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!
(fi-I)2) < (/3 _ 1)-2.1-2E(B(T A S2 A T) - B(T A S1 A T))2. Now if R1 <- R2, are both bounded stopping times, we have 2). The function that shows that this inequality is sharp is simple. Let fe(z) = p(z)l = ((1 + z)l(l - z))fi. Since P E H" for p e (0, 1), > < A-22o) (l K)2) C
E(B(R2) - B(R1))2 = E(B(R2)2 - B(R1)2) = E(R2 - R1), so the above 1)-2A-2 E((T A S2 A T) -(T A S1 A T)) < ( _ 1)-22-2(S2)2P(S1 < oo) 1)-262P(B* > A),
proving the first inequality. To prove the other inequality Let S1 = inf{t : (T A t)112 > .1}. Let S2 = inf{t : (T A t) 112 >
Let T = inf{t : IB(T A t)I > 6.1}.
Again, it is easy to check that P(T 1/2 > #A, B* < 62) < P(T A S2 A
T - T A S1 A T> $2)2 - 22)
<($2_1)-11-2E(TAS2A T-TAS1 A T), and using the stopping time result mentioned in the first part of the proof, it follows that the last expression above _ (/22 _ 1)-11-2E(B(T A S2 A T)2 - B(T A S1 A T)2)
<
($2
-
1)-11-2(S2)2P(S1 < oo)
_ ($2 _.1)-162P(T1n > 2), proving the second inequality.
Remark: The inequalities above are called "good 2" inequalities, although the reason for the name is obscured by our formulation (which is from Burkholder (1973)). The name "good .1" comes from the fact that early versions of this and similar inequalities (see Theorems 3.1 and 4.1 in Burkholder and Gundy (1970), or page 148 in Burkholder, Gundy, and Silverstein (1971)) were formulated as P(f > A) < C$,KP(g > A) for all A that satisfy P(g > A) < KP(g > $2), $, K > 1. The next result shows why we are interested in good A inequalities. First, we need a definition. A function cp is said to be moderately increasing if /P is a nondecreasing function with 9(0) = 0 and if there is a constant K such that
p(21)
(i) T(x)=x",0
MA
6
Hardy Spaces and Related Spaces of Martingales
Proof for (ii)
Since log+ xy S log' x + Iog+ y,
2x + 2x log+ 2x
x+xlog+x
<2+
2 log 2
1+log+x,
The bound now follows by considering two cases, x < 1 and x >_ 1. (2)
If qp is a moderately increasing function, then there are constants c, C e (0, 00) (that depend only on the growth rate K) such that cEgo(T'12) < Eq(B*) < CEgp(T112).
Remark: We will only use the result for 9(x) = xp, but it is nice to know that the only property of xp we need for the proof is that (p(22) < Kq(2) for all
2>0. Proof To prove the result, it suffices to show: (3)
If X, Y > 0 satisfy
P(X>22, Y
Etp(X) < CE9(Y). Proof It is enough to prove the result for bounded gyp, for if the result holds for cp A n for all n > 1, it also holds for gyp. T is the distribution function, a measure on [0, oo) that has
T(h) = foodtp(2) = J
1(h>A)d(p(2). 0
If Z is a nonnegative random variable, taking expectations and using Fubini's theorem gives E4p(Z) = E f
l(Z>A) dgp(A) =
JP(Z> A) d4 (2).
0
From our assumption it follows that
P(X > 22) = P(X > 22, Y:5 52) + P(Y > b2) < 62P(X > A) + P(Y > 62), or integrating dq(2), E4 (2-1X) =
f
P(2-1X < 2) dq (2)
0
< 62E9 (X) + Egp(b-1Y).
Pick 62K < 1 and then pick N > 0 so that 2' > b-1. From the growth condition and the monotonicity of gyp, it follows that
6A Equivalence of Ho too Subspace of .No
115
E9(6-'Y)!5 KNEcO(Y)
Combining this with the last inequality and using the growth condition gives E(p(X) < KEcp(2-1X) < K62Etp(X) + KN+1Eco(f
Solving for E.p(X) now gives KN+1
Ecp(X) <
1-
K62Ecp(Y),
proving (3) and hence (2).
Applying (2) to the case cp(x) = xP and recalling the results of Section 2.11, we get the following inequality that will be useful in studying the spaces Al P.
(4)
There are constants c, C E (0, oo) (that depend only on p) so that for all
0
cE<X)Pl2 < E(X*)P < CE<X>?l2.
As the reader might expect, the constants in (4) are not very good. Ifp = 2, then EX.,, = E<X>, so it follows from (1) of Section 6.2 that E<X%, = EXX < EIX* 12 < 4EXW = 4E<X),,,,,
but in the proof above, K = 4, so if we take S = 1/3 and N = 2, then
C=
KN+1
1 -K62
= 115.2.
Remark: There are other ways of proving (4) directly; see Getoor and Sharpe (1972) for an interesting proof using stochastic integration.
6.4 Equivalence of H" to a Subspace of V'
l
Let B, be a complex Brownian motion starting at 0 and let T = inf{t : I B,I = 1}. In this section, we will show that the mapping f - Ref(B), t < T, maps Ho = { f E HP : f (O) = 0} one-to-one into Al,! = {X: X is a local martingale on [0, T) and X,* = sup,,, I X, I E LP}, and that furthermore, the HP norm off is equivalent to the Al,! norm of its image, that is, if we let u = Ref, U, = u(B,), t < T, and U* = sup«TIUI, then: (1)
There are constants c, C E (0, oo) (that depend only on p) such that for all fe HoP,
cElU*IP
15(t
6
(2)
There is a constant Cc (0, oo) (that depends only on p) such that for all feHP,
Hardy Spaces and Related Spaces of Martingale@
dd(f) < EIF*IP < Cdd(f) Remark: It is curious that the best constant we can obtain from the proof given below is C = e (independent of p).
Proof If r <'l and T, = inf{t : IBI = r}, then IF(T,)I < F*, so we have I
ER I f(re`B) I
P dir(O) = EI F(T,) I P < E(F*)P.
Taking the supremum over r < 1 gives dd(f) < E(F*)P. To prove the other inequality takes some work. We start with the trivial case, p > 1. Since f is analytic in D and bounded on D(0, r) = {z : Izl < r}, it follows from results in Chapter 2 that f(B(t A T,)) is a complex martingale and, hence, that I f(B(t A r,))I is a submartingale. Noting that I f(B(-r,))I e LP, p > 1, and applying Doob's inequality, we get EIF*(T,)IP <(i-f7i) Ej
f(B(T,))IP
Letting r T 1 and using the monotone convergence theorem gives P
EIF*IP <- (i-:P::--,) dP(f)-
proving (1) in the case p > 1 with a constant slightly larger than the one advertised in the remark, since x dx log (x
1)s =
dx (x log x - x log (x - 1))
= log x - log(x - 1) - (x - 1)-1
y-1 - (x - 1)-1 dy < 0.
f,;x 1
To prove the result when 0 < p < 1, we will have to be more devious. First consider 1/2 < p < 1 and let us now suppose f(0) = 1. Since f(B,j is a time change of a complex Brownian motion starting at 1, P(f(B1) = 0 for some 0 < t < T) = 0, and we can define a pathwise square root GG by requiring G
to be continuous and to have Go = 1, and GG = F for all 0 <- t < T. As we mentioned in Section 5.3, (3)
G, I < T, is a local martingale
and furthermore (4)
If r < 1, G(t A T,) is a martingale. (See the remarks after the proofs of (3) and (5) in Section 5.3.) With (4) proved, the rest is easy. Since
6.4
157
Equlvilsoce of He to o Subcpoce of .+Y"
EIG(r,)I2p =
zn
f
lf(re`B)11dm(0) < dp(f)
0
and 2p > 1, it follows from Doob's inequality that 12p2p
1
)2"EIGI2P
Taking a supremum over r and applying the monotone convergence theorem gives
EI F*I P = EIG*(r)I2p <
12p
2p2p- 1
I
dp(f)
This proves the result when f(O) = 1. If f(O) # 0, applying the result above to
g(x) = f(x)/f(0) proves the inequality for f. To extend the result to f with f(0) = 0, consider g,,(x) = f(x) + c and let a -. 0. The argument above can be extended top > 1/n by taking nth roots, with the result that p
EIFF*Ip < np pT1dp(f) Letting n - oo, the last inequality gives El F* I p < edp(f ),
proving (2) with C = e. The result in (2) shows that the Hp norm of f and the .,#,! norm of f(B,), t < r, are equivalent. As we mentioned above, we want to go one step further and consider Hp as a space of real local martingales by using the mapping f - Re f(B,), t < T. Since every harmonic function u has a conjugate harmonic function u defined by the requirements that 4(0) = 0 and f = u + iu is analytic in D, the mapping is one-to-one on He = { fe Hp : f(0) = 0}. The next result shows that if we restrict our attention to this subspace, then the Hp norm off and the .A'' norms of Ref(B,), t < r, are equivalent (i.e., (1) holds). (5)
There is a constant C e (0, oo) (that depends only on p) such that for all f e Ho, EIU*Ip < EIF*Ip < CEIU*IP.
