, (2.9) is an easy consequence of Exercise 3.1 below.
•
10.3. Exercises 10.3.1. Let a: [0, 00) X R'^ ^ 5^ and 6: [0, 00) X /?'' -• R'^ be locally bounded measurable functions. Let i? > 0 be given and assume that k — inf,> 0 infj^i = 1 inf|;c| 0 and that B = sup,>o sup|;,|<j? |^(t,x)| < 00. Suppose that P solves
260
10. Explosion
the martingale problem for a and b starting from (s, x) e [0, oo) x B(0, R) and let T = inf{r > s: \x(t) \ > R}. Show that there exists an e > 0 and a C < oo depending only on d, R, X and B such that £^[exp[£(T — 5)]] < C. There are many ways in which this can be done. One of the most straightforward is the following. Define (p^{x) = 1 - e'^'^"''^'/^ for x,r]e R^. Show that there is an ^/ e R^ and an £ > 0 depending only on d, R, X and B such that z < inf,>o inf|y|^{y). Conclude that £''[exp[e(T - S)](;P^(X(T))] < (p,(x). 10.3.2. Let a: [0, 00) X R*^ -• Sd and /?: [0, 00) x R'* -• R*^ be locally bounded measurable coefficients and let c: [0, 00) x K** -• i^*^ be a measurable function such that ac is locally bounded. Assume that for some (s^, x°) e [0, 00) x R^ the martingale problems for both a, h and a, b + ac starting from (5^, x^) have exactly one solution each, say P^o ^0 and Q^ ^0 respectively. Show that for each t > 5^, P^ ^0 and Qso.jco are equivalent on ^ , and develop a representation for the RadonNik odym derivative dQ^o^^o/dP^o^xo on J^^ (at least as a limit). 10.3.3. To illustrate how the dimension d enters in the criterion for explosion, let a:R^ -* (0,oo) be a locally bounded measurable function such that a is uniformly positive on compact sets of R'^. Let L = ja{x)A. If d equals 1 or 2 show that the martingale problem for L is well-posed. U d>3 and if there is a continuous a: (1, oo)-^(0, 00) such that (x(p) l and | x | = p , and Jf l/a(^/p) dp < CO, then the martingale problem for L has no solution starting from any point. The reason for the difference between d <2 and ^ > 3 can best be understood in terms of the recurrence properties of Brownian motion. Indeed in one and two dimensions Brownian motion is recurrent and so the existence can be seen by the techniques introduced in Section 7.5. On the other hand, if f/ > 3, then it is easy to see that a solution P starting from (0, x) e R'^ would have the property that 00
I
j XB(o,RMt))dt\ = Q | 'B{0,R)
Id-2 I^
a(y)
dy
which means that limE''
XB(0,RMt))dt
< 00
JR-»oo
if there is an a( •) with the stated properties. Of course one can also use Hasminskii's test, i.e., Theorem 10.2.4. 10.3.4. To illustrate how the relations between a and b can influence explosion, let ^ > 3 and take a(x) = (|x | v 1)^"'' and ^(x) = - (p/2) \x \'x for some £ > 0 and p€ R. Show that if p >d — 2 then the martingale problem is well-posed. If p < d — 2, show that there is no solution starting from any point.
Chapter 11
Limit Theorems
11.0. Introduction We have estabhshed conditions on coefficients which guarantee existence and uniqueness of solutions to the associated martingale problem. What we want to do in this chapter is to show that these same conditions also guarantee stability properties for these solutions. For example, suppose that a: [0, co) x R*^ -^ S^ and b: [0, co) X R'^ ^ R^ are continuous coefficients for which the martingale problem is well-posed. Let a„: [0, oo) x K'^-^Sj and b„: [0, oo) x R'^-^ R^ n > 1, be coefficients which tend to a and b, respectively, uniformly on compact subsets. Assume that P" is a solution to the martingale problem for a„ and b„ starting from (s„, x") and that {s„, x") -^ (s, x). If the a„'s and b„'s are bounded independent of n > 1, then {P": n> 1} is relatively compact, and it is clear that any limit point must be a solution for a and b starting from (s, x). Hence P" -* Ps,x^ the unique solution for a and b starting from (s, x). This example is the prototype of one kind of stability result in which we will be interested. A second sort of stabihty with which we will be concerned has as its prototype the celebrated invariance principle of M. Donsker (cf. Exercise 2.4.2). From the analytic viewpoint, this type of stability entails the approximation of solutions to various differential equations involving the coefficients a and b by solutions to finite difference equations. Our presentation will be somewhat complicated by our desire to show that these stabihty results follow from the assumption that the martingale problem for a and b is well-posed and do not depend on any particular condition which might be made to insure that this assumption is justified. In particular, we do not want to assume that a and b are bounded but only that the associated martingale problem admits exactly one solution. The reader should realize that, unlike results about uniqueness, extending stability results from bounded to unbounded coefficients involves proving compactness without the benefit of global bounds on the coefficients. One of the interesting features is that the mere existence of a solution for the limiting coefficients already implies the compactness of any sequence of diffusions corresponding to approximating coefficients.
262
11. Limit Theorems
11.1. Convergence of Diffusion Processes In this section we are going to give our basic convergence result for diffusion processes. Refinements of this resuh will be given in Sections 11.3 and 11.4 under more restricted conditions. In order to have a compactness condition when our coefficients are not necessarily bounded, we will need the following lemma. 11.1.1 Lemma. Let {P": n>\] be a sequence of probability measures on Q and suppose that {x^: k > \} is a non-decreasing sequence of lower semicontinuous stopping times increasing to oo for each co. For each n> \, let {P"''^: n> 1} be a sequence of probability measures such that for each /c > 1, P"' ^ equals P" on M ^^ and {P"''': n > 1} is relatively compact. Finally, assume that P is a probability measure on Q such that for each k> i and any limit point Q^ of {P^'^:n > 1}, Q^ equals P on J^tk' Then {P":n > 1} converges to P. Proof First observe that for each k > 1 and t > 0, {T^ < t} is closed in Q. Thus, since {T^ < t} e J4^^, (1.1)
lim sup P"(Tfc
P(xk < t)
n-*oD
because P"(Tfe < r) = P"' ''(T^ < r), {P"''': n > 1} is relatively compact, and any limit point of {P"'*: n > 1} equals P on ^^^. Now suppose O: Q^R^ is a bounded continuous ^^-measurable function. Then, since ^X(t, oo)('^k) is J^^^-rnQSismable, (1.2)
lE^^O)] - £^[0] I < \E''"'[^] - £^[0)] I + 2||0||P"(Tfc < t).
For fixed A; > 1, choose a convergent subsequence {P" *} of {P"'': « > 1} so that lim |£:^''''[0] - £^[(D]| = lim sup lE^-'^O] - £''[0]|, and let Q'' = lim„'^^ P" *. Since Q'' equals P on Jf^^, we see that lim sup l^^-'t^] - E^[^]\ = I ^ ^ W - E^[(^]\ < 2||a)||P(Tfc < r). Combining this with (1.1) and (1.2), we arrive at: lim sup |£^''[
for all /c > 1. Since P(Tfc < r ) \ 0 as fc -• oo, this completes our proof
Q
11.1.2 Lemma. Let P be a probability measure on Q and for each R> 0 let T^ = inf{r>0: \x(t)\
>R}.
11.1. Convergence of Diffusion Processes
263
Then T^ is lower semicontinuous for all R> 0. Moreover, with the exception of at most countably many Ks.for P-almost all co e {TJ^ < oo}, TJJ is continuous at co. Proof. It is clear that T^ is lower semicontinuous. It is equally clear that T^ = inf{r > 0: \x{t)\ > R} is upper semi-continuous and that T^ = limsx^ T5. Since T^ > T^ , we will be done once we have shown that £''[e~ "^^] = E^[e~ '^] for all but a countable number of R\ because when this equahty holds T^ = T^J , P-almost surely on {T^J < 00}. But R -• E^le'^'^] is a non-negative, non-increasing function, and so £^[e-^«^] = Hm E'^le-''] = E''[e-'''] S\R
for all but a countable number of R's.
D
11.1.3 Lemma. For each n> I let P" be a probability measure on Q and let X„: [0, 00) X Q-^ R^ be a continuous progressively measurable function. Assume that there is an Ro>0 such that ifC = inf{r > 0: \x(t) | > RQ} then sup sup I X„(t A C(co), co) I < 00 n> 1 (t,(o)
and for all n > 1, {X„(t A C), J^t^ P") ^^ ^ martingale. Finally, suppose that P is a probability measure on Q and that X:[0, 00) x Q^ R^ is a progressively measurable function which is jointly continuous in (t, co) such that P„-^ P and X„^ X uniformly on compact subsets of [0, 00) x Q. Then (X(rAC), J^t^ P) i^ ^ martingale. Proof First assume that at F-almost all co e {C < 00}, C is continuous. Then C A t is P-almost surely continuous for all t > 0. Let 0
E''[X(ti A 0 ^ ] = lim E''"[X(ti A C)0],
i = 1, 2.
On the other hand, I £'""[Jf (t, A O * ] - E-"[X„{t, A C)*] I < ||||£T I X(t, A C ) - X„{U A C) I ], and it is easy to see from our assumption that £''"[|Z(r, A f) — X„(r, A C)|] -• 0 as n -• 00. Thus we may replace X(r, A f) by X„(ti A () on the right-hand side of (L3). In particular,
E'[x{t2 A cm = lim E'"[x„(t2 A cm = E''[x{t,At:yt>i
264
11. Limit Theorems
To eliminate the assumption about the continuity of C, we use Lemma 11.1.2 to find R„ /RQ such that C„ = inf{f > 0: \x(t)\ > R^ is P-abnost surely continuous on {C„ < oo}. By the preceding, {X{t A C„), J^t, /^) is a martingale for all n > 1; and therefore, since X(t) is bounded and continuous and C„ /(>, {X(t AQ^J^t^P) must be a martingale. D We are now ready to prove the main resuh of this section. The reader should be aware that the preceding preparatory discussion is necessary only in the case when our coefficients are unbounded. For uniformly bounded coefficients, the next theorem is nearly obvious.
11.1.4 Theorem. Let a: [0, oo) x R^^ S^ and b: [0, co) x R'^-* R'^ be locally bounded measurable functions which are continuous in xfor each f > 0, and assume that for each (5, x) e [0, 00) x R'^ the martingale problem for a and b starting from (5, x) has exactly one solution P^ ^ - Suppose that for each n> 1, fl„: [0, 00) X R^-^Sa and b„:[0, 00) x /?'*-> R** are measurable functions, and assume that for all T > 0 and R> 0: sup sup
sup \\a„(s, x)\\ + \b„(s, x)| < 00
«> 1 0 < s < r |x|
and lim
sup (\\a(s, x) - a„(s, x)|| + |b„{s, x) - b(s, x)\)ds = 0.
n-*cc •'0
\x\
Let P", n> U be a solution to the martingale problem for a„ and b„ starting from (s„, x"). / / (s„, x") -> (5, x), then P" -^ P^, x • Proof For k > sup(\s„\ + \x"\), define Cik = inf{r>0: |x(r)| >k} and Tfc = Cit A k. Let / ( r ) = x(t A TJ^), t > 0, and denote by P"' * the distribution of / ( • ) under P". Clearly P"-^ = P" on J^^^. We are going to show that for each /c, |pn,k. ^ > 1} is relatively compact and that any limit point coincides with P^ ^ on ^^^. By Lemma 11.1.1, this will complete the proof that P" -^Ps,x' We first check that {P"'^:n > 1} is relatively compact. To this end, let / G C^iR"^) be given. Clearly , (r V s„) A tk
f{x{t))-f
L:f(x{u))du
11.1. Convergence of Diffusion Processes
265
is a P"' ^-martingale where L" is defined in terms ofa„ and b„. Thus for each/there is a Ckif), not depending on n but only on the bounds on / and its first two derivatives, such that (/(x(r)) + Q ( / ) r , A , P - ^ ) is a submartingale. Therefore, the relative compactness of {F"*: n> 1} now follows from Theorem 1.4.6. We next turn to the identification on J/^^ of any limit point of {P"' ^: n > 1}. Let Q'^ = lim„-^<„ P"'*, where {P"'*} is a subsequence of {P" ^ n > 1}. Clearly Q^(x(t) = x,0
AT,)) -
[ ' " ' L ^ / W I / ) ) du, M,,
e^)
is a martingale after time s for each/e CQ{R^). By Exercise 11.5.1 this will prove that Q* = ?^ ^ on ^ , , . But if X^{t) =f(x{tAk)) - f'"''"" L:'f(x(u)) du and
X(t) =f(x(tAk)) - r''^\^f{x(u))
du^
then sup sup \X„itACk((ol n'
C0)| < 00
(t, (o)
and X„'^ X uniformly on compact subsets of [0, oo) x Q. Moreover, (r, co)-^ X(t, (D) is continuous and (X„,(fAC,),^„P«''*) is a martingale. Thus by Lemma 11.1.3, (X(t A C^^), -^r^ 6*) is a martingale. That is,
f(x(tAx,))-(^^\,f(x(u))du s
is a Q'^-martingale after time s. The theorem is therefore proved.
D
By taking a„ = a and 6„ = ft for all n > 1, we obtain the following corollary of Theorem 11.1.4.
266
11. Limit Theorems
11.1.5 Corollary. Let a: [0, oo) x R^-> Sj and b\ [0, oo) x R'^-^ R"^ be locally bounded measurable functions which are continuous in xfor each t > 0. If for each (5, x) 6 [0, 00) X R^ there is exactly one solution P^ ^ to the martingale problem for a and b starting from (s, x), then the family {Ps,x'' (^^ x) e [0, 00) x R'^} is Feller continuous.
11.2. Convergence of Markov Chains to Diffusions The purpose of this section is to study the approximation of a diffusion process by Markov chains. For simpHcity, we will be assuming that all of our processes are homogeneous in time. However, the context in which our theorems are stated is sufficiently general to allow one to obtain analogous results for timeinhomogeneous processes by simply considering the time-space process. In order to explain what we are about to do, we will first recast Exercise 2.4.2 into the general setting in which we are going to be working (cf Exercise 6.7.2). Suppose that Xj, ..., X„, ... is a sequence of mutually independent Z?*^-valued normal random variables having mean 0 and covariance / on some probability space (£, J*^, /i). For each h>0 and x e i?^ define ^h^' E-^Q by
%,M) = ^-^h'^
^"^^
I
\t]\
(where Yjk = 1 ^k = 0)- Denote by Pj the probabiHty measure on Q induced from /x by ^f, X. An easy computation shows that if g!^ is the distribution on (i?**)^^' '"' ^ of {x(kh): /c > 0} under P j , then {Qj: x e R'^} is the time-homogeneous Markov chain with transition probability n;,(x, dy) given by:
In particular, if A^f(x) = J (/(>-) -f(x)n^{x,
dy), then
{f{x{kh))-''lAj(x(jh)),J^,,,P'-J 0
is a discrete parameter martingale for a l l / e Co(R^) (cf. Exercise 6.7.1). Indeed, this fact together with Pi(x(0) = 0) = 1 and P;[x(r) = [((k + l)/i - t)/h]x{kh) + ((t - kh)/h)x({k + l)/i), /c/i < r < (/c + l)/i] = 1 for /c > 0 uniquely characterizes P^. Thus, we ought to be able to obtain the conclusion of Exercise 2.4.2 directly from this characterization of P j . Once one realizes that this is a possibility, it is not hard to guess how one should proceed. The crux of the matter is contained in the observation that for each/ e C?(P'^): (2.1)
lA,f(x)-^iAf(x)
as
h\0
267
11.2. Convergence of Markov Chains to Diffusions
boundedly and uniformly. With this remark, one can check that the hypotheses of Theorem 1.4.11 are satisfied by {P^^: h> 0}. Hence {F!^: h>0} is pre-compact. Moreover, for each/e C?(K'^) and r > 0: [tlh]-l
X
A,f(x{jh))^jiAfix(u))du
1 0
uniformly on compact subsets on Q. Thus, any limit point Q of {Pj: ^ > 0} has the property that (/(x(r)) — Jo ^ A/(X(M)) i/w, J^t^Q) is a martingale for all fe CS'iR^ and so it is obvious that Q = iT^^K We have therefore shown that Pj -• i^5f^ directly from the martingale characterization of P^ together with (2.1). In particular, the specific form of n;,(x, •) does not matter once one has (2.1). Obviously, one can carry out this line of reasoning in much greater generaUty, and that is exactly what we intend to do. Throughout the remainder of this section we will be working with the following set-up. For each /i > 0, let n,,(x, •) be a transition function on R*^. Given x e R'^, let Pi be the probability measure on Q characterized by the properties that (0P;(x(O) = x ) = l . . (k+l)h-t (») P'. x(t) = ^^ -f (2.2)
,,.. t-kh x(kh) + x((k + l)h),
kh
\)h = 1 for all /c> 0,
(Hi) Pj(x((/c -f l)h) eT\Ji^^) = nM^h\
T) (a.s., P^) for all /c > 0 and
It is easy to see that (i) and {in) are equivalent to saying that the distribution on ^l^dj{o, ...,n,...} Q|- ^(/^fj)^ /c > 0, is the time-homogeneous Markov chain starting from X with transition probabihty n;,(x, •). In particular, from Exercise 6.7.1, [Hi) is equivalent to saying that k- 1
(f(x(kh))-
^A,f(x(ih)\Jl,„Pl]
is a martingale for a l l / e C"(R ), where (2.3)
A,f(x)=^\(f(y)-f(x))n,(x,dy).