Proof Since l a l < (a2 + b2)1/2 for all a, b, we have I u (z) l < I f (z) I for all z, proving the inequality on the left. To prove the other inequality, we observe that if we let U = u(B,), t < r, then it follows from results in Sections 2.11 and 6.4 that E(U*)p < CEpl2 pl2 < CE(U*)p (here, as before and ever shall be, the value of C is unimportant and will change
1 ui
6
Hardy Spaces and Related Spaces of Martingales
from line to line). It is trivial that E(F*)P 5 2P(E(U*)P + E(U*)P).
Combining this with the other three estimates proves (5).
Given the results above and in Chapter 4, it is now easy to prove the "maximal function characterization of Hr" due to Burkholder, Gundy, and Silverstein (1971). Let S,,(0) be the convex hull of the disk {lzl < a} and the point e`° and let (Nau)(0) = sup{lu(z)I : zeS,,(0)} be the nontangential maximal function. A simple generalization of the argument given at the end of Section 4.5 shows that
SSJ
Po (Osup, u(B.) > .11 < Cl {O: Nau(0) > All,
where C depends only on a, and integrating gives zn
Eo(UU*)P < C
l NQul P(0) d7r(0) 0
(a result that holds for any measurable u). To prove the other inequality, we use (2) from Section 4.3 (which generalizes easily from H to D). This result implies that there is an E > 0 such that if z e S8(0), then
Po(B,, 0 < t < r, makes a loop around z) > E, so it follows from the maximum principle that Pa(U* > Nau(0)/2) >_ E,
and hence that E0(U*)P > E f lNaull(0)dn(0). Combining this with the other inequality and (1) above gives the maximal function characterization of HP cdd(f) <- fo z l NaulP(0)dn(0) < Cdd(f)
6.5 Boundary Limits and Representation of Functions in H" Since HP c N, it follows from (2) in Section 5.3 that (1)
If fe HP, then the nontangential limit off exists at a.e. point of 8D.
In this section, we will investigate the relationship between f (defined in D) and its nontangential limit (defined on 8D), which we will also denote by f.
6.5
159
Boundary Limits and Representation of Function. la HP
There are two reasons why we want to do this: (a) The mean convergence of H" functions to their boundary values (3) and the consequent Poisson integral representation (5) are simple consequences of the equivalence established in Section 6.4, and (b) for developments below, it is nice to know that we can consider a function f e H" as a function in LP(8D). Our first topic in this section is LP convergence to boundary values. As
usual, we start with the probabilistic result and then deduce its analytical counterpart. (2)
Let f c- HP and let F = f(B), t < T. If T. is any sequence of times TT, then as n -- oo, f(BL) a.s. and in LP. Proof The a.s. convergence is a consequence of (2) in Section 5.2. (2) of Section
6.4 shows that F* E LP. Since IF(T) - f(Bz) I < (2FT*)P and F(T) -f(B) a.s., the LP convergence follows from the dominated convergence theorem. (3)
Iffe HP, then as rT 1, I
I .f (re `B) - .f (e `e) I P do (0)
0.
Proof It suffices to show that the result holds for any sequence r T 1. If we let T. = inf {t: I B, I > and apply (2), it follows that -f(B) in LP, or, BT), if we let u denote the distribution of
1-f(y)IPdvn(x,y)- 0. This conclusion is similar to the one desired, but it is different enough to make it painful to obtain one from the other. In the face of this hard difficulty, we will take a soft solution. We recall the following result from real analysis: (4)
If X E LP and X. --* X a.e., then X,, -+ X in LP if and only if EI
P -- EI XI P.
We leave the proof as an exercise for the reader (see Chung (1974), page 97), and then we are done, because
F(T) -f(B.)
in
LP
implies J J
implies
f
I.f(rne's) - f(eie)I Pdxt(0)
0.
The last result is valid for p > 0. When p > 1, we can use the LP convergence to express f in terms of its boundary values.
1M)
6
(5)
Let k9(z) be the probability density (w.r.t. n) of exiting D at e`9 starting from z.
Hardy Spacer and Related Spaces of Martingales
IffeHP f(z) =
p > 1, then d7r(0).
Proof From (2) of Section 6.4, it follows that EIF,*IP < oc, and since p > 1, F, t < t, is uniformly integrable. If v is a stopping time, the optional stopping theorem implies that .f(Ba) = E(.f(BJ13o),
and it follows from the strong Markov property that E(f(BT)1 3o) = Jko(B0)f(eio) dn(0).
The last two results imply that the equality in (5) holds for a.e. z e D. The left-hand side, f(z), is clearly a continuous function of z. In Section 3.3, we showed that the right-hand side is harmonic in D, and hence continuous, so the equality holds for all z e D.
The equality in (5) does not hold when p < 1. The half-plane mapping p(z) = (1 + z)/(1 - z) has boundary limit Example 1
p(e`B) = i sin 0/(1 - cos 0),
so there are two problems :
(i) Rep (z) > 0 in D, but Jko(z)Rep(eio)dir(O) = 0
(ii) Isin 01/(1 - cos 0) - 2/101 as 0 - 0, so Jke(z)IImP(eio)Idir(O) = oo.
The resolution of the first problem is obvious. Rep = ko, so Rep (z) = Jko(z)d50(O).
where So is a point mass at 0. This example suggests that we might generalize
the representation in (5) to allow measures on 8D that are not absolutely continuous w.r.t. it. The next result shows that this generalization does not enlarge by very much the class of functions that can be represented. (6)
The following three classes of functions are the same :
(i) the set of u that can be written as
6.5
Boundary LImlts and Representation of Function In H'
161
U(Z) = fko(z)d!L(o),
where p is a signed measure with finite variation (ii) the set of u that can be written as a difference of two positive harmonic
functions (iii) h' = the set of harmonic functions u with sup flu(reio)IdO < oo. r<1
Proof (i) (ii) (iii) is trivial. Since the proof of (iii) (i) is very similar to the proofs given in Sections 3.3 and 3.5, it is left as an exercise for the reader. Remark: For developments in Section 6.8, it is useful to know that (6) implies that any ueh1 can be written as u1 - u2, where u1, u2 > 0 and 1lull1 = llu1 11 + 1
ll u2ll 1. To prove this, we observe that if u = Jko(z)dP(O). then llull 1 = the variation p = sup,, µ(A) - µ(A`).
The last result completes our consideration of boundary limits of functions in HP. Since we will deal with p >- 1 for most of the rest of this chapter, the important results to remember are (1) and (5). They show that a function feHP, p > 0 has a nontangential limit at a.e. point of 8D and that if p > 1, the values off in D can be recovered from the boundary limits, so we can think off in HP as being a function in LP(8D, it) (and we will do this when we prove the duality theorem). A similar viewpoint is possible for local martingales Xc- ,`f?, p >_ 1. Since this will simplify things below, we will take a few minutes now and spell out the details. X,* e LP, so standard arguments imply that if we let
_
X,
t
t} X, lim
t
IT
then Y is a martingale, Ye #P (the space defined in Section 6.2), and Y can be reconstructed from its limiting value by Y, = E(Y.l JFt).
The last representation is the probabilistic analogue of the Poisson integral representation. When X, = f(B,), t < r, and feHP, p > 1, the relationship is closer than an analogy, since results above imply that Y, = f(BT) (the righthand side being the nontangential limit off evaluated at BT).
Notes: Result (3) about the mean convergence of f(re`B) to f(e`B) was first proved by F. Riesz (1923) (see, for example, Duren (1970), pages 20-22). The key to Riesz's proof was the following factorization theorem :
162
(7)
6
Hardy Specs and Related Specs of Mertlnoalse
Every function f in H° can be written asf(z) = b(z)g(z), where Ib(z)I < 1 and g e H" is a function that does not vanish in D. Once (7) was established, (1) followed from known results. The function b has boundary limits, since it is bounded, and since g is never 0, we can pick n > 1/p and consider g'1" to reduce the result to the easy case where p > 1. The reader should note that taking nth roots to reduce to the trivial case p > 1 was also the key to the proof of (3) in Section 6.2, but in that proof, we
used the fact that Brownian motion did not hit zero to construct a pathwise nth root, so we did not have to factor out the zeros.
6.6 Martingale Transforms Martingale transforms are a natural generalization of the following: Example 1 Let f = u + iv be an analytic function with f(0) = 0, and let B, be a complex Brownian motion starting at 0. Ito's formula implies that u(B,) = f Vu(BS) dB. 0
v(B,) = JVv(Bs).dBs, 0
and the Cauchy-Riemann equations say that Vv =
(_0 0)Vu,
so if we let
H,=Vu(B,) A= (-0
11 0
'
then we can write
u(B) =
JJJ.dB , 0
v(B) = JAI-I.dB3. 0
The last equation obviously makes sense if B is a d-dimensional Brownian motion, H is a locally bounded (R'-valued) predictable process, and A is any d x d matrix. To define the transform of a general local martingale X in this setting, we now recall that since we have assumed that our filtration is generated by a Brownian motion, it follows from results in Section 2.14 that
Xt=Xo+ f 0
Mmrtingde TnMformw
6.6
163
We can therefore define the transform of X by A as
(A*X)t= (whenever this makes sense, e.g., if H is locally bounded). In this section, we will study properties of martingale transforms as mappings between the V' spaces. The results we prove here are analogues of classical results about conjugate functions that we will prove by probabilistic methods in Section 6.7. The first and most basic result is: (1)
If p > 0, X -. A * X is a bounded linear transformation from #P to M". Proof If Xt = 0
then
(A*X)t= f 0
so
IAH s12ds fo'
< C J, I HHIZ ds 0
= C<X>t, where
C= sup{IAYI2:IyI = 1}, so the desired conclusion follows from the equivalence of norms demonstrated in Section 6.3. When p > 1, the norms on . #P and it" are equivalent, so we have : (2)
If p > 1, X A * X is a bounded linear transformation from IC" to IC". The next example shows that this is false when p = 1. Example 2
Let Br be a two-dimensional Brownian motion starting at 0, let
i=inf{t:B' = -1}, and let Xt=BTnL= If we let A = (1
lot
.dBs.
\0 0 I
(Or, if you want, \l
0
then
I I 1
164
6
Hardy Spacer and Related Spacer of Martlnpal"
dB., (A*Xj-JO^T
(, ). 1
Now X, = B;A T > -1 and it is trivial that if a >- -1, l a l < 2 + a, so
EIX,I <2+EX,=2 and Xe.(1. On the other hand, it follows from results in Section 1.9 that (A * X). = B, has a Cauchy distribution
P((A*X) X)= x1
1
n1+x2'
°°
so E(A * X)W = oo, and Fatou's lemma implies that lim inf E(A * X )t+ > oo. t- 00
In the last example,
IIA*XII,r =supEIA *X1, = oo, I
but it doesn't miss by much
2 fo dx2,.,ny-1
I >y)=n
1
as y -- oo. The next result, first proved by Burkholder (1966), shows that this is the worst possible behavior for XE.%r1. The statement and proof given below are from Burkholder (1979a). (3)
If
Proof Doob's inequality implies that AP(X* > A) < supEJX,l = IIXII,, f
so it suffices to estimate the probability of {Y* > A, X* < Al. To do this, we
observe that by stopping at T = inf {t: <X>, > n}, we can suppose that E<X>0 < cc, that is, Xe.#2. Let i = inf {t :1X1 > Al. It is trivial that if we let Y,* = sup fl l Y 1 : s :!g t}, then
AP(Y* > A, X* < A) < ,1P(Y* > A).
Applying Doob's inequality to the submartingale Y' T(Ye 1f2, since
22P((Y*)2 > A2) < EY, . Now X, Ye.,&2, so
EY2=E
6.6
MYrtlnade'rromformn
165
X
Y
Figure 6.1
To finish up, we observe that IX,1 < 2 and IX,I is uniformly integrable, so
EX, <2EIXzl Combining the inequalities above shows that AZP(Y* > A, X* < 2) < AIIXILI
and proves the desired result. A simple example shows that 2 is the best possible constant in (3). Example 3 Let B, be a one-dimensional Brownian motion, and make the following definitions (for a picture, see Figure 6.1):
T,=inf{t:IB,I=l} Tz =inf{t> T, :IB,-B(Ti)I = l}
T3=inf{t>TZ:B,=0 orIB,-B(T2)I=2}
B(T1)=for-1 B(T2) = 2, 0, or - 2 B(T3) = 4, 0, or -4
166
6
Hardy Spac.u and Related Spaces of Martingales
W (s, w)
O(+', (0)
1
1
1
-1
1
1
0
0
S
r
X, =
cps dB,
Y=
0
Os dB.,. 0
Since (p2
we have <X> _
X
B 1,2,4 1,2,0 1,0,2
1,0,-2
1,2,4 1,2,0 1,0,0 1,0,0
Y 1,0,2
1,0, -2 1,2,2 1,2,2
Y* 2 2 2 2
with a similar table for a = - 1. From the last computation it follows that supElX,l = El XXl = 1 t
and
P(Y*>2)=1, so forA=2, 2P(Y* > A) = 2supEpXc1,
showing that (3) is sharp.
(1), (2), and (3) are the main results on martingale transforms, but, of course, it is also possible to consider how X -p A * X behaves on other spaces. We will mention only three results and leave the proofs as exercises. Exercise I If Xe.,#°° = the bounded martingales, then (A * X) may be unbounded. In Chapter 7, we will see that it does not miss by much. (A * X) C- M_#6 and, furthermore, X-+ (A * X) is a bounded map from M.,#0 to M.,#0. Exercise 2 If X e 21 = the uniformly integrable martingales, then (A * X)
need not be in A-'. (Hint: Recall our discussion of A' log A'.) Exercise 3
If X e X log .7f', then (A * X) e .,lf 1. Is (A * X) e .*-log . ''?
6.7 Janson's Characterization of Jj 1 Having seen an example of an X e .K 1 and a matrix A such that A * X 0 .JY 1, it is natural (if somewhat precocious) to ask which X e AA" have A * X e A 1
6.7
167
Jumon'. Chnncterizetlon of .N'
for all matrices A. If we let J1 denote this collection, then (1) in Scclion 6.6 implies that J1 - .,#' and, as you might guess from the title of this rcclion, J 1 = .111 In fact, more is true : There is a finite set of matrices A 1 , ... , A. such that if A, * X E .7Y 1 for i = 1,
..., m, then X E .,6f 1 (and hence A
To discover which sets of matrices have this property, we start by making the trivial observation that A...... A. cannot have a common eigenvector in Rd, for if we have A;y = ) y for i = 1, ... , m, then we can let r = inf {t: y B, = -1 } and X = y B, A . X is Example 2 of Section 6.6, so X E .tY 1 - _#', but
fori=1, ...,m, (A;*X),=
eAL
0
Janson's (1977) theorem says that this trivial necessary condition is sufficient. (1)
Let A,, ... , Am be matrices that do not have a common eigenvector in Rd. Let A° be the identity matrix. If the transforms A; * X, i = 0, ... , m, are all in
*-', then XE.,#1.
The key to the proof is Janson's generalization of the subharmonicity lemma of Chao and Taibleson (1973), which is in turn a generalization of ideas of
Stein and Weiss (1971), who attribute the idea to Calderon.... Using the notation in (1), we can state this result as: (2)
There is a p0 < 1 (that depends only on the matrices A1, ... , Am) such that if F = (1 + "_0 (X02)"2, then Ff is a local submartingale for all p > p0. Once we prove (2), (1) follows immediately. To see this, observe that if we can pick p < 1 so that G, = FP is a local submartingale, then E(Gt 1P) = E RI +
(Xi) 2
1/2
< E 1 + : IXt ]
(since the L2 norm of (1,X°, ... , Xm)ERm+2 is less than its L1 norm). It follows from our assumption that sup, EG, "P < co, and, since 1/p > 1, it follows from Doob's inequality that
oo > E(sup G,)'IP = E(sup F) > E(sup IXI ), so X E .4f 1. If we keep track of the constants, we get
E(sup I X I) < E(sup G,)' '
<
- (1/Ifl- 1
sup EG, /P ,
<-(1 -p) 1(1 + E
IIAj*XII,).
i=o
To get rid of the 1 on the right-hand side and replace p by po, we apply the
168
6
Hardy Spaces and Related Spaces of Martingales
last result to X/e, multiply both sides of the inequality by e, and let a - 0, p -po to get E(supjXI) :!5; (1 -Po)-1 Y
(3)
IjAj*XILr.
j=o
t
Proof of (2) We start by observing that the result is trivial for p = 2 (F = ) and forp > 2 (since a convex function of a local submartingale 1 + o (Xi)2m is a local submartingale), so we assume p < 2. Let g(x) = (1 + 1x12)2. A little differentiation gives Dig= P(l + Ix12)cp-2'22x; 2
DO = 2.p 2 2(1 +
1x12)cp-a'24xtxj
i:Aj
D«g = .P 2 2(1 + 1 x12)cv-a'24x? + 2(1 +I x12)(p-2'22. 2 Applying Ito's formula, we conclude that
FP-FOP =
f tpXsFs-ZdXs i=o ,Jo
+1
2 i=o ;=o
22 f tPP-
1)4XsXgFs
a
d<X`,X')s
0
+ i JPFr2d<xi>. 1
m
2 i=o
o
The first term on the right-hand side is a local martingale, so to prove the result we need to show that we can pick p < 1 so that the second term plus the third term is 0. If X= f Hsd Bs, then
<X= JAHJ2ds o
and
<X`, X'>' = J(AiJ1AJ1)ds. o
To complete our proof, we need to show: (4)
22 - 1) Yi; 4xtx;(At4P, A;(P) + (1 + 1x12) > 1Aj4,12 > 0 1
(P
for all x e R" 1 and cp e Rd. To do this, we observe that if B,; is the angle between xjA;cp and x;A;cp, then Y(x;A;cp,x;A;(p) _ Y1x;Aicp1lx;A;cpl cosBtj <-Y(IxjAj(p I)2 <1x12YIA1w12.
6.7
169
JanMon's Characterization of.#'
Looking at the first inequality above, we see that there is equality Oiily if ©;j = 0 for all i, j and I x,l = cl A;'pl for all i. Now if x, 9 # 0, it follows from the last equality that Ixol : 0, and since Aocp = cp, the first equality in the last sentence implies that cp is an eigenvector of all the A;, contradicting our assumption, so we must have Y_x.x,(A;'p, Aj'p)
< IXIZ
Ai9IZ
for all nonzero x c Rm 1 and rp e Rd. The value of the left-hand side is not changed by multiplying x or 'p by a positive constant, and it is continuous on the compact set {(x,'p)ERm+1+d: Ixl = 1, Icpl = 1}, so the supremum of the expression over K is a number S < 1. Hence the expression in (4) is > (5)
(1+Ixl')YIAi'I2-2(2-1)4a(IXIzYIAiTl2). When p = 1, the coefficient of the second term is a < 1, so we can pick po < I such that (5) > 0 for all p > po, and this completes the proof.