We are now going to make the certain assumptions about llh(x, •) as a function of h>0. Namely, define 1 f *(^) = 7 "
(yi~ Xi){yj - x,)n,(x, dy\
*^|y-jc|
1 I^h(x) = T '* *'|)'-X|<1
[yi- ^i)nh(x, dy\
268
11. Limit Theorems
and
What we are going to assume is that for all jR > 0: (2.4)
lim sup \\ah(x) - a(x)\\ = 0, h\ 0 |;c|<J?
lim sup I bf,(x) — b(x) \ = 0,
(2.5)
h\0
\x\
(2.6)
lim sup Al(x) = 0, h\0
£ > 0,
\x\
where a: /?''-• S^ and b: R'^ ^ R"^ are continuous functions. The origin of these conditions is made clear by the following Lemma. 11.2.1 Lemma. The conditions (2.4), (2.5), and (2,6) are equivalent to the condition that for eachfe C^(R'^) Anf^Lf
(2.7)
uniformly on compact subsets of R^, where i d
L=:~
^2
y a'^(x)^-^^
d
+
Q
yb\x)i-
Proof First assume that (2.4)-(2.6) hold. Given/e CS'(R'^l set
H(x,y)=i(y,-xd§^(x)
By Taylor's theorem there is a Cj < oo such that \ny)-nx)-H(x,y)\
A,f(x)-L,f(x)
\y-x\'n,(x,dy)
''\y-x\
+ [
\f(y)-f(x)\n,(x,dy).
11.2. Convergence of Markov Chains to Diffusions
269
Since (2.4) and (2.5) guarantee that L;, / - • Lf uniformly on compacts, (2.7) will follow once we show that \y-x\'Tl,(x,dy)-^0
f \y-x\
and j
\f(y)-f{x)\U,(x,dy)-.0
^\y-x\>l
uniformly on compacts. But for 0 < e < 1:
1
y-x
\^lly,{x, dy) < e^ + A^x),
y-xui
and clearly f
|/(y)-/(x)|n,(x,c/);)<2||/||A);(x).
Thus (2.6) guarantees the desired convergence. Now suppose that (2.7) obtains. We will prove (2.6) first. Clearly, it is enough to show that if x,, -^ XQ as /i -• 0, then for each e > 0: limA^x,) = 0. To this end, let (p e CQIR'^) be chosen so that 0<(p
(2.8)
lim sup j h\0
(/(>^)-/(x))n,(x,^y)-L/(x) = 0.
\x\
Indeed, we can then simply take / successively equal to x^ and x,Xy, 1 < ij < d, and thereby obtain (2.4) and (2.5) after a little easy algebra. But to prove (2.8)
270
11. Limit Theorems
for a given / e C'*(K'^) and R> 0, we can replace the given / by one which vanishes outside B(0, R + 1); and therefore, in view of (2.7), it suffices to show that for/eCS^(i?'^): (2.9)
Um sup - \ h\0
\x\
^
(f(y)-f(x))n,(x,dy) = 0.
'|j'-^l>i
But (2.9) is obvious, since / is bounded and we have just seen that (2.7) implies (2.4). Thus the proof is complete. D 11.2.2 Lemma. Assume that in addition to (2.4) and (2.5) we have sup sup(||a;,(x)|| + \bf,(x)\) < CO, h>0
xe Rd
and that instead of (2.6) we have lim sup Aft(x) = 0, h\0
8 > 0.
xeR
Let Xh -• XQ as h\0. Then the set {Pj^: h>0} is pre-compact and any limit point as h\0 solves the martingale problem for a and b starting from XQ. Proof L e t / e C^(R''). Then AHf(x)
|V/(x)|
I
(y- ^)n„(x, dy)
\y-x\
i m/dx, dxj ',7=1
2h
[
\y-x\'n,(x,dy)
ly-x\
1 2\\f\\-n,(x,R'\B(x,l)).
+
From this and our assumptions, we see that sup-|^,/(x)| 0
"
where Cf depends only on the bounds on / and its first two derivatives. Since (f(kh) - Y/)~ ^ ^/, f(x{jh)), Ji^y^, F^J is a discrete parameter martingale, we now see that
(f(kh)-^CfkKJ^,,,F
271
11.2. Convergence of Markov Chains to Diffusions
is a submartingale. Next note that P'U\x((k + l)h) - x(kh)I > £) = E''^in,(x(khl R'\B(x(kh), £))] < sup hAl(x). Thus, if r > 0, then X P'J\x(k^l)h)-x(kh)\
>£)<(r+l)supAUx)^0
0<jh
xeR^
as h\0. We can therefore use Theorem 1.4.11 to conclude that {P^^: /i > 0} is pre-compact. To complete the proof, suppose that h„\0 is a sequence such that P„ given by Pj^ with h = h„ converges to P. Clearly P(x(0) = XQ) = 1. Given / e C^iR"^), 0 < fi < t2, and
E'iifixilM)
-f(x(k„h„)) - II A,J{x{jh„m]
=0
where + 1
and
L=
+1
Note that by (2.7):
f(x(l„K)) -f(x(kM) - 'lA,J(x(jh„)) k„
^f{x{t,))-f{x(t,))-CLfix{u))du uniformly on compact subsets of il and boundedly. Thus, since if{x(h)) -f(x(t,))
- fLfixiu)) du)
is continuous, we conclude that E'[(f(x(t2)) -f(x(t,))
- CLf(x(u)) du)^] = 0.
That is, (f(x(t)) - jo Lf(x(u)) du, ^ , , P) is a martingale.
D
272
11. Limit Theorems
11.2.3 Theorem. Assume that in addition to being continuous the coefficients a and b have the property that for each x e R^ the martingale problem for a and b has exactly one solution P^ starting from x. Then Pj -• P^ as h\0 uniformly on compact subsets ofR'. Proof Let h„\0 and x„->Xoas w->oo, and set P" = P^;;. We must show that P"-^P^o as n->oo. For each /c > 1, let (p^ e Co(R^) be chosen so that 0 <(pk
n„,k(x, r) = (iPfc(x)n;,„(x, r) + (i - (pk(x))xr{xl and let {P"'': x e R^} be defined in terms of n„ ^(x, •) by the same procedure as Pj" is obtained from H^J^x, •). Take P"''' = P";*. It is easy to see that if a„^ ^(x) and b„ ^(x) are defined in terms of n„ jt(x, •) as a/,„(x) and /?/,„(x) are from Tlf^J^x, •), then for each k> 1: sup sup (||a„,fc(x)|| 4- |fc„,fc(x)|) < 00 and as n -• 00, an,k(x)^(Pk(xHx) and
uniformly on compacts. Thus, by Lemma n.2.2, {P"' *: n > 1} is pre-compact and any limit point solves the martingale problem for (^fc(*)a(*) and (pfc( • )Z?( •) starting from XQ. By Theorem lO.Ll, this means that any limit point of {P"-^: n> 1} equals P^^ on ^^^, where Tfc = inf{r>0: \x(t)\ >/c}, since (p^ = 1 on J5(0, k). At the same time, it is easy to see that P"* equals P" on J^,^ for all n. Thus, by Lemma 11.1.1, P„ -• P^o. D
1L3. Convergence of Diffusion Processes: Elliptic Case In this section we return to questions about convergence of the sort discussed in Section 1. Where our present treatment differs from the one that we gave there is that we now are going to be assuming that the matrix a is strictly positive definite valued. This assumption will allow us to weaken considerably the sense in which the drift coefficients (i.e. the first order coefficients) have to converge in order that
11.3. Convergence of Diffusion Processes: Elliptic Case
273
the associated diffusions converge. To help in understanding what is happening here, consider the following example. Let L" = ^A + YA= i Ki^^ ') ^/^^i» where b„: [0, 00) X R'^^ R^ is a bounded measurable function for each n > 1. If P" ^ is the solution to the martingale problem for L" starting from (s, x), then we know from Theorem 6.4,3 that for the bounded ^,-measurable O: Q -• C: where K(t) = exp p
'<M«, x(u)l dx(u)} - ^ j ' " \b„{u,
X(M))P du
In particular, if ^„ ^ ^ boundedly and pointwise, then, since
K(t)^R^(t) = Gxp j ' " < % , x(u)l dx(u)) - ^ f 1%, x(u))\' du in I}(i^^f^^); and so it is clear that P"^x^ Ps,x (^ variation) on ^j, where P^^ solves for L, = ^A 4- Yj=i b\t, •) d/dxi starting from (5, x). What is not so obvious is that we can weaken even further the sense in which b„ -• b. The point is that as long as the b„'s all share a common bound, it is quite simple to check (using the Cameron-Martin formula and the explicit expression for the transition probability function going with {i^s,x' (s, X)G [0, 00) x R*^}) that transition probability function P„(s, x; r, •) determined by L" = ^A + ^fi K(ty ') ^l^^i admits a density P„(s, x; r, •) and that {P„(s, x; r, •)}f is pre-compact in I}(R^) for each (s, x) e [0, 00) X R^ and that t> s. Thus, if {^„}f is uniformly bounded and 6„ -• ^ in the weak topology on U'(R% then for all cp e C^iR"^): T
T
lim [ dt [ L';(p(y)p„(s, x; r, y) dy = \ dt \ Lt(p(y)p(s, x; r, y) dy where L, = ^A + YA=I ^'(^» 0 ^/^^i and p(s, x; r, •) is defined in terms of L^. From this observation, one can easily conclude that the associated diffusions converge. The first one to point out that weak convergence of the b„'s to b guarantees convergence of the associated diffusions was M. Freidlin [1967], It is our aim in this section to carry out the proof of Freidlin's idea in considerable generality. Given 0 < A < A < 00 and a non-decreasing function S: (0, 00) -> (0, 00) such that lim^Xo (5(e) = 0, let ^/^(/l. A, S(')) be the set of all pairs {a, b} where a: [0, 00) X R'' -> Sj and b: [0, co) x R^ -^ R'^ are measurable functions satisfying: (3.1) (3.2) (3.3)
^e\^<(e,
a(s, x)e)
(s, x) e [O, oo) x R^ and 6 e R',
||a(5, x) - a(s, y)\\ <S(\x - y\), I b(s, x) I < A,
5 > 0 and x, y e R',
(5, x) e [0, 00) X R'.
274
11. Limit Theorems
For each {a, b} e s^(X, A, d( -)), the corresponding martingale problem is wellposed and the associated family of solutions {F^; J: (s, x) e [0, oo) x R*^} is strongly Markov and strongly Feller continuous (cf. Theorems 7.2.1 and 7.2.4). Let {a„, b„} e s^i{K A, (5(-)), « > 1, be a sequence such that for some {a, h} e s^i{X, A, ^(•)): (3.4)
lim I
I aj/(s, x)(p(5, x) ds dx a'^(s, x)(p(s, x) ds dx,
=
1 < U j < d,
and (3.5)
lim n^ao
b'„(s, x)(p(s, x) ds dx ^Rd
*'o
—
b'(s, x)(p(s, x) ds dx, •'O
i
^Rd
for all (p e Co([0, oo) x R"^). What we wish to show is that if (5„, x„) -• (s, x), then PT„lx"„ -^ Ptx- Notice that stating the sense in which a„-^a as we have in (3.4) is something of a hoax, since the fact that {{a„, b„}]f ^ ^d(>^, A, S( •)) already implies that for all 0 < s < T the family {jj' a^Ks, x)(p(s, x) ds}f is pre-compact with respect to uniform convergence in x. Thus (3.4) should be viewed as simply a method of identifying a as the limit of the a„'s rather than really describing the manner in which the limit is taken. 11.3.1 Lemma. Let cp e C'S(R'^) be a non-negative function such that ^Ri (p(x) dx= 1. For g e B(R% define g,(s, x) = J (p,(x - y)g(t, y) dy where (p^[x) = s~^(p(xl&). Then for each 7 > 0,
1+ 9
J (g - ge)(t, x(t)) dt
tends to zero as s\0 at a rate which is independent of {a, b} e s^i{K A, ^(•)), (5, x) e [0, r ] X R^, and g e B([0, 00) x R'^). Proof. For {a, b} e ^^(A, A, S(*)), let p^^ j,(s, x;t, y) be the associated density function described in Theorem 9.1.9. From Theorem 9.1.15, we know that: T
\ dt \ \p^,5(5, x; t, y + h) - Pa,^(s, x;t,y)\dy '^s
''Rd
tends to 0 as | /i | -• 0 at a rate which is independent of (s, x) e [0, T] x R** and {fl, 6}G j/d(/l, A, <5(-)). Since
275
11.3. Convergence of Diffusion Processes: Elliptic Case
dt \ p^, b(5, x; r, y){g - g,){u y) dy\ < [ (p,(z) dz\ [ dt \ {Pa,b{s, Pa,t(s,x; t, <\\g\\
sup
y))g{t,y)dy
\ dt \
z e supp((pe)
x',Uy-z)
s
\Pa,b(s,x;t,y-z)
Rd
- Pa,b(s.x; t,y)\ dy, it follows that I dt
1 + ^11•^s
p^^t,(s, x; t, y){g - g,){t, y) dy\ ''Rd
1
tends to 0 as e \ 0 at a rate which does not depend on {a, b} e s^^{K A, <5(')), (s, x) G [0, T] X R\ or ge B([0, oo] x R'). But £^?J H
{g - g,)(t, x(t)) dtj T
\ {g - ge){t, x{t)) dt \ (g- g,){u, x(u)) du I ^
<4||sf|| [ s
p^^^(uz;u,y){g-g,){u,y)dy\ dt
sup f du\ 2 e Rd\ t
^Rd
<4(T-s)||^||
sup it, z)e[s,
T]xRd
T
I du I p^,^(r, z; u, y){g - g,){u, y) dy\ U
-^Rd
and so 7P'.i
1 + ii^ii
J {g - gs)(t, x{t)) dt <
(T-s)
T
1/2
rF?'5
1 +M
<2(T-s)l
sup \ ^ "'" 11^11
I ^
\2
{g - ge){t, x{t)) dt
it,z)e[s,T]xRd I
X I du \ p^^i,{t, z; u, y){g - g,){u, y) du\ tends to 0 as e \ 0 at a rate which is independent of {a, b} e s^i{K A, <5( •)), (s, x) e [0, r ] X K\ and gf € B([0, 7] x R''). D
276
11. Limit Theorems
11.3.2 Lemma. Let {a, b} e s^^{X, A, (5(-)) and (s„, x„) e [0, oo) x R'', n> 1. Set P" = p«^"'J|| and suppose that s„^So and P"-* P as n-*oo. Let {g„}T ^ B([0, oo] X R^) be a uniformly bounded sequence such that for some g e B([0, oo] X R^) and all cp e C^([0, oo) x R''): 00
QO
lim n-oo
g„(s, x)(p(s, x) ds dx = \ *'0
^Rd
•'O
g(s, x)(p(s, x) ds dx. ^Rd
Then for all SQ < t^ < ti and bounded continuous J^t ^-measurable O: Q -• K^;
(t>\'\(t.x(t))dt
= lim E^" 0 ) ]
g„(t,x(t))dt
Proof If the ^„'s were continuous in x and converged uniformly in x to ^ for each r, there would be no problem. Thus we introduce Sf„. e(s, x) = I (p,(x - y)g„(s, y) dy and g,(s, x) = j.(p,{x - y)g(s, y) dy. It is easy to check that for each £ > 0,
g„^,(t,x)dt^\
g,(t,x)dt
boundedly and uniformly with respect to x. Thus, by Corollary 1.1.5,
Oj
g,(t,x(t))dt
lim E^" O j
g„^,(t,x(t))dt
It is therefore enough for us to show that
lim sup lim sup e\0
n-»oo
d) ^\„(t, x(t)) dt\ - £''"U j"ff„.,(t, x(t)) dt
277
11.3. Convergence of Diffusion Processes: Elliptic Case
To prove the first of these, choose {h„]f ^ Q([0, oo) x R'^) such that \\h^ \\ < \\g\ h,^g in L'([t„t2] xR'*), and Hm£^ <» [ \ ( f , X(f)) df
fc-oo
.
^n
:
Then, by Lemma 11.3.1 -.'2
<^[(g-
ge)(t. x(t)) dt — Hm
*'ri
= lim Hm £^f(DJ'"\-V.)(r,x(r))^f < ||0|| sup sup fe>l n>\1
L
LMs„vfi
IJJ
as e \ 0 . To prove the second, observe that: J2
hm sup lim sup f. \ 0
ipn
^]
(gn-9n,e)(t,4t))dt
M - • 00 ,5„vr2
= lim sup hm sup lE^" O I F.\0
n -^ CO
I
L
(g„- g„^ ,)(t, x(t)) dt
5nvri
5 lim sup sup||
n>l
L
r'\gn-gn^){tMt))dt\]]
LUsnVri
IJJ
= 0, again by Lemma 11.3.1. D 1L3.3 Theorem. For each n > 1, let {a„, b„} e s/j(X, A, S{')) and (s„, x„) e [0, cc) X R^. Assume that (s„, x„)-> (SQ, XQ) and that there is an {a, b} e s^i{X, A, (5(-)) such that (SA) and (3>.5) hold. Then PJ-J;; -> P%^^^ asn^cc. Proof. Let P" = F^„"; J;;. By Theorem 1.4.6, {P":n> 1} is pre-compact. Thus, all that we need to show is that if {P"} is a subsequence of {P"} which converges to P, then P solves the martingale problem for a and b starting from (5o,Xo). Clearly P(x(f) = Xo, 0 < f < So) = 1. To complete the proof, let SQ < fi < ^2 ^^^ O: Q-*" P^ be bounded, continuous, and .#fj-measurable. Given fe Co'(P''), we must show that
= 0,
278
11. Limit Theorems
where
'•H„|,"''<''%4/,?.''•'•-'s;Since ,'2 VS„'
W{x(t2^s„.))
- / ( x ( t , vs„.)) -
Lrf{x{t))
dt)
for all n\ it is clear that we need only show that Jl'y Sn
0)
Lr'/(x(f)) dt
lim E''"
Oj
tf(x(t)) dt
But this last fact is an immediate consequence of (3.4), (3.5), and Lemma 11.3.2. Thus the theorem is proved. D It should be clear that the conclusion of Theorem 11.3.3 ought to remain true even after we replace our assumptions by their local analogues. To be precise, suppose that R^ ^R and /? ->• A^ are functions on (0, oo) such that 0 < A^ < A^^ for all R, and let K ->• ^^(•) be a mapping of (0, oo) into non-increasing functions from (0, oo) -> (0, cx)) such that lim^ o ^R(^) = 0 for all R > 0. We will say that the pair {a, b} is an element of ^d'^'i^R, /^R, SR(')) if a: [0, oo) x R'^-* S^ and b: [0, 00) X R'^ -^ R"^ are measurable functions satisfying: for K > 0, (s, x\ (s, y) 6 [0, R] X B(0, Rl and 0 e R^ (3.6)
A J 0 | 2 < < ^ , a(s,x)^>
(3.7)
|a(5, x) - a(5, y)||
(3.8)
<SR(\x-y\),
^(5, X) I < A^ ,
and the martingale problem for a and b is well-posed. Note that, given (3.6), (3.7) and (3.8), the martingale problem for a and b starting from (s, x) has at most one solution (cf. Theorem 7.3.1). Thus the martingale problem for a and b is wellposed if solutions exist at all; and according to Corollary 10.1.2, existence is equivalent to non-explosion. Finally, if {a, b} e ^O'^^R ^ ^R ^ ^R( *))»then the associated family of solutions is strongly Markovian and strongly Feller continuous (cf. Corollary 10.1.4). 11.3.4 Theorem. For each n> 1, let [an^K es^r(^R,AR,SR('))and(s„,x„)e [0, 00) X R**. Assume that (s„, x„) -^ (SQ, XQ) and suppose that there exists {a, b} e ^T(x 'R-> A ; j , dR\ ))for which (3.4) and (3.5) hold. IfP" is the solution to the martingale problem for an and b„ starting from (s, x„) and P is the solution for a and b starting from (SQ , XQ), then P" -^ P.
11.4. Convergence of Transition Probability Densities
279
Proof. We will reduce the present case to the situation in Theorem 11.3.3. For each k > sup„ (|x„| + 5„), let rjk e Co([0, oo) x R*^) be chosen so that 0 < ^^ < 1 and ?/fc = 1 on [0, k) X B(0, k). Define a„ ^ = ^k«n + (1 - ^kV ^^^ ^n,k = ^k^n for w > 1, and set a^ = ^/fc ^ + (1 - ^kV ^^^ ^k = ^k ^- For each /c > 1, {{a„ ^, h^fc}}^=i and {flfc, 6 J satisfy the conditions of Theorem 11.3.3. Thus, if P " ' ' = P^'Xt'^'' ^^^ P^^PToX^ then p»^^-^P^ as «-^oo. Finally, if T^ = inf {f > 0:"f >/c or \x(t)\>k\ then F"''' equals P"(x).,(.)PS.''),x"(V\(.), ) and P^ equals P0,^(.)P?J(.^%(,^.),.,. Thus P" equals P"''' on ^ , ^ and P equals P'' on . # , , . By Lemma 11.1.1, we therefore know that P" -^ P. D
11.4. Convergence of Transition Probability Densities In the preceding sections we have seen that solutions to the martingale problem vary continuously with the coefficients and starting place. The type of convergence which we have imposed on the coefficients is different in Section 3 from that in Section 1, but the sense in which the associated solutions converge is the one dictated by the weak topology on M(Q). It is not hard to see that the weak topology is the strongest one which we can afford in general, since, even if the starting point is held fixed, solutions corresponding to different second order coefficients will be singular. Nonetheless, if we do not look at the solution itself but only at the marginal distributions derived from it, we can sometimes do better. In this section we will illustrate this point by looking at the transition probability function for time-homogeneous coefficients which satisfy the conditions of the preceding section. Given 0 < A < A and a non-decreasing function S: (0, oo) -> (0, oo) such that lim^xo ^(fi) = 0, let J f ^(A, A, S(')) stand for the class of {a, b} e s^ j^X, A, (5(*)) such that a and h are time-independent. For {a, b] e Jfd(A, A, ^(•)), we know that the martingale problem is well-posed and that the associated family of solutions {PJ J: (5, x) e [0, oo) X K^} is time-homogeneous in the sense that P%\\ = PJ'' o 0 ~ \ where PJ' ^ = P?;^^ and 0),: Q -^ Q is defined by x(r, 0,(ca)) = x((f - s) vO, co), t > 0. Furthermore, we saw in Section 9.2 that the time-homogeneous transition probability function P^ ^,(5, x, •) admits a density p^ ^(s, x, y) with respect to Lebesgue measure and that Pa,b(^^ x, -) e n{R'^) for all 1 < p < 00. What we are going to prove now is that Pa„,&„(s, ^, ')-* Pa,b{s^ x, •) in 1I(R'^\ 1 < p < 00, if {a„, b„} -* {a, b} in the sense of the preceding section. 11.4.1 Lemma. Let {f„}f be a sequence of non-negative ^ j^d-measur able functions such that J /„(x) Jx = 1 for all n>\ and lim|;,|_o sup„>i J |/„(x -\- h) fn(x) I Jx = 0. Assume that there is anfe I}(R^) such that [ /(x)iA(x) dx = lim [ f„(x)\j/(x) dx for all e C,{R'). Thenf„ ^fin
L'iR').
280
11. Limit Theorems
Proof. First observe that for any £ > 0 we can find an R^ such that sup f
I Ux) -f(x) I dx < sup [ £
e
2
2
I f„(x) \dx+\
I fix) I dx
Thus, we need only show that/„ -•/in L^(B(0, R)) for each R > 0. To this end, let (p e Co(R'^) be chosen so that (p > 0, (p( — x) = (p(x), and J (p(x) dx = 1. Define (p^{x) = e~^(p{xle) for e > 0 and set/„ e = ^t * L and/ = (p^ * f Given ij/ e Q(/?'^), we have: f / . .(^)'A(X) dx = j /„(x)lAe(x) ^X - . J /(X).A.(^) ^X
where il/e = 9e*^' Moreover, it is easy to check that sup„>i {||/„ ^ || + \\Wn,e \\) < ^ for each € > 0. Thus by the Arzela-AscoH theorem, f„^^-^fe uniformly on compacts for each e > 0. Given /^ > 0, we now have: limsupf n-»oo
\fnM-f(x)\ dx B{0,R)
\fn(x)-f„,e(x)\
dx -h [
''B{0,R)
\f{x)-f(x)\
dx
B(0,R)
for all £ > 0. But
j|/W-/(x)Mx-.0 as £\0, and sup f I f„(x) - / „ , ,(x) I dx = sup [ dx I (/„(x) -/„(x - y))(p,(y) dy \ « > 1 *^
n > l -^
< sup f (p,(y) dy \ I f„(x + >;) -/„(x) | dx < sup
sup j I /„(x + y) -f(x) \ dx-^0
«> 1 ye s\ipp{(pt)
as £ \ 0 . This completes the proof. G
11.4. Convergence of Transition Probability Densities
281
11.4.2 Theorem. For each « > 1, let {a„, b„} e Jt^i^, A, d(')) and suppose that {a, b} e Jfj(X, A, 3(')) satisfies:
lim I a'J(x)(p(x) dx = I a'^{x)(p(x) dx,
1 < i, j < d,
lim I b'„(x)(p(x) dx = j b'(x)(p(x) dx,
I < i < d,
n-+oo
for all cp e CJ(R^). Finally, assume that {(s„, x„)}f ^ (0, oo) x R"^ tends to (SQ, XQ) e (0, oo) X R^. Then, for each 1 < p < oo, Pa„,b„(Sr,, x„, ')-^Pa,b(so^ ^o^ ') in U(R'). Proof By Theorem 11.3.3, we know that
I «A(y)Pa. b(5o. >^o>y) dy = lim I iA(>'K„.b>„, x„, >') dy
for all lA e Q(K''). Moreover, by Theorem 9.2.12,
lim sup
|Pa„, J 5 „ , x„, y + /i) - /?«„. J 5 „ , x„, y)| dy = 0.
| / J | \ 0 n > 1 •'
Thus, by Lemma 11.4.1, p«„, J s „ , x„, •)->p^,^(^o. ^o. *) in L^(K'^). Since, by Theorem 9.2.6, we also know that sup„>i j \Pa„,b„(Sn^ ^n^ y)\'^ dy and J I Pa, bi^o ^ ^0».v) 1^ dy are finite for all 1 < p < oo, an easy application of Hdlder's inequality completes the proof. • Finally, we give the localized version of the preceding. Recall the definition of ^d'^'i^R^ A^, 3i^(')) given in Section 11.3, and denote by J^T(^R^ A ^ , ^^(•)) the subset of ^^d^'^i^R, AR , Sn(')) consisting of time independent elements. It is easy to see from Corollary 10.1.5 that is {a, b} e ^Y{}-R^ ^R^ ^R{*)\ then the associated family {?l\\'. (s, x) G [0, oo) x K^] of solutions to the martingale problems forms a strongly Feller continuous, time-homogeneous Markov family. Moreover, by Corollary 10.1.5, we know that the time-homogeneous transition probability function P^fc(s,X, •) admits a density p^ ^,{s, x, y) for 5 > 0 such that p^ ^(s, x, •) e E ^""'{R^) for all 1 < p < 00. Finally, for any 1 < p < oo, 0 < r < R, and (5 > 0,
(4.1)
sup s>d \x\
sup {a, b\ € J¥dloc(XR, \R, dR{-))
j B{0, R)
Ip^,b(5, x,y)\^
dy
282
11. Limit Theorems
11.4.3 Theorem. For each n > 1, let {a„, b„} e :^^r{^R ^ ^R ^ ^R{ ')) «"^ suppose that there is an {a, b} e J^a^'i^^, A^^, SR(')) such that I a'\x)(p(x) dx — lim
al;l(x)(p(x) dx,
1 < i, j < d,
and [ b'(x)(p(x) dx = lim [ bl(x)(p(x) dx,
I < i < d,
n->oo
for all (peCS(R'y If {(s„, x„)}f ^ (0, co) x R' and (5„, x J - > (so, Xo) e (0, o)) X R', then Pa„,b„(Sn. ^n, ')-^Pa,b(so, ^0, *) in n(R') and in H'"'%R')for each 1 < p < 00. Proof In view of (4.1), it suffices to prove that p^^, b„(5„, x„, ')-^ Pa, bi^o» ^o» *) i^ 1}(R'^). We will do this by reducing this case to the one just handled in Theorem 11.4.2. For /c > 1, choose //^ e C^iR"^) so that 0 < yy^ < 1 and rj^ = I on B(0, k). Set
and ^n,k
—
^kbn,k'
Clearly, for each /c > 1, Theorem 11.4.2 applies to {{a„j,, b„j^}: n> 1} and «k = ^fc« + (1 -rjkV and ^/c =
rjkb.
Thus, the proof will be complete once we show that lim sup [ \Pa„,,bJs„, x„,y)-
Pa„,bn{^n^ ^n^y)\
dy = 0.
k / o o n> 1 ''
But
where i^, = inf{r > 0: |x{t) \ > k}. Since FJ; ^" -^ P^''' and lim,/^ ?;'"(!, <sup„^is„) = 0, it is easy to conclude that l i m , / ^ sup„< ^ PS";^;;(T;£ < s„) = 0.
Q
11.5. Exercises
283
11.5. Exercises 11.5.1. Let a: [0, oo) X K^ ^ 5d and /?: [0, oo) X R*^ -^ R"^ be locally bounded measurable coefficients. Assume that the martingale problem for a, b is well-posed in that for each starting point (s, x) there is exactly one solution to the martingale problem corresponding a, b starting from (s, x). Let (5^, x°) e [0, 00) x R'^ be given and let G be any open set in [0, 00) x R"^ containing (5^, x^). Define T = {inf t>s^: (r, x(t)) ^ G}. If P is any measure on (Q, .///) such that P[x(t) = x^ for 0 < r < s^] = 1 and
f(x(TAt))-(^\LJ)(x(u))du is a (Q, , # , , P) martingale for t > s^, then show that P = Pso^^o on . # , where F50 xo is the unique solution to the martingale problem for a, b starting from 11.5.2. Let a:R^ -^ Sd and b:R^ ^ R^ he locally bounded measurable coefficients such that the martingale problem for the time-homogeneous coefficients a, b is well posed and admits a family {P^} of solutions starting from x which forms a continuous family, i.e. the mapping x -^ Px is weakly continuous. If F e ^0+ = nr>o ^t then show that for each x G JR'^, Fx(P) = 0 or 1. This property is known as Blumenthal's zero-one law, and can be estabhshed as follows. Let y) be a bounded continuous map from (R^)^ into R. Let IM > tM-i " • > ti > 0. Consider ^(co) = xp(x{ti),...,x(tM)) mapping Q —• R. For small e, since F e Jt^, the Markov property implies / <^{(D)dPx = J £^^<^) [xp(x{h - £ ) , . . . , x{tM - £))]^Fx . Fixing f 1, ..., r^v ^nd letting a -^ 0 conclude that cl)(w) dP, = P,(F)£'''[(I>(«)]. Now conclude that F is independent of itself and therefore PJ
284
11. Limit Theorems
converges to a suitable constant coefficient diffusion. It will now follow that lim^^Q P^o[(p{x(t))> 0] = ^. Use Exercise 11.5.2 to conclude that P,„[T'(a)) = 0 ] = l . 11.5.4. Using the same notation as in Exercise 11.5.3, let T(CO) = inf [r > 0:x{t) ^ G] and conclude that for each x G G, Px[ti(o) < oo, T'(CO) > T(CO)] = 0. Hint: Use P^[T(CO) < 00, T'((O) > T(CO)] = £^^[F^(,)[T'(CO) > 0], T(CO) <
oo].
11.5.5. Using the same notation, assume in addition that G is bounded. By Exercise 10.3.1, x((jo) is finite ahnost surely with respect to P^ for each x e G. Use Exercise 11.5.4 to conclude that T{(O) is almost surely continuous with respect to P^. Hint: Show that T(CO) is lower semi-continuous but T'(CO) is upper semi-continuous. 11.5.6. If/is a bounded continuous function on dG then u(x) = E^'{f(x(z))] is the solution to Dirichlet's problem, i.e., Lu = 0 in G (in some generalized sense if L does not have smooth coefficients) and w = / on dG. Under the same assumptions as in Exercise 11.5.5, prove that u{x) is continuous in x on G. Also, if PJP^ are diff'usions approaching Px in the sense that Pj^^ —• Px weakly as n —• oo, then lim„_oo£:^^^"^l/(x(T))] = £ ^ ^ | / ( X ( T ) ) ] .
Chapter 12
The Non-unique Case
12.0 Introduction The last few chapters have been devoted to finding out when the martingale problem is well-posed and what can be said when it is. In this chapter, we are going to examine whether anything useful can still be proved when we drop the assumption of uniqueness. In particular, we want to show that even in this case it may be possible to " select" a family of solutions in such a way that the strong Markov property prevails. We also want to describe the structure of the set of solutions associated with given coefficients and a given initial condition. The idea behind our procedure for selecting strongly Markovian versions is the following. Given coefficients a and 6, let ^(s, x) denote the set of solutions to the martingale problem for a and b starting from (s, x). Assuming, as we will be, that a and b are bounded and continuous, it is easy to check that ^{s, x) is a (weakly) compact, convex set in M(Q), the set of probability measures on Q. Thus, when we try to make a nice selection, it is natural to attempt isolating a unique point from each ^(s, x) by insisting that it be that element of ^(s, x) which is characterized by some extremal property. Moreover, since we would Hke the resulting selection to fit together in a Markovian way, the extremal property that we choose should be preserved under conditioning. Finally, we would like to make the selection as well-behaved as possible with respect to changes in (5, x), and so the extremal property should be one which behaves well under weak convergence. That such a program exists and yields the desired results was realized by N. Krylov [1973]. Our treatment does not differ substantially from his. As for the structure of ^^(s, x), what we want to show is that every element of ^'(s, x) can be obtained by "piecing together" elements of ^(s, x) which are members of strongly Markovian selections. What we have in mind is made precise in Section 12.3. One final comment. We are going to assume that the coefficients a and h are independent of time. However, this is not really a restriction, since by considering the "time-space" process, one can easily handle coefficients which are timedependent as if they were time-independent. See Exercise 12.4.1 below for more details on this point.
286
12. The Non-unique Case
12.1. Existence of Measurable Choices Before we can prove the main theorems of this chapter we need some facts about the space of compact subsets of a separable metric space. This section is a digression in that direction. Among other things, we will prove some results about the possibihty of choosing measurable selections from set-valued maps. Let X be a separable metric space with a metric d. We denote by comp(X) the space of all compact subsets of X and define a metric p(Ki, K2) between two points Xj, K2 of comp(X) by (1.1)
p(Ki,K2) = M[8>0:Ki^K'2
and
K2^K\].
Here, for any set A. e X, A" denotes A'- = {y: d(x, y) < e for some x e A}\ in other words, A" is the sphere around A of radius B. It is easily verified that p is a metric on comp(A'). Moreover if two points x, y of X are viewed as the single point sets {x} and {y} in comp(X), then p({x},{y}) = d(x,y). That is to say, X is embedded isometrically in comp(X). We will state a series of lemmas about the metric space (comp(X), p). 12.1.1 Lemma. (comp(JV), p) is a separable metric space. Proof. Let XQ a X bea. countable dense subset of X. One can check easily that the class of all finite subsets of XQ is a countable dense subset of comp(X). D 12.1.2 Lemma. IfC^X is closed in X then {K: K ^ C} is closed in comp{X). If G ^ X is open in X then {K: K ^ G} is open in comp(X). Proof. This is obvious and is left as an exercise. D For any set A G X, we denote by J(A) the set {K: K ^ A} in comp(X). 12.L3 Lemma. IfF^X is any finite set, then the ball S(F,s) of radius s around F in comp{X) is given by S{F. f.) = J{F) n I n [J{B{x, £ r ) ] | Here B(x,e) is the ball in X of radius e around x.