Readers familiar with Janson's (1977) proof (which is for d-adic martingales) should notice that the outline is the same, but two details are different: (a) Ito's formula replaces the computation Janson does for "small Remark:
jumps" (our jumps have size zero), and then, since we do not have a small/large dichotomy, we need only his first compactness argument. (b) The restriction on the matrices sounds the same, but it is different. In Janson's theorem, the matrices do not have a common eigenvector in Ro = {X: Y; x. = 0}. O) has no eigenvector in R2, it follows from Janson's theorem
Since I 0
I
that we have : (6)
If d = 2 and X and its conjugate martingale k are in . ' 1, then X E If d = 3, then any matrix A has a real eigenvalue and, hence, also an eigenvector
in R" (take the real or imaginary part), so it takes at least two matrices to characterize 1. We leave it to the reader to discover what happens if we take 0
1
0
A, _ (-1
0
0 AZ =
0
0
0
0 0
-1
0
1
0
0
0 0.
If you are very clever, you will discover Riesz transforms. If you get stuck, you can find the connection spelled out in Gundy and Varopoulos (1979).
171)
6
Hardy Space. and Related Spaces of Martingales
6.8 Inequalities for Conjugate Harmonic Functions With each harmonic function u, there is associated a unique conjugate harmonic function it, which has (a) u(0) = 0 and (b) u + iu is analytic in D. In this section,
we will investigate conjugation as an operation on hP, the set of harmonic functions u in D with dp(u) = sup f
<
.
Let Ilullp = (dp(u))"P. If p > 1, this equation defines a norm on hP. Our first result, due to M. Riesz (1927), shows that if p > 1, u -+ u is a bounded linear map from hP to hP. (1)
If p > 1, then there is a constant C (that depends only on p) such that Ilullp < CIIuIIP.
Proof Let U = u(B,), t < T, and let U = u(B,), t < T. If llullP = 1, then Doob's inequality implies that E(U*) < (p/(p - 1))P, and it follows from results in Section 6.4 that Ilullp < El
U*IP < KEI U*IP,
so (1) holds with C = K11P(pl(p - 1)). To be fair, we should observe that (1) is easy to prove analytically-there is a simple argument using Green's theorem that is due to P. Stein (1933) and that gives a much better value for Cp, namely,
Cp=2(Pl(P- 1))lh' 1
2 < p < oo.
Since this proof has some interesting probabilistic aspects, we present it here. If the theorem is true for some 1 < p < oo, then it is also true for the conjugate index q = p/(p - 1) with C. = Cp (exercise for the reader: See Zygmund
(1959), page 255, for the answer), so it suffices to prove the result when 1 < p < 2. In view of the decomposition of functions in hl given in Section 6.5, we can also assume that u(z) > 0 in D. To prove the result, we compute 82
_
OZ7
1
u P(z) W
Pu
= az. (
(Z) azi
=P(P- 1)uP z(Z)(az/z+pup-1(z)8z? Summing and noticing that Au = 0, we get
(a) AuP(z) = p(p -
1)u"-z(z)IVu(z)I2.
Similarly, if we let v = u, then
6.8
Inequalities for Conjugate Harmonic Functions
171
a2 1 flP = a2 u2 + v2)p/2
04
Oz?
_
i
(p22 (u+ v2)(P-z)/z (2u Z + 2llZi
_ .p
2 2. (u2 + v2)(p-4)/2I 2u -Z + 2v
+ P (U2 + vz)(P-z)/z 2 2
(u az,)
z
+ 2u
Z
azZ + 2 az;
(av)z zJ
z \\
+ 2v
Zl. aZ; JJ
Summing and using the relationships Au = 0, Av = 0, Vu Vv = 0, and lVul = I Vv 1, we get
(b) Al./ I P =p(p - 2)(u2 +
v2)(P-4)/z(u2JVul2 + v210u12)
+ p(u2 + v2)(p-2)/2(lVu 12 + 0 + 1Vu12 + 0) =P2 (U 2 + v2)(p-2)/21ou12.
Since p - 2 < 0, it follows from (a) and (b) that we have flp < p
(c) Al
p IAlulp.
This is the key inequality for the proof. From here, completing the proof is easy, and how you do it is a matter of background. A probabilist would reason as follows: Ito's formula says that if h is C2
in D, then fort < r=inf{t:B,0D}, h (B,) - h (Bo) =
JAh(Bs).dBs+iJVh(Bo)ds 0
(here we have written B, as a real two-dimensional Brownian motion). The first term on the right is a martingale for t < r, (since
Eoh(B,) = h(O) + 2E
Ah(Bs)ds. 0
Letting h = up, and then h = If lp, and using (c), gives fo2n f2x
p
I
I
I
o
since 1u(0)l = l f(0)1.
An analyst would use the divergence theorem 0zx
a
JP_(reb0)rdo = f
Acp(z)dA(z),
J IzI
where A' denotes Lebesgue measure(.' This implies that znIf(re`B)lpdO
ar
J
p p -
1
ar J 0
172
6
Hardy Spicea and Related Spaces of Martingales
ivp(Bt)
Figure 6.2
so integrating from 0 to r and recalling that v(0) = 0, we have a second proof of the inequality
J:(re1I0< pp1
f lu(re`Pdo. 0
Since Riesz's theorem is an old and important one, there are many different proofs. Calderon (1950c) has given another proof that you can find in Zygmund (1959), pages 253-255. Pichorides (1972) has given a refinement of Calderon's
proof and has found the best possible constant: CP = tan(n/2p) if 1
f# E H° for p e (0, 1/#). Since p maps D one-to-one onto {z: Re z > 0} = {re'B : r > 0, 0 < n/2}, f,, maps D one-to-one onto the cone IF,, = {re`O : r > 0,
0 < fln/2}. From the last observation (and Levy's theorem), it follows that j,,(Bt), t < 7, is a time change of a Brownian motion C, that starts at 1 and runs until it leaves F''. Drawing a picture (see Figure 6.2) reveals that if up _ Re f,, and vp = Im f,, , then
limIvP(B) I = tanfl imlu,,(BJ)I. tT=
If 1
Itan7E l
IlupiLP,
6.8
173
Inequalities for Conjugate Harmonic Functions
so the optimal constant CP z tan(n/2p) for all 1 < p < oo. Repeating the argument above for ip(z)' shows that CP > cot(ir/2p) for all 1 < p < oU, so Pichorides's constants are the best possible. As the constants in the last remark might suggest, Riesz's theorem is false for p = 1. We have seen the counterexample many times: the half-plane map p(z) = (1 + z)/(1 - z). By computations in Section 6.1, u(z) = Re p(z) Chl, but u(z) = Imp(z) is not, since
n({9:Iu(e'B)I > A}) - 2/n2 as A -+ oc.
The next result, due to Kolmogorov (1925), shows that (up to a constant multiple) this is the worst behavior we can have for ueh1. (2)
There is a constant C such that if ueh1, then n({e: lu(e`a)I >_ 2}) < Ca.-11IuII1.
Remark: In the jargon, this is called a weak type (1, 1) inequality. We will give Davis's (1974) proof, because it has the advantage of identifying the best
constant C = 1.347... . Proof Let ZZ = Xt + iY be a complex Brownian motion, and let a = inf{t : I Y I = 11. Since t - I X, I is a submartingale and the stopping time a depends
only upon the Y component, it is easy to see that if x and - 1 < y < 1 are real numbers, (i) EoIXaI >EIYIXal (ii) Es+i,IXal > EirI X1I.
The next step is to show :
(iii) Let x be a real number and fi a stopping time for Z. Then EXIX API > EOIX.IPX(, ? a).
Proof By (ii) and the triangle inequality, EOIXXI < E.IXal < EXIX..pl + EXI Xa - X.A$I.
By the strong Markov property, Exl X8 -
Xa
-
and we have, using (i),
a> (the centering makes it look like starting at 0, but I Y,,, until a). Combining the results above, we have EOIXQI <ExIXZAPI +EoIX2IPP(«> fi),
proving (iii).
> 0, so there is less time
174
6
Hardy Spaces and Related Spaces of Martlnaalee
We are now ready to prove Kolmogorov's inequality. Let f = u + iu. By Levy's theorem, if we let a(t) = f O I f'(B5)IZ ds fort < T, then Z,(,) = f(B,), t < T,
defines a Brownian motion run for an amount of time y = Q(T). If we write Z, = X, + iY and let T, = inf{t : IB,I > r}, then IX",(tr)I = iu(BtAT,)I is a bounded submartingale, so EI X.Aa(t,)I <_ EI XC(tr)I = EIU(Bt,)I
Ilulll.
As r T 1, T, T T and Q(T,) T Q(T) = y, so using Fatou's lemma gives
(iv) EIXanvl < IIulIi.
We have finally assembled all the ingredients to complete the proof: n({O: Iu(e`a)I > 1}) = P(I YYI >- 1) P(YY 1) = P(a < Y),
so it follows from (iii) and (iv) that the above <- (EopX.l)-'EolX2Arl
< (EofXal)-'Ilulll. Substituting u/), for u, we have n({e: lu(e`B)I z A)):5
(EoPXaI)-'llulil/2.