12.1. Existence of Measurable Choices
287
Proof. Clearly K e S(F, e)op(K, oK
F) < s
^ F^
<=>K ^ F'
and and
F c K^
for some
e' < £
F ^ K\
Now X cz FMf and only if X e J(F'). On the other hand, F ^ K^ox e K^ oK
for each
x eF
n B(x, e)i= (j) for each
oK ^ J(B(x, ef) oK
6 [J(B(x, sy)Y
oKe
for each for each
f] [J(B(x, 8Y)Y.
x eF xe F xe F
D
X€F
12.1.4 Lemma. The class of sets of the form J (A) where A runs over all the sets A in X which are either closed or open in X, generates the Borel o-field of comp(X). Proof It is clear from Lemma 12.1.3 that the cr-field generated by J(A) as A runs over the class of sets that are either open or closed in X contains sets of the form S(F, fi) for any £ > 0 and any finite set F. The class of sets of the form 5(F, £) as F runs over finite subsets of X and £ runs over all positive numbers is clearly a basis for the topology of comp(A'). Any c-field that contains a basis in a separable metric space must contain the Borel cr-field. G 12.1.5 Lemma. IfG ^ X is open, then J{G) is a countable union of the form |J„ J(C„) with C„ closed in X. If C ^ X is closed then J{C) is a countable intersection of the form P)„ J{G„) with G„ open in X. Therefore, either one of the classes {J(A): A open in X} or {J(A): A closed in X} generates the Borel o-field of comp(X). Proof If G £ X is open we define
C„=jx:fi(.x,;)^G]. Then each C„ is closed and \J^ Cn = G. Moreover, since K is compact, it is easy to check that, if X c: G, then K a Cn for some n. Therefore J(G) = n J(C„).
288
12. The Non -unique Case
On the other hand, if we write C = f] G„ with
G„ open in
X,
n
then J(C)=f]J(G„y
D
12.1.6 Lemma. Let A a X be a closed subset ofX. Let S = {K: K n A i- 0]. Then i is a closed subset of comp(X) and the map K-* K n A is a Borel map ofS into comp{X). Proof. Since S = [J(^'')]^ it is closed. In order to prove the measurability of the map K-^ K n A, it is enough to show that for any G ^ X which is open in X, {K: K n A ^ G] is a. Borel subset of comp(X). Clearly K n A ^ G if and only if K^Gu A'. Therefore {K: Kn A^G} is the same as {K: X c G u A'} = J(G u A% and J(G u A^) is open in comp(X). D 12.1.7 Lemma. Let f(x) be a real valued upper semi-continuous function on X. For K £ comp(X), set JK = sup/(x), xeK
and define f : comp{X) -^ comp(X) by f{K) = {y:yGK Then the maps K -^fi^ and K -^f(K) comp(X), respectively.
Proof.
and
f{y)=fK}-
are Borel maps of comp(X) into R and
{K:f^ < 1} =\K: sup f(x)< I] \.
xeK
=
I
{K:K^{x:f{x)
= AG,) where G/ = {x:f{x) < 1} is open in X. J(G/) is therefore open in comp(X) and therefore X - • / ^ is a Borel map. On the other hand, if G £ X is open, then {K:f{K)
^G} = {K: K^G}u = J(G) u
| x : K n G'=/= 0,
{K:KnG'+0]n
sup f(x) < sup >K:f,^,,
f(x)\
12.1. Existence of Measurable Choices
289
By Lemma 12.1.6, K-^ K n GMs a Borel map of comp(X) into itself on {K: K n G*" i^ 0 } . From the first part of the lemma, K - • / K is a Borel map of comp(X) into K. Therefore {K\f(K) ^ G} is a Borel set in comp(X) and we are done, n 12.1.8 Lemma. Let Y he a metric space and ^ its Borel a-field. Let y^Kybea map of Y into comp(X)for some separable metric space X. Suppose for any sequence y„ -• y and x„ e Ky^, it is true that x„ has a limit point x in Ky. Then the map y ^ Ky is a Borel map of Y into comp(X). Proof. Let G £ X be open. We will show that {y:Ky ^ G] is open in 7 ; or equivalently that {yiKynG"" =^ 0 } is closed in Y. Let Ky„ n G"^ ^ 0 for any n and y„ -> y. Choose x„ e Ky^ n G^. By our assumption, x„ will have a limit point x which will be in Ky. But, since G*" is closed, x G G*" as well. Therefore Ky n G" is non-empty and we are done. D 12.1.9 Lemma. Let (£, J^) be a measurable space. Let q -• h(q) and q-^ K^ be measurable maps ofE into some separable metric space X andcomp(X), respectively. Then the set {q: h(q) e X j is a measurable subset of E. Proof. It is enough to verify that the set of pairs (x, K) in X x comp(A') with X e K is a Borel subset of the product space. But it is in fact closed, because if x„ -^ X and K„-^ K with x„e K„, it follows from the definition of p(K, K„) that d(x„, K) -• 0. Therefore x e K. D We will complete this section with a theorem which will be important for us in the rest of this chapter. 12.1.10 Theorem. Let (£, ^) be any measurable space and q-^ K^a measurable map ofE into comp(X), where X is some separable metric space. Then there is a measurable map q -• h(q) of E into X such that h(q) e K^for every q e E. Proof. Let / i , / 2 , . . . , / „ , . . . be a countable sequence of continuous functions on X which separate points in X. Take Kf^ = Kq and define inductively
(The notation here is that of Lemma 12.1.7.) By Lemma 12.1.7, for each n,q-^ Kf^ is a measurable map of E into comp(X). Moreover for each q.K^^^ is a nonincreasing sequence of non-empty compact subsets of X. Therefore K'^ = fin ^^"^ is non-empty and one can verify that X ^ is the limit of X^"^ as w -• oo, in the space comp(X). Therefore the map q^ K^ is also a measurable map of £ into comp(X). We will complete the prooof by showing that for each q, K^ consists of exactly one point of X. If we denote that point by h(q), then q -> h(q) will be a measurable map of E into X (remember that X is embedded isometrically in comp(X) and h(q) e Kq for each q).
290
12. The Non-unique Case
To see that K^ consists of exactly one point, we reason as follows. By our construction,/„(x) is constant on K^^K Since K^ is contained in Kf^ for every n, all the functions/„(x) are constant on K'^. Since/i,/2, ... separate points in X, it follows that X^ can have at most one point for each q. But K^ is non-empty for each q and must therefore consist of exactly one point. This completes the proof, n
12.2. Markov Selections We are now going to adapt the preceding results to the situation arising in the study of the martingale problem. For each (s, x) e [0, oo) x R"^, suppose that ^(s, x) is a non-empty compact subset of M(Q), the space of probability measures on Q, and assume that the family {^{s,x):{s,x) G [0,oo) x R"^} satisfies the following conditions: (a) (s, x) -* ^(s, x) is measurable on [0, oo) x R*^ into comp(M(Q)), (b) P e ^ (0, x) if and only if P
x{t, (^sco) = x{(t - s) V 0, co);
r > 0,
(c) if P G ^(0, x), T: Q ^ [0, oo) is a finite stopping time, and {Pj is a r.c.p.d. of P|./#,, then there is a P-nuU set N e . # , such that ^x(t(co),to)®T(co)^a) ^ ^(z((o), X(T(CO), CO)) for each co ^ N, (d) if P e ^(0, X), T is a finite stopping time, and co -^ Q^ is an ^^-measurable map such that S^^,^^^ ^^^ ,(^^Q^E^(X(O)1 X(T(CO), CO)) for all co, then
P®,e.e<
291
12.2. Markov Selections
12.2.2 Lemma. Let {^(s, x): (s, x ) e [0, oo) x R"^} satisfy (a) through (d) of (2.1). Given A > 0 and a hounded upper semi-continuous function f: R** ->• jR, define u(s, x) =
sup E'
\
e-'f(x(t-hs))dt
.0
P e "^{s, X)
Then u{s, x) = w(0, x)for all s >0 and x e R^. For each (s, x), set r(s, x) = [P e ^(5, x): E'
+ s))dt = u(s, x),
re-'r(x(t
Then ^'(s, x)e comp{M(Q)) for all (5, x) and the family {^'(s, x): (s, [0, 00) X R^} again satisfies (a)-(ci) of (2.1).
x)e
Proof That u(s, x) = w(0, x) is obvious from (b) for {^(s, x): (s, x) e [0, 00) x R^}. Each ^'(s,x) is in comp(M(Q)) and (s, x)->> ^'(5, x) is measurable because of Lemma 12.1.7 and the easily verified fact that e-'f(x(t
(s, P) -^ E'
+ s))dt
is upper-semicontinuous. Condition (b) for {^'(s, x): (s, X)G [0, 00) x R*^} is immediate from (b) for {^€(s,x)\ (s, X)G [0,00) x jR**} plus the identity u(-, x) = M(0, X). To prove that (c) is satisfied, let P e ^'(0, x^) and a stopping time i be given. Suppose {Pf^} is an r.c.p.d. of P \ M^ and define N = {OJ: P^ ^ ^(1(0;), X(T(CO), OJ))}
and A = {oje N': P^ i T(z(co\
X(T(CO), CO))}.
By (c) for {^(s,x): (5,x) G [0,oo) x R^}, N e ^r and P(N) = 0. Moreover, NUA = {coiPoj ^ ^'(T:(CO),X(T(CO),(O))}, and therefore, by Lemma 12.1.9, N u A G .£^. Thus .4 G ^#,. We want to show that P(A) = 0. To this end, use Theorem 12.1.10 to choose a measurable map (s, x)->- R^.x so that R^^ e ^'(5, x) for all (5, x). Then co -> R^ given by 1^0)
—
^O) ® T ( c o ) - ^ T ( c a ) , JC(Tr(cj),
co)
is ./#^-measurable. Define jR„
if (oe N u A
\P^
if
CO ^ iV u ^ .
292
12. The Non-unique Case
By (d) for {^(5, x): (s, x)e [0, 00) x R% Q = P(S),Q.e ^(0, x^). Thus w(0, x^)>£^ f"e-Y(x(r))Jr i*'o
At£Q.
+ £''
e-''/(x(r))rft + £' g-At£P. Z' / X, -
AT
I
rR
fe-"f{x{t
+ T))dt
j e-^'/(x(f + T)) dt
-E'
Nt//1
= u(0, x° + E' e-"\u{x,x(x))-E'
Jr
Since
w(0, x^) = E^
e-r{x(t))dt \e-''fMt))dt
+ £' £ - ' • ' £ '
^f
and P(N) = 0. Thus ^^(M(T, X(T)) -
£^ j'"e-Y(^(f + T)) ^f I
<0.
On the other hand, e-'YMt-\-T))dt
< W(T(CO), x(T(a;), o)))
'0
for (JOE A, and so we conclude that P(A) = 0.
,A
293
12.2. Markov Selections
Finally, we must show that {^'(5, x): (5, x) e [0, 00) x R*^} satisfies (d) of (2.1). For this purpose, let P e ^'(0, x^) and suppose co-^ Q^ is an ^^-measurable map such that ^x(r(co).CO) (8)t(co)60 e ^'(T(CL>), X(X(CO), (O)) for all co. Set Q = P(x),g.. By (d) for {^(s, x): (s, x) e [0, 00) x R% Q e ^(0, x^). Moreover,
w(0, x^) > £^ 0
-^rj^Q.
= E' >E^
e-''f(x(t))dt
+ £''[e-'"u{T,
X(T))]
\e-'r(x(t))dt
+ £'
^ - xATirP r£
j ^-^:^(x(r + T))^f
= E' =
M(0, X"),
since, by (c) for {'^(s, x): (s, x) e [0, oo) x R''j M(T, X(T))
> £'
We have therefore shown that Q e ^'(0, x^).
^r
(a.s., P).
Q
We are now ready to prove the main theorem of this section. 12.2.3 Theorem. Let a: P'^-^S^ and b: P'^-^P'* be bounded continuous functions. Given A > 0 and a bounded upper semi-continuous function f: R^ -^ R^, there is a measurable map (s, x) -> P^^x of[0, co) x R^ into M(Q) such that (0 ^s.x solves the martingale problem for a and b starting from (s, x), is the map described in b) of (2.1), (a) P, ^ = P Q ^ : (5>~ ^ 5 > 0, where ^,:Q-^Q (Hi) I/T: Q -• [s, 00) 15 a stopping time and {P^^} is a r.c.p.d. ofP^^ ^ I -^t»then there is a P-null set N e Ji, such that P^ = ^c,®r(co)^r(co).x(t(co).co)/o^ «/^ co ^ N, (iv) for each (5, x), £^^ I j ? e-'f(x(t + s)) c/r] maximizes £''[J? £-''/(x(f + s)) t/f] as P Dar/^5 oi;er solutions to the martingale problem for a and b starting from {s, x).
294
12. The Non-unique Case
Proof. Let {(7„}f be a dense subset of (0, oo) and {(p„}T a dense subset of Co(R'^% and let {(A^,//v): N > 1} be an enumeration of {(a^, (p„): m > 1 and n > 1}. Set XQ = X and/o = / . Take ^oi^^ x) = {P: P solves the martingale problem for a and b starting from (s, x)} and UQ(S, X) — supj£^ fe-'^%(x(t
+ s))dt
P e %(s, x)\
and define inductively: e-'^%(x(t
^^^,(s,x)=^.Pe%(s,x):E^
+ s))dt
UN(S.
X]
and %+i(s, x) = sup £^ re-'^^%^Mt
+ s))dt
:Fe^;v+i(5,x)
By Lemmas 12.2.1 and 12.2.2, for each N > 0, {^^(s, x): (s, x) e [0, oo) x R^} is a subset of comp(M(Q)) which satisfies (a)-(d) of (2.1). Since C;v + i(s, x) ^ '^^[s, x\ it is clear that ^^(5, x) = flJ" ^;v(s, x) e comp(M(Q)) for all (s, x) and that {^00(5* ^ ) ' (5' ^) ^ [0' ^ ) >< ^'^l satisfies (a)-(d) of (2.1). Thus, if we can show that ^00(s, x) has at most one member for each (s, x), the theorem will be proved. Suppose that P, Q e ^^(s, x). Then, for all m > 1 and n> 1,
fe-'-YMt
+ s)) dt\=EQ
\fe-'-%(x{t
+ 5)) dt
Since {A^}f is dense in (0, 00), it follows from the uniqueness of the Laplace transform plus the continuity of E''[f„(x{ -))] and E^[f„(x( •))] that E'[fMt))] for all n>l.
= E%f„(x(t))l
t>0.
But {f„}f is dense in Co(K'') and so E'[f(x(t))]
= E<^[f(x(t))l
t>0,
for all / e B(R'^). We can now proceed in exactly the same way as we did in the proof of Theorem 6.2.3 to prove that each ^00(s, x) contains exactly one element. In fact, the proof here is easier since we know a priori that for each t > 0 and fe B(R'') the function u(s, x) such that E^[f(x(t))] = w(s, x) for all P e ^^(s, x) is measurable. The details are left to the reader, n An interesting consequence of the preceding is the following.
295
12.2. Markov Selections
12.2.4 Theorem. Let a: R^ ^S^ and b: R'^ ^ R'' be bounded continuous functions. Then the martingale problem for a and b is well-posed if and only if there is exactly one strong Markov, time homogeneous measurable Markov family {P/. x e R^} such that for each x e R^, P^ is a solution to the martingale problem for a and b starting from X. Proof We already know that the " only if " assertion is true. To prove the " if " assertion, note that if the martingale problem is not well-posed, then by Corollary 6.2.4 and Lemma 6.5.1 we can find an x° e R'*, a A > 0, a n / e CQ(R% and solutions P and Q to the martingale problem for a and b starting from (0, x°) such that
fe-'f(x(t))
dt\>E<^ \fe-'r(x(t))
dt
Applying Theorem 12.2.3 with this A and/, we know that there is a strong Markov, time-homogeneous measurable family of solutions {P]^^: x e R*^} such that
£nV re-'f(x{t))dt\>Enre-'f(x(t))dt Applying Theorem 12.2.3 with this X and —/, we get a second such family {Pi^^: X e R'^} such that £:n^c^ re-'f(x(t))dt\<E<^\\e-'f(x(t))
dt
In particular, {P^^^: xeR*^} cannot equal {P^^^: x e R**}, and so the proof is complete. D Less striking, but nonetheless useful for the purposes of the next section, is the following refinement of Theorem 12.2.3. 12.2.5 Theorem. Let a: R^ -^S^ and b: R'^-^ R*^ be bounded continuous functions. Given X > 0 and a bounded upper semi-continuous function f : R^ x R^ —• R \ there is a measurable map (s, z, x) - • P^^ of [0, oo) x R^ x R^ into M{Cl) such that for each z: (i) Pl^ solves the martingale problem for a and b starting from (s, x), (n) p^ ^ = p j ^ o 0)- \ 5 > 0, where O^: Q -• Q i5 the map described in (b) of (2.1) {Hi) if x: Q-* [s, co) is a stopping time and {P^^} is a r.c.p.d. ofPl^ ^ \ M^, then there is a PI ^-null NeM^ such that P^ = b^ (x),^^) P\^^^^ ^(,(^), „) for co^N, (iv) for each (s,x), E^Hf^ e-^'f(z,x{t + s))dt)] maximizes E^[f^e-^'x f{z,x{t-\-s))dt] as P varies over solutions to the martingale problem for a and b starting from (s,x).