It is clear from the argument above that the inequality is sharp. If we let g(z) _ (2/7r) log p(z), then g maps D one-to-one onto the strip {z : -1 < Im z < 11, so
1 =n({O:1u(eie)I > 1})=P(a<-y) = (EojXal)-'EoiXal = (EOIXaI)-'llulll.
To compute the constant, observe that by (4) in Section 5.1, the probability density of X, is ex'/2/(I + exn), so
_
zi
(-
8
n=0 °°
z
n=0
000
(-1)n(2n +
1)-2
If you recognize that nz/8 = En '=O (2n + 1)-2, then you can write the constant as (1-3-z+5-z-7-z+...)/(1+3-z+5-z+7-z+...).
Although the inequalities (i)-(iv) are important, the two main ideas in the proof of (2) are the observation that : (a) If we time change f(B), t < T, to obtain a Brownian motion Z, = X, + iY, t < y, then
6.8
175
Inequrlltlai for Conjugate H.rmonlc Functions
n({H: u(e")I >_ 1}) = P(I YYI >_ 1) 5 P(Y* > 1),
and (b) forueh',P(Y*>_ 1)/Ilulllis largest when y=a=inf{t:IY,I= I;. With this philosophy as a guide, the reader should have no problem proving +i onesided version of Kolmogorov's inequality. Exercise l
If u c h' and u > 0, then
({O: u(e`") > }) < 2IIu111
7r(l
+y2),
JA
and this constant is the best possible.
Hint: The function that shows that this inequality is sharp is
f(z) = it + (1 -
z)'
i+
which maps D one-to-one onto {x + iy : x > O, y < 2} and has f(0) = 1. In this case, the reflection principle shows that
n({O:u(e`")>2})=2 f Z
dy + y2). n(1
Applying (2) to u(rz) shows that if 1lull1 = 1, then 7C({O :
I u(re")I > A}) < CA-',
soifp<1 2n 0
u(re`")IPdn(O)= f"Op
< [P2P1dA+
>.l})dA
Jp1 C` d= 1
P,
and we have proved another inequality due to Kolmogorov: (3)
If u e h1 and p < 1, there is a constant C (that depends only on p) such that dd(u) < CIIull f.
The argument above proves (3), but since it is not very interesting and does not give a very good value of the constant, we will give two more proofs of (3). The first is a purely probabilistic one, and because of this, it gives a
crude value for the constant. The second is an analytic translation of the probabilistic argument and gives the best constant for positive u. Proof 2 Again without loss of generality, we can suppose that u > 0 in D and IIuhl1 = 1. In this case, u(0) = 1 and u(0) = 0, so (u(B,), u(B,)), t < 'r, is a time change of a two-dimensional Brownian motion starting from (1,0) and running for an amount of time y. Since u >- 0, y must be smaller than 71, the time it takes a Brownian motion starting at 1 to hit 0, and it follows that
176
n
1lardy Spacra and Related Spacca of M.rtlnialra
dp(u) < /;(U*)p < CE 200< -CETv1Z.
In Chapter 1, we found that the probability density of T1 is (tit) - 3/2 t- 3/2 e- 1/2t ,
l
so ifp < 1, ETf2 =
(2n)-3/2
f aD
t(p- 3)/2 e - 1/21 dt.
0
To evaluate the integral, let u = 1/2t to obtain (2ir)-3/22(1-p)12
r u-(1+p)l2e-"du J0
and observe that the value of the integral is I'((1 - p)/2) if p < 1. The reader should observe that if u(x) = Re p(x), then y = T1. This suggests that for nonnegative functions Ilullp/Ilullp should be largest for the Poisson mapping. The proof above cannot be used to prove this fact, since we have used the clumsy estimates dp(u) < EI U*Ip < CEc/2. This inaccuracy can be avoided if we abandon the correspondence and translate the proof into analytical terms. Proof 3 The idea for this proof is due to Littlewood (1926). Again without loss of generality, we can suppose that u > 0 in D and II U 111 = 1. A function f analytic in D is said to be subordinate to g if there is an analytic function co with Ico(z)I < Izl such thatf(z) = g(w(z)).
One reason we are interested in this concept is that
(i) If u > 0 and u(0) = 1, thenf = u + iu is subordinate to p(z) _ (1 + z)/(1 - z).
Proof Let w(z) = p-1(f(z)); co is analytic and maps D into D and 0 into p-1(1) = 0. Let a(z) = w(z)lz, z 0, and a(z) = w'(0) at z = 0; o is analytic in D and has I u(z) I < 1 on 8D, so it follows from the maximum principle that Io(z)I < 1 in D, that is, Iw(z)I 5 Izl. Another reason for our interest in subordination is (ii) If f is subordinate to g, then f02
I f(re`a)I 'd7r(0) < I
I g(re`a)I pd7r(0)
for all r < 1.
0
Proof Let B, be a complex Brownian motion and 'rr = inf{t : IBI > r}. By hypothesis, EI J (Br)I p = El g(w(B1r))I p
6.8
Inequalities for Conjugate Harmonic Functions
177
If we subject w(B,), t < z to Levy's time change, we get a Browniun mot ion B,' run for an amount of time y,. Since Iw(z)I < Izi, it follows that y, r; inf{t : IB;I > r}. Combining this with the fact that Ig(B',Tr)IPisa submurtingulc gives
EIg((o(B,,))IP=EIg(BY.)IP<EIg(Bt)IP and proves (ii). Combining (i) and (ii) proves (3) with CP = dP(q), where q(z) = Imp (z).
To compute the value of the constant, write p(z) = R(z)e"(z), where R(z) = I p (z) I and cb(z) E (- n/2, 7r/2). Since there is no confusion about which root to take, F(z) = R(z)Pe'P(z) is analytic in D. By the mean-value theorem,
1 = Re F(0) = f
Re F(reie) d7r(O)
,J n n
= f
R(re'B)Pcos(p(D(re`B))d7r(O). n
As r
1, cb(re'B) -+ sgn(9)n/2 and R(re'B) -> Iq(e's)I. Since gehP and Icos(p(D(reie))I < 1, it follows from the dominated convergence theorem that
1= f
cosP2 lim f rT1
,1
Iq(reie)IPdm(0), n
so dd(q) = sec(pn/2).
Remark: If at the beginning of the computation of the constant we take an arbitrary f with u = Ref > 0 and use the inequality I ((re') I < n/2, we get a purely analytical proof with the constant given above. This argument is due to Hardy (1928). Proof 3 shows that (3a)
If u > 0, dd(u) < sec (12) IIuIIl
(3b)
If uEh1, dP(u) < 2 sec 12 I IIuIIl
.
The Poisson kernel shows that (3a) is a sharp result. Since (3b) is obtained from (3a) by using the triangle inequality, we should expect that (3b) is not sharp, and indeed it is not. Burgess Davis (1976) solved the problem of finding
the optimal constant in inequality (3b). He showed that the smallest value for C, is IIull,,, where u = (k1 + k_1)/2. In this case, the corresponding analytic function is g(z)
1 (I +z 1-z =2 1-z+1+z
2z (1-zZ),
178
6
Hardy Spaces and Related Spaces of Mortinplea
which maps D one-to-one onto S = C - {x + iy : x = 0, IyI >- 11. To prove his theorem, Davis uses Levy's theorem to reduce the result to an optimal stopping problem for Brownian motion, which is solved by considering related discrete time problems. Since the argument is rather lengthy and the improvement on (3b) is rather slight, the reader is referred to Davis's paper for the details or
Davis (1979b) for a sketch of the proof. A. Baernstein (1978) has given a purely analytical proof of this result. Up to this point, we have only discussed u e h° for p > 1. For 1 < p < oo, the class hP was preserved under conjugation. For p = 1, this was false, but it
was almost true: uehP for all p < 1. When p < 1, things fall apart-there is an analytic function f = u + iv such that uehP for all p < 1, and yet foN (hence v 0 hP for any p > 0). The example is a randomly chosen function f(Z, (0) =
(4)
n=1
Sn(w)g(ZZn)
where g(z) = z/(1 - z2) and fin, n > 1, are independent random variables with
-1)= 1/2 (for analysts, let S = [0, 1] and n((o) = sgn(cos(2nw)), the nth Rademacher function). I claim that (4a)
For every w, Ref (. , co) a hP for all p < 1 and
(4b)
With probability 1, Yo = (0: lim,fl f(re`B, (o) exists} has Lebesgue measure 0.
The proof of (4a) is a straightforward, but somewhat lengthy, calculation and is therefore left to the reader (see Duren (1970), page 66). We will proceed, then, with the more interesting claim (4b), which is implied by the following result. (5)
Let gn(z), n > 1, be complex-valued and continuous in IzI < 1 except at a finite number of points z with IzI = 1.
If (i) Y' O, I g (z) I < oo and for all r < 1 the convergence is uniform on D, {z : IzI < r}, and (ii) for each N, as r T 1 we have oo
gn(re`B)I2. oo
uniformly in 0,
n=N
then with probability 1, f(z, (o) = Y',
(w)gn(z) has a radial limit almost
nowhere.
Proof Let E = {(0, (o) : lim,.1 f(re`B, co) exists}. It suffices to show that for almost every 0 the section Ee = {w : (0, co) e E} has probability 0, for then the desired conclusion follows from Fubini's theorem.
6.8
179
Inequalities for Conjugate Harmonic Functions
Suppose 0 is such that eie is not a discontinuity point of any of the an, i,ntl P(EB) = a > 0. We are going to construct a decreasing sequence of events An. n > 0, with AO = EB and do this in such a way that limn-,. P(A,) Z a/2. I'or
k,n> 1, we let B,(k) = {coeAn_1 : I f(re`B,w)I < k for all r < 1}.