296
12. The Non-unique Case
Proof. One can obtain this result as a consequence of Theorem 12.2.3 by the following trick. Define coefficients a-.R^""^ -^ SNU and b:R^-^^ -^ R^ by 5U(3^)^|a'^W \ 0 and
li(^^^[h\x) \ 0
if N + \
j
if N H - l < i < i V + rf otherwise,
where x = (z,x) ^ R^ x R^, Cleariy a and b satisfy the conditions of Theorem 12.4.3. Thus we can find a strong Markov, time-homogeneous, measurable family {Ps^: (5, x) G [0,00) X R^+^} of solutions for a and b such that E''^[J^ e-^'f(x(t + s))dt\ maximizes E^ [J^ e~^^f {x{t-\-s))dt] over all P which solve the martingale problem for a and b starting from (s,x). But, for each z € R^, there is an obvious one-to-one correspondence between solutions for a and b starting from a given point (s,x) and those for a and b starting from (s, (z,x)). Using this correspondence, the proof is completed by simply setting P^^ = Ps,iz,x)- D
12.3. Reconstruction of All Solutions Let a: R"^ -^ S^ and b: R"^ -^ R'^ be given bounded continuous functions. For each X G R'^, let ^(x) denote the set of all solutions to the martingale problem for a and b starting from x. We have just seen that there is at least one measurable selection x->P^e^(x) such that the resuhing family {P^:x e R'^} forms a timehomogenious strong Markov process. We have also seen that there is exactly one such selection if and only if each ^(x) contains only one element. The purpose of the present section is to refine these statements by showing that in a certain sense every element of ^(x) can be obtained by " piecing together " homogeneous strong Markov selections. We now give a precise formulation. Let s^ be an index set for the time-homogeneous strong Markov selections. That is, for each oc e j ^ there is exactly one time-homogeneous strong Markov selection {P^: x e R'^}. A mapping cp from a measurable space (£, ^) into s^ will be said to be measurable if (q, x)^ P^^^^ is measurable on (E x R'^,^ x Mj^^) into M(Q). Given x e R^ n > 1, 0 < fj < •• < f„, oL^e s^, and measurable maps ay: (Q, Ji^^) -^ s/, 1 <j
(3.1)
P7®nPtl%.-)®nPttJit.-)®'-'®n"! Jc(f„, • ) ' •)
where Ply = PJ, -^ O,"^ for any a e ja/, f > 0, and y e R'^ (cf b) of (2.1) for the definition of ^,). By Theorem 6.1.2, the measure in (3.1) belongs to ^(x). We will denote by ^^ the set of all measures of the form given in (3.1) generated by n > 1, 0 < ti < • * • < r„, OLQ e s^, and the maps oLy. (O, ^#,^) -• j / . It is our purpose to prove the following theorem.
297
12.3. Reconstruction of All Solutions
12.3.1 Theorem. For each x e R'^, <^(x) coincides with the closed convex hull S^ of
We first observe that proving Theorem 12.3.1 reduces to showing that for any N > 1, 0 < tj < • • < f;v, and bounded continuous/: (R^f -^ R^: (3.2)
sup E'[f{x(t,l
..., x(t,))] = sup E'[f(x(t,),
P 6 '^(x)
...,
x(tM
PeQ-,
To see that (3.2) is enough, we invoke the Hahn-Banach theorem for locally convex topological spaces to show that ^(x) = Q)^ is equivalent to sup £^[F] = sup P€<6{x)
E\¥\
Pe9,
for all bounded continuous F: Q -• R^ If one now notices that there is for each e > 0 a compact set iC in Q such that P(Q\K) < e for all P e ^(x) (and hence for P e Sx) and that by the Stone-Weierstrass theorem, every bounded continuous function on Q can be approximated uniformly well on K by a function of the form in (3.2) having the same bound, it is clear that (3.2) suffices. We will prove (3.2) in two steps. 12.3.2 Lemma. Letf: R^ x R'^ -^ R^ be a bounded continuous function. Given X>0 and n>2, there is a measurable map (si, . . . , s „ _ i , z, x ) ^ P J ^
'"-''' e 9,
of[0, oo)"" ^ X K^ X R'^ into M(Q) such that the function 00
(3.3)
„(z,x) = A"
, 00
•••
ri'
X e-^^''^-^'"Usi
"~''[f{z,x{s,
+ --+s„))]
••• ds,
is upper semicontinuous and satisfies. (3.4)
0„(z, x) > sup E^ Pe '6{x)
1"
(n-iyJo
s"-'e-''f(z,x(s))ds
Proof Define S; ) for bounded upper semicontinuous cp: K x R'^ -^ R^ by: S^(p(z, x) — sup XE^ P€'6{X)
I e '^^(p(z, x{s)) ds
298
12. The Non-unique Case
Clearly 5;^ (p is upper semicontinuous and is bounded by the same bound as (p. Thus, by Theorem 12.2.5, we can find for each 1 < m < n a measurable map a-R^-^j^ such that J Sr'"'''f(z.x) = XE^'^"'''
e-^'Sr'"f(z,x(s))ds
Define PsXi
s„-i,z
_ p a i ( z ) /-N p(X2(z) — •* X ^ti^tuxitu
•)
(x),,---(x),„_,Fr:i^> , , „ _ , . „ where t^ = YA ^J ^^^ ^"ify = ^7^'^ ° %^ ^^ ^^ ^^^y ^^ ^^^ ^^^^ (^i' • •' 5«-1' ^' ^) -> Pt^' '^"-1'^ is measurable. We will next show that 00
(3.5)
i"[ '0
. OC
• • • [ E^'^'- '-'•y(z.x{s,+
-+sM
*^0
X e A(s, + - + s„)^5^ • • • JS„ = 5 ^ / ( Z , X).
This will certainly prove that the „ in (3.3) is upper semi-continuous. It will also enable us to prove (3.4). To prove (3.5), we will show that for 1 < m < M — 1:
Jo
Jo X e-^^''^-^""Usi
(3.6)
Jo
••' ds,
Jo
x[sr'"-y(z,x(5i + --- + 5,,i))] X e
- / ( S l + - - - + S m + l)
^'^ dsi'-'
ds^+i
where psi,
..., Sm- 1, z
/^x^^^^0r,/^r?.%.,®r/--(x)r._,/^:
(z) l.x(f^-l,.)
Since pj'- '^"-'-- = P^Jv '""-''^ and
5",/(z,x) = Ar^^*^'
^ ~ ' ' 5 r y ( z , x(s))^s
12.3.
299
Reconstruction of All Solutions
(3.5) certainly follows from (3.6). To see (3.6), note that P^^Viix'"'^ equals
.00
,00
,00
,00
^ r I •••I
E^^V.---^
XE^fzUl^i
e-''S"-'"-^f(z,x(t^
+ s))ds
0 A(si + - - - + S m )
X e
00
= r
, 00
• • • I £^s,'.- • — [ s r ' " / ( 2 , x{s, +••• + s„))] "0
0
X e
A ( s i + ••• + Sm)
dSi
••• dSr
In view of (3.5), (3.4) will be proved once we have shown that if m > 1, then jjm
SXz, x) > E'
,00
(m-l)!Jo
for all P e ^6(x) and bounded upper semi-continuous (p: R^ x R'^ ^ R^. This is obvious for m = 1. Assuming it for m, we have for P e ^(x): ^m +
s'"e-''(p(z,x(s))ds ml .'o ]m
.00
I r - 'e-'^E'' UE^' I e - X z , x(t + 5)) ds m
1 )! *^n jjm
dt
U n
,00
where {Py is a r.c.p.d. of P\,#,. (3.4). D
We have therefore completed the proof of
12.3.3 Lemma. For every N > 1, 0 < f i < - • < r ^ , and bounded continuous / : (Ry^R\ (3.2) holds. Proof. Clearly it suffices to show that the left hand side of (3.2) dominates the right. We will do so by induction on N.
300
12. The Non-unique Case
If iV = 1 and if r > 0 and a bounded continuous/: R*^ -* R^ are given, choose (si, ..., s„_i, x)-• PJ^' ' *"" ^ as in Lemma 12.3.2 relative t o / a n d ^„ = n/t; and define 00
.00
•'o
•'o
Note that QJ e S^ and that
'0
"0
< A" r ••• f V ^ — [ | / ( x ( r ) ) - / ( x ( 5 , + •••+ s„))|] -'o •'o X e-^"^''^-^'"Usi '" ds„. By the weak law of large numbers:
as n -> 00 for all e > 0. Thus, since ^(x) is compact, (3.7)
X:C--C •'O
sup£''[|/(x(i))-/(x(5,+ *'0
-+s„))|]
Pe<€{x)
as n-^ CO. In particular, sup £''[/(x(0)] > lim sup E<^'[f(x(t))] Pe3,
n-*oo 00
= limsupAjJ
^00
•••! E^'^'' '-y(x(s,
+
X ^ - A „ ( 5 , + - + s„)^^^ • • • t / 5 ,
> lim sup E^ j^^j
s^-'f(x(s))e-^-ds
for all P 6 ^(x). Since - 1 ^ - A „ 5
[" ~ ! ) • *'{|s-r|>£} '{\s-t\>e}
as n -• 00, we have now proved (3.2) when N = 1,
^s->0
"+sM
301
12.3. Reconstruction of All Solutions
Next assume that (3.2) holds for N and let 0 < ^i < ••• < tj^+i and a bounded continuous/: (R'^Y^ ^ -^ R^ be given. By the induction hypothesis we know that if ^x(h^ •••^ h) and ^;c(^i» •••» h) denote, respectively, the images of ^(x) and S^ under the mapping induced by co -• (x(ti, co), ..., x(tj^, co)) then ^x(h^ • • •» ^N) = ^x{tu ", IN). NOW choose (si, ..., s„_i, z, x) -^ psu...,sn-i,z f^^ j ^nd Xn = n/{tN+i — IN) with z = (x^... ,x^) and x = x^+^ and define 0„(z, x) = AJ / • • • / Jo Jo
E^^E^^' "" |/(z, x(5i + • • • 5„))]
Since 0„ is bounded and upper semicontinuous, we can find P" e S^ such that £^"[a)„(x(ri),...,x(riv),x)]= sup £^[0„(x(ri),...,x(tiv),x)]. Next, define
It is obvious that 00
.00 Sn-l)
is an element of ^^. Moreover, by the same argument as we used when iV = 1 to prove (3.7), we have lim |£:^"|/(x(ri),...,x(tA,+i))]-£^"[0„(x(ti),...,x(r^),x)]| = 0 . Thus, it only remains to check that lim sup E^" [0„(x(ti),..., x(tN), x)] > sup E^ IfM^il • • •»^(^N+i))]n-»>oo
Pe'i^ix)
But if F € ^(x), then E^" [^n(x(ti),...,
x{tN), X)] > E^ [^n(x(tl),...,
>£^ gp'"
x/(x (ti,
•).
x{tNh x)] in
/•oo
.(«-!)! Jo ..., x(fN, O'^C'N + s))
302
12. The Non-unique Case
where {F'J} is a rx.p.d. of P \ M,^. That is, for P e ^(x) £^"[a)„(x(ti),...,x(rN),x)]
\yn
V). vQ
and so the desired result follows from the easily seen fact that
n — Ij! ^
= E^ \i{x{tx\..., We have therefore completed the proof of (3.2).
x(tN), x(riv+i))].
D
12.4. Exercises 12.4.1. Use Exercise 6.7.2 and the results of Section 12.2 to show that for any bounded continuous a\ [0, oo) x K^-^S^ and h\ (0, oo) x R^^K^ there is a measurable family {Ps,x' {^^ x) G [0, oo) x R^\ of solutions to the martingale problem for a and b such that for any (s, x) and any finite stopping time t > s: {<5co (8)r(co) ^T(a>), X(T(CO), W)} IS a r.c.p.d. of Pj x\'^x- ^l^o, notc that all solutions can be recovered from members of such strong Markov selections in the sense described in Section 12.3. Finally, observe that one does not really need a and h to be continuous with respect to i so long as they are continuous with respect to x. However continuity in t is needed if one wants to reduce the inhomogeneous case to the homogeneous case. 12.4.2. It is reasonable to ask if one cannot always make a Feller continuous selection of solutions when the coefficients are continuous. After all, uniqueness guarantees Feller continuity automatically. However the following counterexample shows that this may not always be possible. Let f/ = 1, a = 0 and h(x) any bounded continuous function which equals x/|x|^ = (sgn x)|x|* for |x| < 1 and which is continuously differentiable for |x | > 1. Show that if x 7^ 0, the solution to the martingale problem for a, h starting from x has exactly one solution P^^. Next show that P^ has a limit as x -• 0 through positive or negative values, but that these limits are different. Conclude from this that no Feller continuous choice exists. 12.4.3. Let a\ K^ ^S^ and h\ K^ -^R^h^ bounded measurable functions such that a(*) is uniformly positive definite. Show that there is a measurable strong Markov family {P^\ xe R^\ of solutions to the martingale problem for a, h. To do this, it suffices to treat the case of 6 = 0. Using 7.3.2, choose C(T, R)< 00 hr T, R>0
303
12.4. Exercises
such that for all x e R'^ there is a solution to the martingale problem from (0, x] satisfying (4.1)
j /(x(t)) dt
for all/e Co(B(0, R)). For each x e R'' let
Appendix
A.O. Introduction This appendix is devoted to the derivation of the inequalities on which the results of Chapters 7 and 9 depend. To be precise, let c: [0, oo)-^ S^ be a measurable function for which there exist 0 < A < A < oo with the property that (0.1)
/l|^|^<<^, c(r)^>
r>0
and 0 e R'.
Define C(s, t) = J^ c(u) du for 0 <s
^(^)(5, x; r, >;) = (2n)-"^dot C(s, r))"* (y-x,C(s,t)-'(y-x))
X exp
for 0 < 5 < r and X, y G R". F o r / e C^([0, oo) x R'') and T > 0, set (0.3)
GJ)/(5, x) = fdt\
f(u y)g''\s, x-Uy) dy Rd
on [0, T] X R^. What we have to prove is that (0.4)
5^GJ,/ dxi dxj
X, A)||/||LP([o,r)xi?d)
T]xRd)
for all 1
l/f */|ii.P(RxR|
1 < P < 00,
A.O. Introduction
305
where (0.6)
^x2/2s
/c(s, x ) = X(-00, o)(s)
An indication that (0.5) cannot be entirely elementary is contained in the observation that " /c * / " cannot be given an immediate meaning. Indeed, k is not an integrable function and so the integral defining k * /is not a Lebesgue integral and must be defined via a principle value procedure. To be precise, set k^^\s, x) = Xi-1/£, -e)(s)k(s, x) for e > 0. Then k^'^ e 1}(R x R) and so k^^^ */not only makes sense, but also, by Young's inequahty, satisfies (0.7)
/|C<^> *f\\LHR>^R) ^ WnLHR.RAf\\mR>^RV
1 < P < 00.
What we want to do is define (0.8)
k*f=\imk^'Uf.
However, it is not clear that the limit in (0.8) exists, since the right hand side of (0.7) explodes as e \ 0. It should by now be clear that in order to get anywhere we are going to be forced to take advantage of cancellations; putting in absolute values is just too crude. To see how cancellations enter, we compute k^^^ */. (Throughout we will use the convention:
J mi
Rd
and
^^«) = ?2iyl/-""^W''^ for/e L^(/?'').) A simple computation yields: IA^)(T
;=\ -
^1^1 (M2iz-\^\2)I2 _
(2iT-K|2)/2n
In particular,
limfe<->»/(T,^) = ,../l^i,J(T,^) To'
' ^•'"
2iT- k
exists in I}{R x R) for/e I}{R x i?), and we have (0.9)
sup|i/c«*/|U,,,.R)<2||/|U,,«.«,. e>0
306
Appendix
Of course, the derivation of (0.9) involves ParsevaFs equality and the cancellations missed by (0.7) are hidden in the Fourier analysis. Nonetheless, this simple computation demonstrates that there are cancellations to be exploited if one is clever enough to do so. The archetypical example of this sort is the Hilbert transform; and it was through a thorough understanding of the Hilbert transform that Calderon and Zygmund [1952] were able to make their original discoveries in the area which is now known as the theory of singular integrals. The type of singular integral which we want to study does not fit into the scheme of Calderon and Zygmund but is very closely related. A way to handle them was first given by B. Frank Jones [1964] and was fully developed by Fabes and Riviere [1966a]. The approach that we are going to adopt is based on the ideas of Fefferman and Stein [1972] and is an adaptation of the presentation in Stroock [1973]. We are grateful to E. Fabes who communicated the proof of Lemma A.2.2.
A.l. Lp Estimates for Some Singular Integral Operators In the preceding section we saw that, with the aid of Fourier analysis, it is possible to prove estimates of the form (0.4) when p = 2. In order to get other p's, we want to apply interpolation theory. However, (0.4) is false for p = 1 and p = oo; and therefore the Riesz-Thorin theorem is not directly applicable. One is therefore forced to adopt another form of interpolation theorem. The one usually used is that of Marcinkiewicz for which one must estabhsh a weak-type (1, 1) estimate. For an elegant treatment of this method, the paper of Riviere [1971] is recommended. In order to introduce what we consider to be an interesting technique to a new audience, we have decided not to take the Marcinkiewicz route but, instead, the newer Fefferman-Stein approach. As indicated by Exercise A.3.2 below and the papers of Burkholder and Gundy [1970] and Garsia [1973], these ideas ought to prove useful to probabilists in the future. We have therefore made our presentation look as much as possible like " probabihty theory." A. 1.1 Lemma. Let (£, ^, fi) be a probability space and {<^„}o a non-decreasing sequence of sub a-algebras. For each n > 0, let X„ be an ^^-measurable random variable. Assume that there exists a fi-integrable non-negative random variable Y such that (1.1)
sup E[\X„ -X^.,\
I J ^ J < E[Y\^J
(a.s., fi)
n>m
for all m> 1. Then, for every a, )5 > 0; (1.2)
Jsup \X„-Xo\ \n>0
>oi-\I
p]<-E ^
y,sup \x„-Xo\ n>0
>p
A.l. Lp Estimates for Some Singular Integral Operators
307
Proof. Given N > U let a = mm{n: n>N or \X„- Xo\ > p} and T = mm{n: n> N or \X„- Xo\ > (x + p}. If B e J^^, then E[\X,- X,_,l B]< E[\X, - X,.,\, = l^E[\X^-X^^,\,Bn{a
B] + E[\X^ - X,l B]
= m}]+ Y^ E[\X^ - X„l B n {x = n}] M = 1
m=1
< t E[Y, Bn{(j = m}]+ f £[7, B n m=1
{T =
n}]
n=1
= 2£[y, B], since B n {cr = m} and B n {T = n} = ((J" B n {cr = m}) n {T = n} are elements of ^„ and #^„, respectively. We have therefore proved that (1.3)
E[\X,-X,^,\\^,]<2E[Y\^,]
(a.s.,n).
In particular: J sup \X„-Xo\ >a + li\ = n{\X,-Xo\ > p and \X,-Xo\ >a + H) \o
/
= ?£ a
Y, sup |X„-Xo| >p 0
Letting AT/^oo, we arrive at (1.2). Q A.1.2 Lemma. Let (£, =^, /i) an^ {^n}o be as in Lemma AAA. Assume that ^0 = {<^> E} and ^ = cdJcT ^^n)- Oiyen a fi-integrable random variable X, define
308
Appendix
X„ = E[X I^„l « > 0, and set X^ = sup„>i E[\X-X„.,\\
J^J. Then for
1 < p < 00;
(1-4)
^-^^E[\X-E[X]\r-'<E[\X^\r-'^-^lE[\X-E[X]\r''
in the sense that if any of these is infinite then they all are. Proof First observe that we need only prove (1.4) under the assumption that X is bounded, since we can always reduce to this case by truncation. In particular, we will not worry about any of the quantities in (1.4) being infinite. Next, note that for any 1 <m
\^„] <E[\X-
X„\ IJ^J + E[\X-
X„_, I J ^ J
= E[E[\X- X„\ |i^„. J l i ^ J + E[\X<2E[X^^„]
X„
^m\
(a.s.,/i).
We can therefore take 7 = 2ZMn Lemma A. 1.1 and thereby obtain: /ilsup \X„-Xo\
>a + )9)<-
Xt,sup
\X„-Xo\>P]
n>0
for any pair a, j5 > 0. Set X = sup„>o \X„- Xo\. Then (1.5)
fi(X>{\ + (x)R) < — E[X\ X>R] (X.R
for any a > 0 and R>0. Proceeding as in Exercise 1.5.4, we have from (1.5) that
for any 1 < p < oo and a > 0. Noting that X - E[X] = lim„^„ X„ - XQ (cf. Exercise 1.5.10 and note that XQ = E[X] (a.s., /i)) and taking a = l/(p — 1) in (1.6), we obtain the left hand side of (1.4). To get the right hand side of (1.4), observe that X^ <2 sup„ E[\X - XQ] \^„] and, therefore, by Doob's inequahty: £[|j^t|P]l/P<2£
sup E[\X -
Xo\\^„r
p-1
A.1.3 Theorem. Let (£, ^, ft) be a probability space and {^„} a non-decreasing sequence of sub a-algebras such that ^Q = {0, E] and ^ — G{\]Q ^„). Given a
A.l. Lp Estimates for Some Singular Integral Operators
309
fi'integrable random variableX,setX^ = sup„>i E[\X - E[X\^„- J | |^„].Suppose that T is a linear map defined on I?(fi) n L*(/i) into l}{fi) and assume that M«||/|U«„„
feL^(p.)nI?(ji\
IIT/|U.„, < M, ll/IU,,,,,
/ 6 I}(n) n L"(M),
for some M^ and M2. Then for each 2 < p < 00;
(1.7)
||T/- £[T/]||„,,, < ^JP^M^-<^">(4M,)^"'||/||„,„.
Proof Given a ^^^ x ^^-measurablefif:Z"^ x £ - • C satisfying \g\ = 1 , define [/,(«)/= E[(Tf-
E[Tf\^„.Mn)\^„]
for / e L'^(fi) n L'^in). Clearly, |l/g(n)/| < (r/)t. Next, given g and an immeasurable T:£ -• Z+, define [Ug{T)f]{q) = [C/g(T(^))/]te), ^ € £. Again: |i/,(T)/|<(r/)t. Thus, by assumption, ||t/,(T)/|U.w<M„||/||,.,„,. Also, by assumption and the right hand side of (1.4),
l|i/,(t)/lk.(„<||(r/)t|U.,„, <4||r/-£[r/]||,.<,, <4||r/||,.„<4M211/11 Thus, by the Riesz-Thorin interpolation theorem, applied to Ug(x), (1-8)
l|l^,(t)||L„., < Mi-<^"'>(4M,)^"'||/|U^,,
for 2 < p < 00. Maximizing the left hand of (1.8) over all choices of g and T, we arrive at (1-9)
ll(r/)l.„„ < M-<^""{4M,)^/''||/||,^,,
for 2 < p < 00. If we now apply the left hand side of (1.4), (1.7) results.
D
We now want to develop the context in which we will be applying Theorem A. 1.3. Given a bounded measurable set S ^ R x R^ and a locally integrable/on R X R**, define/s to be the mean value of/on S. (That is,/5 = 1/151 Js /, where
310
Appendix
ISI is the Lebesgue measure of S.) Let ^ stand for the collection of all" parabolic" cubes Qin R X R^ of the form
and
Xj — S < yj < Xj + ^, 1 <j < d}
for some (s, x) e R x R'^ and S > 0. F o r / e LlX^^ x R% define (1.10)
||/|U = s u p ( | / - / e | ) Q .
Note that || • ||^ is a semi-norm, and that ||/||^ = 0 if and only if/is equal almost everywhere to a constant. Hence, if 0>(R x R'^) = {fe Ll,(R x i?'*)/C: \\f\\^ < 00}, then ^(R x R"^) together with || • ||^ forms a normed linear space. When talking about elements of ^(R x K'^), we will indulge in the usual abuse of identifying the equivalence class of / with / itself. To get some feeling for the position of 0^(R x R*^) among more familiar function spaces, note that n^(R X R'^yc c ^(R X R'^) (in fact ||/||^ < 2||/||jroc); but the opposite inclusion fails since/(s, x) = log |xi | is an element of ^(R x R'^) which certainly is not in I^(R X R*^). (Exercise A.3.1 shows that log|xi | is just about as singular as an element of ^(R x R'^) can be.) The importance of ^(R x R"^) to us is contained in the next lemma. A.1.4 Lemma. Given N>0, let E^^^ = {(t, y) e R x R'^; - 4 ^ < r < 4 ^ and - 2^ < yj < 2^, 1 < ; < d} and let ^^^^ be the Borelfield on E^^\ For n > 0, let ^^^^ be the partition of E^^^ consisting of sets Q having the form {{t,y):lo^'-"
+
l)4'-" and
/y2^-"
l<j
< d. with r = (/o, ..., /j) 6 Z*^^^ satisfying - 4 " < /Q < 4" and -2" < Ij <2M<j Set J ^ f ^ = (T(^^^^). Finally, let fi^^^ denote normalized Lebesgue measure on (E^^\ i^^^)) (i.e., //(^>(r) = I r I /1 £<^) \ for re J^<^>). Then Fff > = {0, £:<^>}, J^f > e J^^^^ j, and J^<^) = (T([J^ ^^P). Moreover, iffe L^(/z<^>) and
f\,, ^ sup ^ n i / - ^n/i^^i^-Mi i^n then /,V, < 2''^2 sup sup ( | / - / e | ) G . « > 0 Q€^„W
Froof, We will drop the superscript N during the proof. The assertions about the ^„'s are obvious. To prove the stated estimate, let Q„(r, y), n > 0 and (r, y) e £,
A.l. Lp Estimates for Some Singular Integral Operators
311
denote the unique element of ^„ containing (r, y). Then:
£"[ I / - £"[/ | ^ „ . J I I J^„](t, y) = (\f-fa,_,,, ,>| W , ,
= 2
(|/~/<2n-i(r,y)l)Q„-i(r,y)
since e„(f,y)eg„_i(r,y) and
en(^y)i
= 2''-^^. D
We are now ready to prove the interpolation theorem which we will be needing in the next section. A.1.5 Theorem. Let T be a linear mapping of I}(R x R^) n L*(R x R^) into l3(R X R^)for which there exist M^ and M2 such that r/|U(Kxi,.) <M^||/||^oo(«.;,.),
/ e I}(R X R') n L-(R x R'\
and
Then if2
(1.11)
00;
||T/|U.,«.«., <^^^(2''*^Mj'-<^"'>(4M,)^"'|i/||„,,.,.,.
Proof. Given N > 0, define T*'*>/to be the restriction of T{fxEm) to £*">. Then
and, by Lemma A. 1.4, IK7^~'AVIit-.„.«„^2^^^sup sup (|7^^!/"-(7^^'/)cl)e
Thus, by Theorem A. 1.3, lipmf_ E^-X-P^m,,,,,.,
<^
^
(2^-^M J'-'^'")
X (4M,)^"'||/||i,,„«,
312
Appendix
for 2 < p < 00. In other words LPiR X Rd)
Since ( T ( / - X£w))z£(iv)-> T/in l}(R x R% (1.1) will follow if we show that \\{T(fXEiN)))EmXE(N)\\ LP{R X K«0
^
as iV-* 00. But
\\(TfXEiN)))E{N)XE(N) WLHRXR^
^
|E;(N)|1-1/P
L2(Jl X Rii)
|£(^)|i-l/p ^ 2 \ \ J
\\L2(RxR
^
|£;W|i-i/p
since p > 2. Thus the proof is complete.
D
We conclude this section with the proof of a criterion which ensures that an operator satisfies the hypotheses of the preceding theorem. Before doing so, we need to introduce the "parabolic metric'' on R x R^. For (s, x) e R x R'^, define
,U2,
p,.„=(lit±(lill±^)'.
Note that p(s, x) is determined by the property that: (113)
-f 1^1 = 1p'^is, x) p^(s, x)
from which it is easy to see that p(s -{-1, x + y) < p(s, x) + p(t, y) and that for each (s, x)e R X R^iO, 0)} there is a unique rj = (rio,tj)e S'^ such that (5, x) = (p^(s, x)rio, p{s, x)fj). Finally, given (5, x) e R x R**, the Jacobian of the transformation (f, >;)-• (5 + p^(t - s, y- x)rio, x-\- p(t- s,y- x)fi) is
(1-14)
p'^'Uri)
A.l. Lp Estimates for Some Singular Integral Operators
313
where Ja(ri) is a smooth function on S*^ satisfying: (1.15)
l<J,(ry)<M„
rjeS'.
This metric p and its associated " paraboHc polar coordinate system " is a useful tool when dealing with singular integrals having parabolic homogeneity. We will be making use of it in the next section; the reason for our introducing these notions here is mostly based on aesthetic considerations. What they do is enable us to make a clean translation of the " almost 1} " condition of Hormander [1960] into the parabolic context. These tricks are due to Fabes and Riviere [1966a]. A. 1.6 Theorem. Suppose that k: {R x R'^Y -*C is a measurable function such that ^ ^ \k(s,x;t,y)\
dt dy < co
for all (5, x)e R X R"^. Define the operator K on I}(R x R"^) n If(R x R'^) by Kf(s, x) = j j k(s, x; r, y)f(t, y) dt dy, and assume that (1-15)
||K/||,.,«
If there is a Bfor which: (1.16)
I |/c(5, x; r, y) - k{a, c^; r,
j
y)\dtdy
p(t-s,y-x)>2p(
for all (cr, i) and (s, x) in R x R'^, then (1-17)
l|K/IU^K.«^, ^ (^3^(MJ'-<^""(M,)^"'||/||,^,.,.,
for 2 < p < 00, where (1.18)
M , = 2-^'{\{(t, y): p{t, y) < l(d + 1)}|*/1 + 2B)
and (1.19)
M2 = 4A
Proof In view of Theorem A. 1.5, all that we have to show is that (1-20)
\\Kf\\
<(\{{t,y):p(t,y)<2{d+l)}\*A
+ 2B)\\fU co(R X Rd) •
314
Appendix
To this end, let fe U'(R x R'^) n I}(R x R^) be given and suppose that Q = {(t, y): s- S^
(p=fXs^
and
^^f-cp.
Then (\Kf-{Kf)Q\)Q<{\K
+
{\Kxj^-{Kxli)Q\)^
Note that (I A> - (Kq>)Q \)Q<(\KCP-
(KCP)Q \%<(\K(p
H|
and 15'
\{(t,y):p(t,y)<2(d+l)}\.
Next, observe that if (cr, i) e Q and (t, y) ^ S, then p(t — 5, y - x)> 2p(G — s, { — x). Hence, if we use the fact that
{\KXP-{KXP)Q\)Q
< :^J
\Kxp(c7,0-Kxp{s,x)\dc7di,
then (1.16) yields :
e.. <
2
X
\k((T, ^; U y) - k(s, X, t, y)\ dt dy
j VO-s, y-x)>2p(
< 2B\\ I
\\L^(RxRd)'
315
A.2. Proof of the Main Estimate
Combining these, we arrive at: (\Kf-
(X/)ei)Q < (|{(f, y): p{t, y) < 2{d + 1)}|*^ + 2B)||/|U.,«XR.,
and so (1.20) is now proved.
G
A.2. Proof of the Main Estimate We now have made all the preparations necessary for the proof of (0.4). What remains is to see how to apply these preparations. In order to eliminate unnecessary sub- and superscripts during the actual proof, we begin by establishing the notation which will be used throughout this section. Let c: [0, GC)->SJ be a measurable function which satisfies (0.1) for some 0 < A < A < 00. Extend c to i? by taking c( — s) = c(s), 5 > 0, and set C{s, t) = \l c(u) du for s
ds\ g(s, x; r, y)f(s, x) dx. •^-00
^R
In order to prove (0.4) for all 1 < p < oo, it is sufficient to show that for all 1 < i, j
1 dxi dxj \LP(R
< Q ( p , A, A)i|/||Lp,R.R., xRd)
and
(2.2)
1 d^G*f 1 ^yi ^yj \LP{R 1
316
Appendix
for all 2 < p < 00, where Cj(p, A, A) depends only on d, p, A, and A. Indeed, a simple duality argument shows that (2.2) for a given p imphes (2.1) for the conjugate p' of p (i.e., (1/p) + (1/p') = 1 ) . Since there is no difference between the derivations of (2.1) and (2.2), we will restrict our attention to the proof of (2.1) for 2
For 0 < e < 1 and 1 < i, j < d, define k\f(s, x;t,y)-= X[e, m{t - ^ ) ^ ^ ^ ^ (^> x; t, y) =
Xie,m(t-s)[((C(s,t))-'{y-x)), x((C(s,t))-'(y-x))j - Sij(C(s, t)~%]g(s, x; t, y).
Clearly jj \k^\s, x; t, y)[ dt dy < CO for all £ > 0, 1 < j , ; < d, and (s, x) e R x R'^. Define
K^f(s.x) =
\jk^(s.x;t.y)f(t,y)dtdy
f o r / e If'(R X R'^). Our first lemma shows that the proof of (2.1) reduces to proving (2.4)
sup i|KW/i LP{R X i?d) ^ Q(P, A. A)||/|| LP{R X Rd)
0<e< 1
for all 1
(2.5)
| J ^ ( s , i) = -^,ij j%-«.c,.,,K>/2;.(j^ ^) rf,
5Xt ^Xj
and .S + (l/E)
(2.6)
KJJ^ '(.S, 0 = -<^.^; j ' -"e-<4,C(.,,K>/2/.(j^ ^) ^,.
(Here ^^(s, c^) = J«, e'<^'^>(p(s, x) ^x, (J e R^ /or
A.2. Proof of the Main Estimate
317
Proof. The first assertion as well as (2.5) and (2.6) are easy consequences of the spacial translation invariance of the operators G and K\f. Finally, from (2.5) and (2.6), it is clear that
in I}(R^) at a rate which is independent oi s e R. Hence the last assertion follows from the Fourier inversion formula. Q We now turn to the proof of (2.4). The idea is to show that k\f satisfies the hypotheses of Theorem A. 1.6 with A and B independent of 0 < 8 < 1. The first step is the following. A.2.2 Lemma. (Proof communicated by E. Fabes). For all 0 < e < 1 and feU'iR xR'')n l}(R x R^): (2.7)
m^fU^.^^
< 2(AM^)*||/|U.<^x«^.
Proof Let/e C^{R x R^). By Parseval's equality: (2.8)
mn\\LHK>^R^^ ^ ^ [ \
1 1 ^ ^^*' ^^"'^^^^ "^^r
By (2.6), \K^f ^(s, 01 < 1^1' [^-(^^<^)-^^(^))/2|/-(r, 01 dt ''s
where C^(t) = jo (i^ c{u)^} du, t e R and ^ e R\0}. Changing variables, we get \K,j
( s , o i < K l j^j
CHQ-HO)
= (p * »A^(Q(5)), where (p(t) = e"^X(-oo,o)(t) and
^^(^)--^-Q(cnor"Note that
Ht) ^
r(Q'(tu)i
318
Appendix
and so l-HR)
1 \\
W
||/1C,-'CU)IU.,K,
2A*
^ I T 11/1'
|i^2(/?)