B,(k) T An_1 as k T oo, so if we have already constructed An_1 with P(A,,_1) z
(1/2 + 1/(n + 1))a, then we can pick kn so that P(B,(kn)) > (1/2 + 1/(n + 2))a and let An = B,(k,J. Let B =n,, 1 An. By construction, P(B) > a/2 and, for all n, z
00
i 9m(re') m(w)
1B < k2P(B).
m=n
The functions bi4, 0 < j < k < oo, are an (incomplete) orthonormal set in L2(S2), so
Y
(E1B
m)2 _< II1B 2.
n
0<m
It follows that we can pick M large enough so that
Y
(E1Bnm)2 <
(PB)2MSm
Since B c AM, we have (since m = 1) kMP(B) > E I
Y gm(ret
Z
m=M
1B/
g g Y gm(re` )gn(re`
ao
= EI
\m=M
gm(re`B)21B + 2
MSm
Cauchy-Schwarz implies that the last expression above 2
>
M<m
<2 m=M
Igm(re` lI2 gn(reiO)I2) 1/2
I gm(re`
Y
Y (E1BSm
(M<m5n
II2)1/2(Ign(rei0)I2)
n)2) 1/2
/2 P(B)
1
n=M
Combining the results above, we see that
kMP(B) ?
-2), m=M
contradicti ng (ii) and proving (5). The next exercise recaptures the main aspects of the last proof in a simpler setting.
Exercise 2 If W 1 Ia.I2 = oo, and SN = N 1 with probability 0.
then limN-,, SN exists
I NO
6
Hardy Spaces and Related Spaces of Martingalee
Remark: Probabilists will recognize this exercise as a special case of the Kolmogorov three-series theorem (see Chung (1974), pages 118-119).
Proof Let C > 0 and B = {w : supN I SN((O) I < C }. If (o e B, I S,,(w) - Sm(w) 15 2C for all m, n, and K
4C2P(B) >- E ( k=m+1
2
ak 4
1B )
n
= P(B)
akI2 + 2
YY
m+15j
k=m+1
Again, SJSk, j < k, is an orthonormal system in L2(Q), so we can pick M large enough such that P(B) z,
YK M+Yj
and can apply Cauchy-Schwarz to show that 4C2P(B) > (k= Y_
I akI z)
P(B) (1
- 2),
Mn+1
which is a contradiction, since we have supposed that Y_ IakI2 = oo. There are many other results for conjugate functions. At this point, we have
not covered all the inequalities on the first three pages of Chapter VII of Zygmund (1959) ! Another inequality that can be proved, using results from Section 6.2, is the following. Exercise 3 Zygmund's Inequality. If gpeL1(8D) and we let p denote the conjugate function on 8D (i.e., the boundary limits of the conjugate of Y(p), then there is a constant C such that zn
f2,,
I:edo
(e`e)log(p(e1e)dB.
6.9 Conjugate Functions of Indicators and Singular Measures In this section, we will investigate the conjugate functions of u = 91A, A C 8D,
and u = /j where u is a measure with µ(8D) < oo, which is singular w.r.t. surface measure. In each case, it turns out that (*) qp(O) = limu(re`B) rt l
exists for a.e. 0
and the value of rz(O : qp(O) < y) denotes only on n(A) or µ(8D), respectively.
6.9
Conjugate Functions of Indicators and Singular Measures
181
The result for 1,, was first observed by Stein and Weiss (1959), who proved the result by computing the distribution when A = Ui (ai, bi) is a disjoint union of intervals (see pages 273-276 of their paper) and observing that the distribution depended only on Yi Iai - bit. The result becomes transparent if we look
at f = u + iu through the eyes of a Brownian motion B, starting at 0. Doing this, we observe: (a) If r = inf {t : B, 0 D}, then f(B,), t < r, is a time change of a Brownian motion C, starting at f(0) = n(A) and run for an amount of time
y= J
I .f'(B,,)IZds.
(b) If 0< ir(A) < 1, then 0< u < 1 in D and u(B,) a (0, 1) for all t < r. This implies that y < T = inf It : Re C, 0 (0, 1) }. (c) Since 1A E L1(8D, n), it follows from results in Section 6.5 that
lim Re C, = lim u(B,) = 1A(BJ a.s., t1= tty so we have y T. (d) Since T < oo, it follows that lim u(B,) = lim Im C, = Im CT, tfv
tfT
and using the equivalent of Brownian and nontangential convergence in d = 2 proved in Section 4.3, we conclude that (*) holds and that (1)
7r (0 : OP (0) > y) = PP(A) (Im CT > y).
To compute the value of the right-hand side, we observe that z --, exp(in(x - 1/2)) maps the strip {0 < Rez < 1} one-to-one onto the half space {Rez > 0} (see Figure 6.3) and sends 1
-.i
1 + i t -> ie-z' 2
a -- exp(in(a - 1/2)),
so if we let a = n(A), b = n(a - 1/2), and c = -2n/2, then it follows from results in Section 1.9 that _fec
CT > y) =
1
cos b
n (cos b)2 + (y - sin b)2
dy,
and changing variables x cos b = y - sin b shows the last integral
=1
((tan-, e`
sin cos b b)
- tan'
(-e-
- e cos b
in b)l
Remark: To be fair to the analysts, we should say that the proof above is essentially due to Calderon (1966) and was later rediscovered by Davis (1973b),
who wrote the proof in probabilistic language. We leave it to the reader to
1$2
6
Hardy Spaces and Related Spaces of Martlnf{ales
a 0
1
exp(iir(z - 1))
-le-ami:
exp(i17(a -
Figure 6.3
show that (1) is consistent with the result given by Stein and Weiss (1959) (see (4.3) on page 273) that sinh Y + i sin
/
expl tl{O:IcP(e)I >Y}I) = 2
2
A2
sinh 2 - i sin 12 I
At this point, the reader can probably guess how we are going to prove the result for u = 9aµ when u is singular. Let f = u + iu and observe that
(a) f(B,), t < t, is a time change of a Brownian motion Cr starting at f(O) = µ(8D) and running for an amount of time
y= Jf(B)I2ds. (b) If 0 < µ(8D) < oc, then 0 < u < oo in D and u(B) a (0, oo) for all t < T. oo)}. This implies that y <_ T' = inf{t : Re (c) Since 9au E h', it follows from results in Section 6.5 that
6.9 Co*gate Functions of Indicators and Singular Measure.
lim Re Ct = lim u(B1) = 0
ttr
183
a.s.,
ttt
and so we have y = T'. (d) Since T' < co, it follows that lim 4(B) = lim Im Ct = Im C(T'), tt=
tty
and using the equivalence of Brownian and nontangential convergence in d = 2 proved in Section 4.3, we conclude that (*) holds and that (2)
7r (0 : p(0) > y) = PP(aD) (Im C(T') > Y)
This time it is trivial to compute the right-hand side. If we let a = µ(8D), then
Pa(Im C(T') > y) = l JJ
a 1t
a2
x2
dx.
The proof above is due to B. Davis (1973).
7 H1 and BMO, Jf 1 and Rd(
7.1
The Duality Theorem for # 1 In Chapter 6, we saw that ifp > 1, X -. X.,, maps .,!!P one-to-one onto in such a way that the., #P norm of X is equivalent to the LP norm of X.. From this observation, it follows immediately that every continuous linear functional qP on MP can be written as cp(X) = E(X,, Y), where YEL9, q = p/(p - 1). When p < 1, the equivalence of MP and LP breaks down and the reasoning above fails to identify the dual space (,#P)*. In this section, we will consider the problem of describing (.,lfl)*. The first step in the solution is to introduce a decomposition due to Bernard and Maisoneuve (1977), which expresses a general XE.,!!1 as a .,#1. As will be the case many times below, (a) sum of very simple elements of the probabilistic definition was developed after and imitates the definition in-
vented by analysts (see Coifman (1974), Latter (1977), Coifman and Weiss (1977)) and, more embarrassingly, (b) we will often assume (e.g., in the proof of (4)) that our martingales start at 0 at time 0 and forget to mention this. (1)
A martingale A E.,kl is said to be an atom if there is a stopping time T such that
(i) A,=Oif t
If A is an atom, IIAII, = EA* < 1.
It follows from (2) and the triangle inequality that (3)
If A" is a sequence of atoms and c" is a sequence of numbers with Y-1c.1 < oo, then X= >c"A"E.,ff' and 11X111 < >1c"1-Clearly, di1. The (3) allows us to construct many examples of martingales in next result shows that every XE if 1 can be built up in this way and furthermore
thatY-lc"1
7.1 no Dudity 'Theorem for .,N'
(4)
185
For all X e #', there is a sequence of atoms A", n e Z, and a sequence of constants c", n e Z, with y_n Ic"I < 6IIXII1 such that as N, oo, N
I c"A" - X in fl'. n= -N
Remark: I think the proof of this result is beautiful. It is a trivial computation, but it is also an ingenious idea. To convince yourself of the latter, you should put the book down for a few minutes and try to construct your own decomposition.
Proof One answer to the problem is: for each n e Z, let
T"=inf{t:IXI>2"} and let An _ (X(t A
Tn+1)
- X(t A T"))/c".