It is evident from this and (2.8) that (2.7) holds.
Q
The final step is to check that there is a B, independent of 0 < £ < 1, for which (1.16) holds when k is replaced by k\f. A.2.3 Lemma. There is a constant ^^(A, A) such that (2.9)
kfj\s,x;t,y)-kr>(<j,i;t,y)\
j p(t-s,
y-x)>2pia-s,
dt dy
^-jc)
< B,(l A) for all 0 < e < U 1 < U j < d, and (s, x), (a, {) e R x R'*. Proo/. Let %(5,x;r,>;) = x(o,oo)(t-s)[(C(s,0~'(>' - x))i{C{s,t)-\y
- x))j
-dijC{sa)n'\ X ^(s, x;t, y).
Note that for s < r: 1
1
k(s.x;f,y)|
ly-xP/ZAir-s)
and so
^
1
+ l(t
- s)
|y-x|2/2A(f-s)
319
A.2. Proof of the Main Estimate
Clearly: \k^\s,x;t,y)-k^(G,i;t,y)\ < \kij(s, x;t,y)-
/c,-,(cr, i; t, y)\
+ IX[e, i/s]{t -s)-
X[e, m(t - cr) I Ikij(s,
x;t,y)\.
We will first show how to estimate f *'p(t - X, y - x) > 2p(
\Xls,l/e](t-s)-X[e,m(t-(r]
kij(s, x;t, y)\ dt dy. Obviously, this comes down to looking at
{Ink)
^p{i-s,y-x)i2a
X j X ( 0 . co)(f -
X) I XU. m(t
{\y-AY
-
S) -
X[e, l / e l ( t " < T )
exp[-b-xP/2A(t-5)]
1
(t-sY
A(t - s)
U(t-s)/
d/2
"'^ ^y
r>0
exp[-|y|V2Ar] ^d/2
^f ^y
•'p(l,y)>2a/ri
r1 X J y I %U, l/^](0 - /[e, l/a](^ " (^ " 5)) I
where a = p(o- — 5, (^ — x). Since lo" — 5| < a^, the last expression can be written as the sum of four terms, each of which is less than or equal to:
-'^CH
.J(¥)'^l}
-|yP/2A
|y|2/2A
dy dy
\yW2\ dy.
320
Appendix
Using the fact that p(U y) < I + ly I, one can easily check that |>'|2/2A
t •'p(l,y)>2/fi\\
^ /
dy,
t > 0,
^/
is bounded by a constant depending only on d. A, and A. Thus, it remains only to estimate: kij(s, x;t, y)p(t-s, y-x)>2p{a-s,
kij{(j, i; t, y)\ dt dy.
^-y)
Note that: \kij(s,x; (2.10)
t,y)-kij{G,
i; t,
y)\
(-ao,r)(5A(7)|
sup SA<7
4- X(-00,r)(^)\^-x\
dk - ( T , x;r, y) dx
sup IV/c,>, (l-y)x
+
yi;t,y)\.
0
Given T < r, we have: -
dk • ^ (T, x; r, y) = [ C ( T , r)" ^C(T)C(T, t)-'(y
-
x)l
x[C(x,t)-'(y-x)]jg(T,x;t,y) ^[C(x,t)-'(y-x)l X [ C ( T , r ) - ^C(T)C(T, r ) - ^(y - X)],-^(T, X; U y)
-d,,(C{x.t)-'c{x)C{x,t)-%g{T,x-Uy) + i < y - X, C(T, r)- ^ C(T)C(T, t ) - Hy - x) > /C,;(T, X; r, y)
^-[detC(T,r)] 2
det
/ciy(T, x ; r, y )
r)
C(T,
Thus for T < r - ( T , X;
dx
r, y)
1
/
A
^lA'^ \
^-|y-x|2/2A(f-T)
(r-T)^'72 + 2 1 / 5A ^lA'^ \ (2^\2F^^"^2?^^^^;
321
A.2. Proof of the Main Estimate
X _iZ_^_L_^-|>'-^P/2A(r-t) d/2 + 3
(t-z)'
where Cd(A, A) depends only on d, A, and A. We now switch to the parabohc polar coordinates described in the paragraph preceding Theorem A. 1.6. Let 0- = 5 H- OC^OQ , ^ = X -\- 0(3, t = s + P^WQ , and y = x -\- pat where a, p > 0 and (^0» ^)^ (<^o. ^ ) e 5'^- Assuming that p > let and that o^ A s < r, we have for (T A S < T < ( a v s ) A f :
^ C,(/l, A)
,
w
2A 2
CO
h-^)
12/
'="h-^)
d/2 + 2 + /
Clearly CO
exp 1 —
r^ I -7
'
ito
T
CO |2i ,d/2 + 2 + /
1 4«-^)i-(-^) T-S\>
is bounded so long as (COQ — (T — s)/p^Y > ^ . On the other hand
> — ,
CrA5
16
since a < p/2. Thus |co|^ > ^ marks, we have
sup tTAS
if (COQ - (T - s)/p^f < ^ . Combining these re-
dk ;^(T,x;t,3;) dx
y.CAK A) ,d + 4
Appendix
if p > 2a. Hence sup
\s-c\j y-
x)> 2p(a -s, t>(T AS
^ — x) a AS<x<(a
< y,C,(X, Ay
dt dy v s) At
f p-^ dp \ J,(o)) do
where y'^ equals y^ times the surface area of {(r, y): p(u y) = !}• We have used here the fact that dt dy = p^^^Ji(m) dp dm, where Jd(co) is discussed before Theorem A. 1.6 and d(o denotes surface measure on S^. By essentially the same argument, one can show that there exists Cd(A, A) such that sup | V M ^ , ( l - y ) ^ + y ^ ; f , y ) | < ^ ^ 0
P
when p = p(t — s, y — x)> 2p((T — x, ^ — x) and s < t. Thus: IX - >;I [ pit-s,
sup IWkij(s, (1 - y)x + yi;t,y)\ y-x)>
2p(a -s, i~x) t>s
dt dy
0
< Q(A, A)a
p~^ dp \ Jd(co) dco
= iQ(A, A) f J,(co) dco, where a = p{(7 - s, i — x). Combining this with the preceding, we see that the lemma now follows from (2.10). D We have at last made all the calculations necessary for the proof of (0.4). We give a precise statement in the following. A.2.4 Theorem. Let AJ^X, A) = 2(A/A^)^ and let Bj,X, A) he as in Lemma A.2.3. Set M^ = 2'-'( I{(r, y): p(r, y) < 2(d + 1)} |M,(A, A) + 3B,(A, A)) and M2 = 4A,(X, A).
A.3. Exercises
323
Then, for 2 < p < oo, (0.4) holds with C,(p, A, A) = ^
^
(M^)^-^/^(M,p;
and for 1 < p < 2, (0.4) holds with Ca(py K A) = Q(/7', A, A) where p' is the conjugate of p. Proof As we pointed out before Lemma A.2.1, proving (2.1), and therefore (0.4) for 2 < p < 00, is equivalent to proving (2.4) for this range of p's. But (2.4) with the specified choice of Cd(p, >l, A), 2 < p < oo, is immediate from Lemmas A.2.2 and A.2.3 together with Theorem A.L6. To get the desired result when 1 < p < 2, we have to prove (2.2) for 2 < p < oo; and as we said before, this is accomphshed in the same way as we proved (2.1). It turns out that (2.2) holds with the same Ca(p, A, A) as (2.1), and this is the reason that we can take Ca(p, A, A) = Q(p', 1, A) for 1 < p < 2. D
A.3. Exercises A.3.1. Starting with Lemma A. 1.1, check that if (£, #", /i) is a probability space, {^„}o is a non-decreasing sequence of sub a-algebras of ^ , and {X„}o' is a sequence of integrable random variables such that X„ is #^„-measurable and such that
supsupE[\X„-X„_,\\^„] nt>0
n>m
is bounded //-almost surely by some constant M < oo, then for all y < l/2Me: exp|ysup | Z „ - X o | j
where Cy(M) < oo is a constant depending only on y and M. In particular, if J^o = {0» E} and ^ = (T([JO ^n\ apply this observation to show that for integrable X with |A:^||i^oo(^) < 00: £[exp(yX*)] < C(||Xt||,.(,)) when y < (2||X^||i^oo(^)e)" ^ (The notation here is borrowed from Lemma A.1.2.) Of course, this proves that for such an X and y: £[exp(y|X-Xo|)]
324
Appendix
The preceding inequality is, for the present context, the John-Nirenberg [1961] inequality. A.3.2. The analogues of Lemma A. 1.1 and the preceding for continuous parameter families of random variables can be derived when some regularity condition is imposed on the family as a function of the parameter. This can be done by a limit procedure starting from the discrete case. It is interesting that if enough regularity is assumed then a direct derivation can be easier. For instance, let (£, .^, //) be a probabiUty space, {^,: T > 0} a non-decreasing family of sub cr-algebras of ^ , and 0: [0, oo) X E -• C a continuous, progressively measurable function. Suppose that there is an integrable function Y such that sup E[ I e(t) - e(s) \\^s]<E[Y\
J^J
(a.s., fi)
t>s
for all s > 0. Show that for all a, j5 > 0: (sup \9(t)-e(0)\
>(x-l-p]<-E /
\r>0
^
y,sup \e(t)-d(o)\ >p r>0
The proof can be accomplished by the same procedure as we used in the proof on Lemma A.1.1, only this time one should exploit continuity. Once one has this inequality, it is clear that the analogues of Lemma A. 1.2 and the preceding exercise follow easily. In particular, if sup£[|^(O-0(5)||^J<M
(a.s.,/i)
for all 5 > 0, then exp|ysup |^(r)-0(O)|) < C,(M) for all y < l/lMe. An interesting application of this result is the following observation due to Gaveau [private communication]. Let a: [0, co) x R^ ^ S^ and h: [0, co)x R!^ -^R^ be bounded measurable coefficients which determine a measurable Markov family {Ps,x • (s, x) € [0, QO)XR^} of solutions to the martingale problem for a and h. Suppose that / : [0, oo) x l?*^ -• C is a measurable function such that £P...
|V(", Au)) du
<M
for all 0 < s < r and X G R\ Show that for all y < l/2Me, s >0, and x e R'^: expl y sup f{u, x{u)) du \ t>s rs
<
Cy(M).
325
A.3. Exercises
The idea is to set 6(t) = J^ /(M, X(U)) du and check that: E'^'i I e(t2) - e(t,)\ \Jl,^<M
(a.s., P, J
for all 5 < ?! < ^2 • Note that if/> 0, this result proves that sup £''*•* C f{u,x{u))du
< 00
if and only if there is a y > 0 such that sup £^^*^ e x p M
f{u,x{u))du\
< 00
(s,x)
A.3.3. Let (£, J*^, /i) and {^„}o be as in Lemma A.L2. Note that the same argument plus an obvious change in notation enables one to prove Theorem A. 1.3 when T is defined on functions from (£, J*^) to a complex separable Hilbert space jfi into functions from (£, ^) into a second complex separable Hilbert space ^ 2 - That is, if T: l}(^\ ^ j ) n L"(/z; Jff^)-^I}(^, 3^2) and T satisfies:
||(T/)].JL...<M, tJtr2\\L^iti) ^ ^"-'oo
1^1 IIL«(/i)
and l|7y|UjU.„,<M,
.^1
liL2(M)5
where ((^)]^2 = sup„>i E[\\(p - E[(p\^„. JIU^l<^J for cp e I}(ii\ J^JI then for 2 < p < 00
7y|UallL,w<(J^Mi-^"'(4M,p|
^1
l|LP(/x)-
As an application of this remark, we outHne the following derivation of Burkholder's inequaUty (cf 4.6.12). Take Jfi = l^(C) and 3^2 = C, and consider the operator T'.L^{n\J^i) -• V-{ii\J^2) defined by
nm)=I
(£ui^j - E[f,\^,.,]).
Check that ||r({A}?)IU.(,)<
fc/l
IIJTI ||L2(M)
and that E[ I T({/,}r) - £[T({/,}r) I ^ „ - . ] PI ^ J < 4|
fc/l
WJUCX | | L < » ( , / )
326
Appendix
Thus, by the preceding remark:
\\TU}T)\U^
<_Z£___2i-2/P42/P|
k/l II Jfi II LP(M)
(p-1)'
for 2 < p < 00. Next, show that T*: I}{n) ->• l}(n; Jfi) given by
r*/={£[/|i^,]-£[/|i^,.J}r is the adjoint of the map T. Also, check that
ll(r*/)].,|k.,„,<2
L«(^)
Hence, for 2 < p < oo
Working by duaUty, conclude that for 1 < p < oo:
||T({A})||„,„>
lll|iVIU.IU,(„,
if 2 < p < 00, and Cp = Cp-, if
T*f= T*{f- £[/]) and TT*f=f-E[f]. Thus
T*/IU, IUP,,, < 11/-£[/]|U^„ < c,|| ||r*/|| This is essentially Burkholder's inequahty. To get from it to the form in which the inequahty is usually stated, let (X„, J^„, fi) be a martingale. Given N > i, set f=Xj, and note that X„ = E[f\^„l 0
r/
\p/2 1 1/p
(si^„-^„-.r)
<£[|X^-Xon^ ip /N
\p/2] 1/p
(l|x„-x„..pj j
327
A.3. Exercises
Since the middle term is a non-decreasing function of N, this now yields:
(l|X„-X„_,|^j
C.
Pl2
1/P
<supE[\X„-Xon''' n>l
(£ix„-x„.,i^)
p/2
1/P
for 1 < p < 00. Note that the middle term is commensurate with £[sup„> 1 \X„ — XQ l^y^ for p's in this range, but that it no longer is when p = 1. A beautiful fact, discovered originally by B. Davis [1970], is that Burkholder's inequality remains true when p = 1 after one brings the " sup *' inside the expectation. For an elegant derivation of Davis's inequahty, the reader is referred to A. Garsia [1973]. A.3.4. The reason for our working with parabolic scaling can be most easily appreciated if one goes back to the kernel k in (0.6). One can easily check that k(X^s, Xx) = (l//l^)/c(s, x) for any A > 0. This type of homogeneity meshes perfectly with the homogeneity of the metric p introduced before Theorem A. 1.6 and is the reason why p is the natural metric with which to study such kernels. In connection with the preceding paragraph, suppose that k: R x R\0,(!)}-^C is a smooth function satisfying k(?i^s, Xx) = X'^^-^kis, x) for all X>0. Let k^^\ £ > 0, be given by k^'\s, x) = X[e, m(p(s. x))k(s, x). Show that there is a B for which k^% -t,x-y)'Pit-s,
y- x)>2p(<j-s,
k^'\
-t,i-y)\dtdy
4- X)
for all (c7, i), (s, x) e R X R'^ and £ > 0. Assume in addition that there is an A such that
for all £ > 0 and (a, ^) e R x R'^. Show that 11^^'^ */II LP(i?x/?-0 ^ <^pl|/|Up(/?x«d),
1 < P < 00,
where Cp is independent of £ > 0 and depends only on A and B as well as p.