The definition is arranged so that >c,,A, is a telescoping series, so we have N
X, - Y c"Ai = X(t) - X(t A TN+1) + X(t A T-N). n= -N
The last term on the right is <2-N and, hence, in as N - oo. To estimate ' norm of Y = X(t) - X(t A TN+1), we observe that Y = 0 on the {X* < 2N+1 } and Y* 2X*, so by the dominated convergence theorem, EY* < E(2X * ; X* > 2N+') -. 0 as n -> oo. Up to this point, the values of the cn's and the precise form of the stopping times have not entered into the proof. We must now choose the c"'s to make the A"'s atoms. To do this, we observe that
X(t A Tn+l) - X(t A T")I < I2n+1 _ (-2")I = 3 ' 2", so if we want IAt I < P(Tn < co)-', we must pick cn = 3 - 2"P(Tn < oo).
Having done this, we find that
EIcnl=Y3'2"P(X*>2") n 2"
2n
P(X* > 2")dy
= n
:!5; 6
J
f2"-
P(X*>Y)dy=611XII1.
0
Remark 1: It is important to observe that we use 2", n e Z, and not just n > 0. We do this (that is, use n < 0) and use a sequence that grows geometrically, so
that the picture remains the same, if we multiply by 2'. The last feature is
1N6
7
H' wd BMO..t' snd 44&J
crucial if we are going to prove an estimate like (4), which is unaffected if the quantities under consideration are multiplied by a constant. Remark 2: While it is important to let T" = inf{t : I X,I > a"} and to pick c" = (a + 1)a"P(T" < oo), the actual choice of a is not crucial. If we repeat the last computation above in this generality, we find that Y_ICI<
(a+l)
(1 - a-')
EX*.
The constant is optimized by taking a = 1 + -,,,[2-, and for this value of a we get a constant = 3 + 2.,/2- = 5.828... , which hardly seems worth the effort.
With the decomposition in (4) established, it is "easy" to find the dual of .A". A linear functional 'p will be continuous if and only if (5)
sup {IT (X)I : Xis an atom} < oc.
To be precise, a linear functional defined on the linear span of the atoms will have a continuous extension to df' if and only if (5) holds. As in the case of the decomposition, it is hard to guess the answer (the reader is again invited to try), but if somebody tells you the answer and shows you which atoms to use as test functions, it is not hard to fill in the details.
We say that Y has bounded mean oscillation (and write Ye .AV) if YE.,lf2 and there is a constant c such that for all stopping times T, (6)
El Y, - YTI < cP(T < oo).
The infimum of the set of constants for which (6) holds is called the M.,616 norm of Y and is denoted as 11 YII* This definition may not look very natural now, but it will by the end of the next proof. (7)
Let sd be the set of atoms. For all Ye _#', III YII* < sup(IE(X, YJ : Xed) < II YII*. Proof We will first prove the inequality on the right. If X e sad and T is a stopping time for which (1) holds, then
E(X, YT) = EE(XJ E(YTE(XXI FT)) = E(YTXT) = 0,
since XT = 0. From this it follows that EX. Y.1 = I EX.(Y. - YT) 1
< E(X*I Y. - YTI) < P(T < oo)-'El Y. - YTI <- 11
YII*.
7.1
The Duality Theorem for.,N'
187
To prove the other inequality, let T be an arbitrary stopping time, let Zoo = sgn(Y, - YT), and let Z, = E(Z,,,,I.) be the martingale generated by this random variable. Since IZ,I < 1, X, = (Z, - ZTA,)/2P(T < oo) is an atom. I'rom the definition of Z,,,,, it follows that
EIY. - YTI =E(Zc(YY- YT))=EZ,Y.-EZ1YT. The first computation in the proof shows that E(Zao YT) = E(ZT YT) = E(ZT Y.),
so we have
EIY. - YTI
=E((ZW-ZT)YT),
and it follows that
E Y. - YT = E ((Z. - ZT) 2P(T < oo)
1\2P(T < oo) Y-)
= E(X Yom).
Taking the supremum over all stopping times now gives the desired result. With (4) and (7) established, it is now routine to conclude that R.,#&. To prove (41)* c RMO, we observe :
(a) .#2 c ill, with IIXII2 =
(EIX*I2)I12
> EI X*I = IIXIII, so if co is a continuous linear functional on #', cp induces a continuous linear functional
on M2 with IIPII2 = sup{Iw(X)I :IIX112 < 1} < 119111(b) from the duality theorem for ..1f2, it follows that there is a YE .% 2 such that
cp(X) = EXm Y. for all X E JI'2. Since d c M', it follows from (7) that YE'4A' V.
(c) from the atomic decomposition, it follows that .,6l2 is dense in M, so the correspondence cp -+ Y defined in (b) is one-to-one.
To prove that (A'l)* = R.,!!U, we now have to prove that all the linear functionals given above are continuous. This follows from the next result (Fefferman's inequality). (8)
If X, YE.,#2, then I E(Xa YaD)I
611XII I
II Y.
Proof From (4), it follows that X can be written as 1" c"A", where All, n E Z, is a sequence of atoms and N
Y_ c"A< < X* + 1
for all N, t.
.=-N
Since X E A'2, we have X* E L2, and it follows from the Cauchy-Schwarz inequality and the dominated convergence theorem that
EX. Y. = Y c"E(A" Y.). n
INN
7
H' and BMU, .41 nd w..4(J
Using the triangle inequality now with the results of (7) and (4) gives the desired conclusion : IEX. Y. I
EIcnI IE(A"YY)I n
I IcnI II YII* < 611XII1II
YII*.
n
Remark: What we have shown above is that if (p e (.,#1)*, then there is a Ye-V.,#O such that cp(X) = E(X, Y.) for all Xe.A'2. In Section 7.2, when we give the "classical" proof of the duality result, we improve this conclusion slightly by showing that qp(X) = E<X, Y>W for all Xe.441. (9)
If there is a constant c such that for all stopping times T
(*) EI Y. - YTI < cP(T < oo), then it follows that we have (**) E(I Y. - YTII 30"T) < c
a.s.
for all stopping times.
Proof Applying (*) to the stopping time
T' =
(T
ifE(IY(-
oo
otherwise,
we see that if P(T' < oc) > 0, then cP(T' < oc) Z E(I YO -YT E(E(I Y. - YTII JET) 1(T' x,))
> cP(T' < oo), a contradiction, so P(T' < oc) = 0.
7.2 A Second Proof of (,&')* = R,& O In this section, we will give a second proof of the duality theorem for -#' following Meyer (1976). This approach starts with a somewhat different definition of -4,V9.
Let Xe.,#2 with X0 = 0. We say that Xe-"02 if there is a constant c such that, for all stopping times T, (1)
E(XQ - XT)2 < c2P(T < oo).
The infimum of the constants with this property is called the _q,#02 norm of X and is denoted by <<X >> *.
7.2
A Secood Proof of (. 4')* - 4Y4N
189
Remark 1: From Jensen's inequality for conditional expectations, it follows
that if Xe. .,lf02, then Xe-V..lf0 and <<X>> Z IIXII*. An inequality in the other direction, <<X>> 5 CIIXII*, is also true and is a consequence oI' the two proofs of the duality theorem. We will also give a direct proof of the second inequality in Section 7.6. XT Remark 2: Since E((X, the definition of -4.,k02 can be written as
(1)
IJIT)
= E(<X )m -
E(<X> - <X )r) : c2P(T < oo) or, in view of (9) in Section 7.1, as
E(<X> -
c2
a.s.
The first step in our second proof of the duality theorem is the same as in the first proof. We show that every continuous linear functional on .,K1 comes from a Ye 9.,0 02. (2)
If 9 is a continuous linear functional on A1, then there is a YE-V.,#02 such that for all Xe.df2 cp (X) = EX. Y..
Proof It suffices to prove the result when <<9>> 1 = sup { I (p(X) I : <<X>> 1 5 1 } = 1. Jensen's inequality implies that «X))2 = (E<X)W)112 z E(<X>10012) _ <<X>>,, so tp induces a continuous linear functional on .,#2, and it follows as in Section 7.1 that there is a Ye.112 such that 9(X) = EX, Y. for all Xe A'2.
To show that Ye"02, let T be a stopping time and let X, = Y, - YT,,,. X is the stochastic integral J Y where J = 1[T,.,), so using our formula for the covariance of two stochastic integrals ((3) of Section 2.6), we have <X, Y> = JJd
<X,X)r =
J,2d3. J0
Since J32 = J, it follows that
<X,X>.=<X,Y)m=
it follows from the inequalities of Doob, Cauchy-Schwarz, and KunitaWatanabe that
EZ-E<X,Y).=EX.Y, . On the other hand, we have EX. Y. = (p (X) <- << X ))1 = E(Z 1/2),
191)
7
H' mnd BMU, .N' rnd M,
and since Z = 0 on IT= oc }, it follows from the Cauchy-Schwarz inequality applied to Z 1(T< .) that E(Z112) < (EZ)112P(T < 00)1/2.
Combining the last three results proves that EZ < (EZ)112P(T < 00)112,
that is,
EZ = E(
If XE., (1 and Ye .,lf02, then
J'ld<X,Y> l
E
Proof By stopping, it suffices to prove the result when X, Y, <X>, and
EJ
Since <X>, has bounded variation, ordinary (Riemann-Stieljes) integration gives <X>-112
d <X )t = 2<X> a1o/2,
J0 so
E
f
<X>t 1/2d<X>, = 2<<X>>1
0
To estimate the second integral, we fix co and integrate by parts to obtain E J oo <X>112 d
=E
r
J0
2
-
fOD
d<X>i12
0
-
At this point, it is very easy to complete the proof if we leave out one detail. I claim that since <X> is adapted to F, the last expression
7.2
A Second Proof of (.t') - :M.A'
= E f " E(< Y>.