Bibliographical Remarks
Chapter 1. The study of weak convergence of probability measures on complete separable metric spaces is carried out in detail in the works of Prohorov [1956] and Varadarajan [1958] and [1966]. One can find a convenient treatment of the topic in the books of Parthasarathy [1967] and Billingsley [1968]. For a detailed study of the theory of martingales the reader is referred to the books of Doob [1953], Meyer [1966] and Neveu [1975]. The compactness criterion for Markov processes that we have presented in terms of martingales is a modification of the results contained in Prohorov [1956]. Standard references for the extension theorems of Kobnogorov and Tulcea are Kolmogorov [1956] and Doob [1953]. Chapter 2. The regularity of paths in the Brownian motion case goes back to Wiener [1923]. Other basic references for results on the regularity of sample paths are Slutzky [1949]; Dynkin [1952] and Kinney [1953] for Markov processes, and the more recent work of Garsia, Rodemich and Rumsey [1970], for general stochastic processes. Chapter 3. For partial differential equations there are several sources. For example Protter and Weinberger [1967] on the maximum principle, Friedman [1964] for the existence and properties of solutions to partial differential equations with Holder continuous coefficients and Ladyzenskaja, Solonnikov and Ural'ceva [1968] for the existence and uniqueness of solutions in Sobolev spaces. For degenerate parabolic equations, the result we have proved is taken from Oleinik [1966]. These results, at least when the diffusion matrix {a'-'} has a smooth square root, have been proved probabilistically in Gikhman and Skorohod [1973] and McKean [1969]. Chapters 4 and 5. The classical reference for the theory of stochastic integration is Ito [1951]. By using the theory of martingales the proofs were streamlined by Doob [1953]. Other references for stochastic integrals are Kunita and Watanabe [1967], Strassbourg seminar notes by Meyer [1974-75] and McKean [1969]. The existence of Lipschitz continuous square roots for smooth coefficients {a'^} was proved in a paper by Phillips and Sarason [1967]. Chapter 6. The idea that one should look for distribution uniqueness for solutions of stochastic differential equations goes back to some papers by Girsanov [1960]
Bibliographical Remarks
329
and [1961]. The martingale formulation and the systematic use of the techniques of this chapter were initiated in Stroock and Varadhan [1969]. The existence of solutions to the martingale problem is a reformulation of the generalized existence theorem for Ito's equations proved by Skorohod [1961]. The technique of random time change goes back to Volkonskii [1958] and was used by him and by Ito and McKean [1965] to study one dimensional diffusions. The Cameron-MartinGinsanov formula was derived in a particular case in Cameron and Martin [1953]. Girsanov [1960] and Skorohod [1957] have made computations of RadonNikodym derivatives for stochastic processes. Chapter 7. Tanaka [1964] and Krylov [1966] constructed a Markov semigroup (without proving uniqueness) for an arbitrary set of time homogeneous elliptic coefficients assuming for smoothness only the continuity of the second order terms. Some of the results proved in this chapter have been extended to diffusions on manifolds in Yang [1972]. Chapter 8. The results of this chapter have their origin in Watanabe and Yamada [1971]. Chapter 9. Most of the estimates on the transition probabilities that we derive in this chapter can also be derived by using standard techniques in the theory of partial differential equations. See for instance the works of Fabes and Riviere [1966] and Ladyzenskaja, Solonnikov and UraFceva [1968]. Chapter 10. When the dimension of the space is 1, Feller obtained complete results concerning conditions for explosion. These can be found in Ito-McKean [1965]. In the muhidimensional case the main results are due to Khasminskii [I960]. They may be found in the book by McKean [1969]. Chapter 11. The idea of proving limit theorems for processes goes back to Erdos and Kac [1947] and Donsker [1951]. The papers by Prohorov [1956] and Skorohod [1956] were very important in the development of the subject. The technique we use to prove limit theorems is very close in spirit to the proof of Trotter [1959] of the central limit theorem. These same martingale techniques have been used effectively to study the central limit problem for random variables on a Lie group in Stroock and Varadhan [1973]. Chapter 12. The problem of Markov selections was first solved by Krylov [1973]. Chapter 13. The results on singular integrals and their application to partial differential equations go back to Calderon and Zygmund [1952] in the elliptic case and Jones [1964] in the parabolic case. The results on singular integrals go back even further to Mihlin [1938] and Giraud [1932]. There is, of course, a considerable amount of material on Markov processes in general and diffusion processes in particular that has not been covered in this book. The expository articles by Dynkin [1962] and Williams [1974] contain excellent references. The books by Dynkin [1965], and Gikhman and Skorohod [1974], [1975], [1978] have a much wider scope than is attempted here. The two volumes by Friedman [1976] deal with several applications of stochastic integral
330
Bibliographical Remarks
equations besides developing the theory. Stratanovich [1968] and Lipster and Shiryayev [1977] and [1978] deal with the problems of control and filtering of Markov processes. For a more detailed study of one dimensional diffusions Ito and McKean [1965] and Dynkin [1959] are good sources. For diffusions with boundaries Watanabe [1971], Sato and Ueno [1965], Wentzel [1959] and Stroock and Varadhan [1971] are some of the relevant references.
Bibliography
Alexandrov, A. D. 1963 The method of "supporting" mapping in the study of the solutions of boundary problems. Outlines Joint Symp. Partial Differential Equations (Novosibirsk, 1963) 297-302. Billingsley, P. 1968 Convergence of Probability Measures. New York, John Wiley. Burkholder, D. and Gundy, R. 1970 Extrapolation and interpolation of quasiUnear operators on martingales. Acta Mathematica 124, 249-304. Calderon, A. P. and Zygmund, A. 1952 On the existence of certain singular integrals. Acta Math. 88, 85-139. Cameron, R. H. and Martin, W. T. 1953 The transformation of Wiener integrals by nonlinear transformations. Trans. Amer. Math. Soc. 75, 552-575. Davis, B. 1970 On the integrability of the martingale square function. Israel Jour. Math. 8, 187-190. Donsker, M. D. 1951 Four papers in probability. Mem. Amer. Math. Soc. No. 6. Doob, J. L. 1952 Stochastic Processes. New York, John Wiley. Dynkin, E. B. 1952 Criteria for continuity and lack of discontinuities of the second kind for trajectories of a Markov stochastic processes. Izv. Akad. Nauk SSSR. Ser. Mat. 16, 563-572. 1959 One dimensional continuous strong Markov processes. Theor. Prob. and Appl. 4, 1-52. 1960 Foundations of the Theory of Markov Processes (English translation). Oxford-London-New York-Paris, Pergamon Press. 1962 Markov processes and problems in analysis. Proc. Int. Cong, of Math. Stockholm, 36-58. 1965 Markov Processes, Vols. 1, 2. Berlin, Springer-Verlag. Erdos, P. and Kac, M. 1946 On certain limit theorems of the theory of probability. Bull. Amer. Math. Soc. 52, 292-302. Fabes, E. 1966 Singular integrals and partial differential equations of parabolic type. Studia Math. XXVIII, 6-131. Fabes, E. and Riviere, N. M. 1966 System of parabolic equations with uniformly continuous coefficients. Jour. Analyse Math. XVII, 305-335. 1966a Singular integrals with mixed homogeneity. Studia Math. 27, 19-38.
332
Bibliography
Fefferman, C. and Stein, E. M. 1972 H" spaces of several variables. Acta Mathematica 129, 137-193. Feller, W. 1936 Zur Theorie der Stochastischen Prozesse [Existenz und Eindeutigkeitssatze]. Math. Ann. 113, 113-160. Freidhn, M. 1967 On small (in the weak sense) perturbations in the coefficients of a diffusion process. Theory of Prob. and Appl. 12, 487-490 (English Translation). Friedman, Avner 1976 Stochastic Differential Equations and Applications, Vols. 1 and 2. New York, Academic Press. Friedman, A. 1964 Partial Differential Equations of Parabolic Type. Englewood Cliffs, NJ, Prentice Hall. 1969 Partial Differential Equations. New York, Holt, Rinehart and Winston. Garsia, A. 1973 Martingale Inequalities. Reading, MA, W. A. Benjamin. Garsia, A., Rodemick, E., and Rumsey, H., Jr. 1970-71 Indiana Univ. Math. Journal 20, 565-578. Gihman, I. I. and Skorohod, A. V. 1973 Stochastic Differential Equations. New York, Springer-Verlag. 1974, 1975, 1979 The Theory of Stochastic Processes, Vols. I, II, and III. Berlin, Springer-Verlag. Giraud, G. 1932 Sur une extension de la theorie des equations integrates de Fredholm. C. R. Acad. Sci. Paris, 195, 454-456. Girsanov, I. V. 1960 On transforming a certain class of stochastic processes by absolutely continuous substitution of measures. Theory of Prob. and Appl. 5, 285-301. 1961 On Ito's Stochastic Integral Equation. Soviet Mathematics, Vol. 2, pp. 506-509. Halmos, P. R. 1950 Measure Theory. New York, Van Nostrand. Hille, E. and Phillips, R. S. 1957 Functional Analysis and Semigroups. Providence, American Math. Soc. Hormander, L. 1960 Estimates for translation invariant operators in Lp spaces. Acta Math. 104, 93-140. Ito, K. 1951 On stochastic differential equations. Mem. Amer. Math. Soc. 4, xxx-xxx. Ito, K. and McKean, H. P., Jr. 1965 Diffusion Processes and Their Sample Paths. Berlin, Springer-Verlag. Jacod, J. and Yor, M. 1976 Etude des solutions extremales et representation integrate des solution pour certains problemes de Martingales. C. R. Acad. Sciences, Paris T. 283 Series A, pp. 523-525. John, F. and Nirenberg, L. 1961 On functions of bounded mean oscillation. Comm. Pure Appl. Math. 14, 785-799. Jones, B. F. 1964 A class of singular integrals. Amer. Jour, of Math. 86, 441-462.
Bibliography
333
Khasminskii, R. Z. 1960 Ergodic properties of recurrent diffusion processes and stabilization of the solution of the Cauchy problem for parabolic equations. Theor. Prob. and Appl. 5, 179-196. Kinney, J. H. 1953 Continuity properties of sample functions of Markov processes. Trans. Amer. Math. Soc, 74, 280-302. Kolmogorov, A. N. 1931 Uber die Analytischen Methoden in der Wahrscheinlichkeitsrechnung. Math. Ann. 104, 415-458. 1956 Foundations of the Theory of Probability. New York, Chelsea. Krylov, N. V. 1966 On quasi diffusion processes. Theor. Prob. and Appl. 11, 373-389. 1973 The selection of a Markov process from a Markov system of processes (Russian). Izv. Akad. Nauk SSSR, Ser. Mat. 37, 691-708. 1974 Certain bounds on the density of the distribution of stochastic integrals. Izv. Akad. Nauk, 38, 228-248. Kunita, H. and Watanabe, S. 1967 On square integrable martingales. Nagoya Math. Journal 30, 209-245. Kuratowski, C. 1948 Topologie, Vol. 1. Warzawa, Monografie Matematyczne. Ladyzenskaya, O. A., Solonnikov, V. A., and UraFceva, N. N. 1968 Linear and quasilinear equations of parabohc type. Providence, Amer. Math. Soc. Translations of Math. Monographs, No. 23. Levy, P. 19-8 Processus Stochastiques et Mouvement Borwnien. Paris, Gauthier-Villars. Lipster, R. S. and Shiryayev, A. N. 1977, 1978 Statistics of Random Processes, Vol. I, Theory, and Vol. II, Applications. Berlin, Springer-Verlag. McKean, H. P. 1973 Geometry of Differential Space. Ann. Prob. 1, 197-206. McKean, H. P., Jr. 1969 Stochastic Integrals. New York-London, Academic Press. 1975 Brownian Local Times. Advances in Mathematics, Vol. 16, pp. 91-111. Meyer, P. A. 1966 Probability and Potentials. Wahham, MA, Blaisdell. 1974-75 Seminaire de Probabilites, Universite de Strasbourg. Springer-Verlag, Lecture Notes No. 511. Mihlin, S. G. 1938 The problem of equivalence in the theory of singular integral equations. Mat. Sbornik N.S., 45, 121-140. Neveu, J. 1975 Discrete Parameter Martingales (translated by T. P. Speed). Amsterdam, North-Holland. Oleinik, O. A. 1966 Alcuni risultati sulle equazioni lineari e quasi lineari ellitico—paraboliche a derivate parziah del second ordine. Rend. Classe Sci. Fis. Mat., Nat. Acad. Naz. Lincei, Ser. 8, 40, 775-784. Parthasarathy, K. R. 1967 Probabilit) Measures on Metric Spaces. New York, Academic Press.
334
Bibliography
Phillips, R. S. Sarason, L. 19xx Elliptic Parabolic Equations of Second Order. J. Math, and Mech. 17, 891-917. Prohorov, Yu. V. 1956 Convergence of stochastic processes and hmit theorems in probability theory. Theor. Probability Appl. 1, 157-214 (English Translation). Protter, M. H. and Weinberger, H. F. 1967 Maximum Principles in Differential Equations. Englewood Cliffs, NJ, Prentice-Hall. Pucci, C. 1966 Limitazioni per soluzioni di equazioni ellitiche. Annali di Matematica pur ed applicata. Series IV, LXXIV, 15-30. Riviere, N. M. 1971 Singular integrals and muhiplier operators. Arkiv fiir Matematic 9, 243-278. Sato, K. and Ueno, T. 1965 Multidimensional diffusions and the Markov process on the boundary. Jour, of Math., Kyoto Univ. 4, 529-605. Skorohod, A. V. 1956 Limit theorems for stochastic processes. Theor. Prob. and Appl. 1, 261-290. 1957 On the differentiability of measures which correspond to stochastic processes, I, Processes with independent increments. Theor. Prob. and Appl, 2, 407-432. 1961 On the existence and uniqueness of solutions for stochastic differential equations. Sibirsk. Mat. Zur. 2, 129-137. 1961 Existence and uniqueness of solutions to stochastic diffusion equations. Sibirsk Math. Zh. 2, 129-137. Slutzky, E. E. 1949 Some statements concerning the theory of random processes. Publications of Middle Asian University, Mat. Ser. (5) 31, 3-15. Stratanovich, R. L. 1968 Conditional Markov Processes and Their Applications to the Theory of Optimal Control. New York, American Elsevier Publ. Co. Stroock, D. W. 1973 Applications of Fefferman-Stein type interpolation to probabihty theory and analysis. Comm. Pure Appl. Math. XXVI, 477-496. Stroock, D. W. and Varadhan, S. R. S. 1969 Diffusion processes with continuous coefficients, I and II. Comm. Pure Appl. Math. XXII, 345-400, 479-530. 1970 On the support of diffusion processes with applications to the strong maximum principle. Proceedings of Sixth Berkeley Symp. on Mathematical Statistics and Probability, Vol. Ill, pp. 333-360. 1971 Diffusion processes with boundary conditions. Comm. Pure Appl. Math., XXIV, 147-225. 1972 On degenerate elliptic-parabolic operators of second order and their associated diffusions. Comm. Pure Appl. Math. XXV, 651-714. 1973 Limit theorems for random walks on Lie groups. Sankhya 35, Ser. A, 277-294. Sussman, H. 1977 An interpretation of stochastic differential equations as ordinary differential equations which depend on the sample point. Bull. Amer. Math. Soc. 83, 296-298. Tanaka, H. 1964 Existence of diffusions with continuous coefficients. Mem. Fac. Sci. Kyushu Univ., Ser. A., 18, pp. 89-103.
Bibliography
335
Trotter, H. F. 1959 An elementary proof of the central limit theorem. Archiv der Mathematik 10, 226-234. Varadarajan, V. S. 1958 Weak convergence of measures on separable metric spaces. Sankhya 19, 15-22. 1965 Measures on topological spaces. Transl. Amer. Math. Soc, Ser. II, 48, 161-228. Volkonskii, V. A. 1958 Random time changes in strong Markov processes. Theor. Prob. and Appl. 3, 310-326. Yang, W. Zu. 1972 On the uniqueness of diffusions. Z. Wahr. Verw. Geb. 24, 247-261. Watanabe, S. 1971 On stochastic differential equations for muhidimensional diffusion processes with boundary conditions. Jour, of Math., Kyoto Univ. 11, 169-180. Watanabe, S. and Yamada, T. 1971 On the uniqueness of solutions of stochastic differential equations. Journal of Math, of Kyoto Univ. 11, 155-167. Wentzel, A. D. 1959 On boundary conditions for multidimensional diffusion processes. Theor. Prob. and Appl. 4, 164-177. Wiener, N. 1923 Differential space. J. Math. Phys., 2, 131-174. 1930 The homogeneous chaos. Am. J. Math. 60, 897-936. Williams, D. 1974 Brownian motions and diffusions as Markov processes. Bull. Lond. Math. Soc. 6, pp. 257-303. Yamada, T. and Watanabe, S. 1971 On the uniqueness of solutions of stochastic differential equations, II. J. Math. Kyoto Univ. 11, 553-563.
Subject Index
Backward equation 3 Brownian motion 56 ,5-
57
, J-dimensional 57 , estimates on modulus of continuity of , martingale characterizations of 84 , strong Markov property of 59
63
Cameron- Martin—Girsanov formula 152, 154, 155 Compactness criteria for Markov chains to diffusions 41 measures on C([0,oc), /?") 39 weak convergence of probability measures on a Polish space, Prohorov 's condition 10 Conditional expectation and probability 12 Conditional probability distribution 12 , regular (r.c.p.d.) 16 Consistent distributions 37 Continuity of trajectories, criteria for , , of Garsia et al. 47 , , Kolmogorov 's theorem 51 Doob's martingale theorems , martingale inequality for stopping time 34
21
Extension of measures , Kolmogorov's theorem on, for consistent distributions 19 Tulcea's theorem on 17 Feller continuity 152 Feynman - Kac formula 114 Finite dimensional martingales 145 Forward equation (Fokker-Planck equation) Homogeneous chaos
2
120
Integration by parts formula for martingales Interpolation theorem 311,313 Itd's formula 105, 115
25
Ito processes , definition of 92 , estimates for 86 , relationship of to random time change 112 , stochastic integral representation of 108 , stochastic integration with respect to 96 Itd's stochastic differential equations (see Stochastic differential equations) ltd uniqueness 195 Kolmogorov's criterion for continuity 51 Kolmogorov's extension theorem 19 Local time, existence and properties of
117
Markov processes with given transition probability function 52,54 , continuous 54 ,(time-) homogeneous 54, 158 Markov selections of solutions (in non-unique cases) 290 Martingales 19 —and conditional probability distributions 28 —, Doob's theorems for 21, 34, 44 —, super- and sub- 20 Martingale problem 13 8 , consequences of uniqueness of: Markov and strong Markov properties 145 , examples of non-uniqueness 169 , Feller continuity 152 , global case, 187 in one dimension 192 in two dimensions 193 , local case 192 , proof of uniqueness, localization procedure 161 — —, properties under conditioning 140 , relation to ltd stochastic differential equations 152 , special techniques for proving uniqueness: random time change 159
Subject Index
338 Martingale problem [cont.] , existence of solutions to 143 , uniqueness of solutions to 148 , well-posed 138 Maximum principles , strong (Nirenberg's version) 169 , weak 65 Non-explosion of diffusions, conditions for Polish spaces 7 Prohorov's theorem 9, 10 Progressively measurable functions
255
19, 20, 45
Random time change 112, 159 Regular conditional probability distribution (r.c.p.d.
ofPl^j)
16, 34
Right-continuous function
19
Semi-groups 70 Simple functions 92 Singular integrals 306 Square root of a non-negative definite matrix valued function 131
Stochastic integration with respect to an ltd process 96 Stochastic differential equations , existence and uniqueness of solutions 124 , relationship to uniqueness of solutions to the martingale problem 202 Stopping times 22 Strong Feller property 123, 204 Strong Markov property 145 for Brownian motion 59 Submartingales 20,29 Submartingale problem 167 Supermartingales 20 Time-space process 166 Transition probability function 51 , convergence of densities 282 estimates on densities 220, 231, 240, 247 Tulcea 's extension theorem 17 Weak topology on probability measures Wiener measure 56
7, 9