191
-
If you accept this, then there is nothing left to show, for definition (I") of 5U. WC) 2
implies that the above is < KY>)*E(<X>1/2) 00 = <<X»1<
Combining the inequalities above shows that I d<X, Y>sI < (2<<X>>1)1J2(<<X»1<
EJ 0
which is the desired inequality. To complete the proof of (3), it remains only to justify the equality claimed above. To do this, we will prove a general result: (4)
If Z is a bounded random variable and A is a bounded increasing process adapted to IF with AO = 0, then E fo'O E(Z
I S)
Some care is needed in defining the integrand, since A may be singular and, for each t, E(Z I.;) is defined only up to a null set. In our situation, there is no problem. The Brownian filtration admits only continuous martingales, so we take versions of E(Z I.) that are continuous in t for each w. In the language of the general theory, we are taking the optional Technical Remark:
projection of the process YY = Z (which is constant in time), but in our situation
we do not need this notion, since there is only one reasonable way to define E(Z I.) for all t simultaneously.
Proof Suppose without loss of generality that Z >_ 0.
EZA = E
A ko
\n / -A Ck
n 1/
=Ek>E(Z(A\n/-ACkn = E kj (A 1 =1
/ ki)
1/)I,
- A (k n
n
1) E(Z
I Ak/n)
-> E f 0 E(Z I.,) dA1 0
as n - oo, by the dominated convergence theorem.
192
7
H' aad BMO, .N' rnd 91.+YPV
Taking Z =
J
- J F
0
0
=J
E(
)d<X)''2
0
=f
-f
0
-
)d<X>t
which completes both the proof of (3) and the proof of the duality theorem. As we mentioned in Section 7.1, one advantage of the new proof is that it gives
a formula for the linear functional that is valid on the whole space and not just on a dense subset.
7.3 Equivalence of BMO to a Subspace of L# Our next goal is to prove the duality theorem (H1)* = BMO. One-third of the work for the proof of this result was done in Sections 6.4 and 6.5, when we showed:
(a) f - Re f(B,), t < r, maps Ho one-to-one into .,ktl and, furthermore, the //T' H' norm off is equivalent to the Xe..,R#', norm of its image (b) if we let X, = Ref(BfAT), then so if we let M: f-+ Ref(BEAT), then the results in (a) hold when the r is erased.
It follows from (b) that M(H0) is a closed subspace of .#1 and that all continuous linear functionals on Ho have the form A(Mf), where A e (.,#1)*. Since (.,ff 1)* = _4.,#0 (the second third of the work), it then remains to identify M(H, )* c -4.,#(9 and to show that we can map BMO one-to-one onto M(Ho )*
in such a way that the BMO norm of a function is equivalent to the .4,f(9 norm of its image. The answer to the first question can be found by very naive reasoning :
M(HH) = {XE.,ffl : X0 = 0, X, = h(B,), t <,r, and X is constant for t > r}, a space we call lf,; , so if L.,L(O,, is defined in the obvious way, we should have
When we prove that (H1)* = BMO in Section 7.4, we will show that this reasoning is correct. To prepare for that result, this section is devoted to determining which harmonic functions h give rise to martingales in (.,ff,; )* =
R.,1f0,, (the last third of the work). The first baby step in doing this is to observe : (1)
If we let ('
Yf= J ke(z).f(e`B) dn(0), then h = Yip where 9 e LZ(8D, n).
7.3
Equivalence of BMO to a Subspace of A WO
193
Proof c .,llti , so the result follows from (2) in Section 7.3. The last result is not much but, in view of the remarks at the end of Scct ion 6.5, it allows us to think of h as being a function (p e LZ (8D, n) and to conclude that t < T
XJh(Bt) t
- gq(Bz)
t > T
is a continuous martingale e.,ll2. A more substantive conclusion results if we use the definition of M.,KO: for all stopping times T, E((Xao - XT)ZI FT) S C2.
So it follows from the strong Markov property that
E((XX -
`YT)2I_f
T) = W(BTnz)
where w(z) = 0 on OD and w(z) = Jko(z)(co(ei) - YTV(z))'21r(O)
when z e D. So in terms of w, the condition for x to be in M.,lfO is (2)
For all z e D, w (z) < c2.
The aim of this section is to show that the last definition is equivalent to the following notion in analysis. (3)
T is in BMO if there is a constant c such that for all intervals I, 19
-WIIZI
f
I
where
=
dO (P
(P,
I
is the average value of p on I. The smallest positive constant with this property is denoted by Conditions (2) and (3) have a very similar form. To emphasize this we will introduce some notation. Let II
.le(re`l') =
jl/(l - r) if Iii - OI < (1 - r)n 0
otherwise
Ap (z) = JJo(z)(e)d7t(O) x'(z) = JJ9(z)(4(ei9) - y (z))2drt(O).
194
7
..e :ir.4u
H' .ed 3MO,
If we recall that dn(0) = d0/2n and observe that as z ranges over D, the supports of the maps 0 -' je(z) run through all possible intervals I, then we see that (3) says simply that w(z) 5 c2 for all z e D. The first step in understanding the relationship between (2) and (3) is to
understand the relationship between ke(z) and je(z). To begin, we will look at the asymptotic behavior of k9(r) as r - 1. (4)
ke(r) =
I lr-
erelz
1 - r2 (r - cos 0)Z + (sin 0)2
1 - r2 1 + r2 - 2r cos O
(1+r)(1-r) (1 -r)2+2r(1 -cos0) 1+r (1+2r(1-cos0)1-1 1-r (1-r)2 J ' so if we let 0 = y(1 - r), then
(1 - r)kl(l-,)(r) -- 2(1 +
y2)-1
2n times the density of a Cauchy distribution with parameter 1. The last result should be no surprise. If we approach the boundary of the disk and rescale the picture so that we are at a distance 1 from the boundary, then the boundary will approach a straight line, so the rescaled exit distribution will approach the Cauchy distribution and the factor of 2n arises since on OD we are looking at the density w.r.t. A = d0/2n.
The computations above show that the natural place for viewing ke(r) is at 0 = y(1 - r). Since this is the width of the support of je(r), the strict positivity of the Cauchy density leads easily to : (5)
There is a constant A such that, for all z and 0, je(z) 5 Ake(z).
Proof From (4) we see that je(z) < C,ke(z) for all Izl < r, so it suffices to consider what happens when r - 1. From (4), it follows that
ke r
O
_ l + r (1 + 2r(1 - cos 0)1 r)2
1-r
J
Now 1 - cos 0 = e(0), where e(0) _ 02/2 as 0-0, so Is(0)15 CO' for Oe [0, 2n] and it follows that if r < 1 and 101 < (1 - r) 7r, then
1+2r(1-r)Z0)51+(1 2r)2C(1-r)Zn2, proving (5).
7.3
Equivalence of BMO to a Subspace of :,Y.,NtV
195
With (5) established, it is easy to prove: (6)
For all zED, w(z) < Aw(z).
Proof By (5), Aw(z) ?
- 9a(p(z))2j9(z) dir(e)
J((e1°)
- Ap(z))2je(z)dn(O),
since a - $ ((p(e`) - a)2j9(z) drz(O) is minimized at a = 9atp. In more familiar terms, the mean p = EX minimizes E(X - a)2, since
E(X-a)2=E(X-p+p-a)2=E(X-p)2+(p-a2). Having proved (6), we turn our attention to the other comparison. From the proof of (6), it is clear that it would be enough to show that there is a constant B such that (*)
BO(z) >- J(q(eio) - #(p(z))2ke(z) dir(e),
for this would imply, by the argument above, that i(z) z w(z). The proof of (6) was easy, because all we had to do was show that some multiple of ke(z) was >- je(z). This statement is false if we interchange the roles of je and k9, so we will have to work harder to prove the result we want-we have to add up a large number of functions of the form cl[a,b] to make something z ke. Because of this difficulty, we will prove a slightly weaker result than (*) that is still sufficient to prove that the norms are equivalent. (7)
There is a constant B such that for all z e D, B (sup
)?
(z)
J(cp(ei0
- #(p(z))2ke(z)d7t(9).
Proof As we mentioned above, we prove this by adding up multiples of lla bi to make something > ke(z). The first step is to introduce the intervals. Let
Io = [-(1 - r)n,(1 - r)ir] and for n > 1, I,, = 2"I0, let N be the largest integer such that 2N(1 - r) < 1, and write
F. =
10 + n 1 + fl--,-]-IN For a picture of the decomposition, see Figure 7.1.
19b
7
H' ud sM0, .,#' and iM.,re)
FT
Figure 7.1
k8(.9) = 19/(1 + 180(1 - cosO)), Io =
N= 3
Estimating the integral over Io is easy. If 0 e Io, then (4) implies that ( ao)
1-r2 < 1-r2 ke((1-r)2+2r(1-cos0) r) = (1-r)2 = i + r < 2je(r),
so we have (co(eie)
(bo) 20(r) ? J
- Pl(o(r))2ke(r) dn(0).
to
(The reason for this strange numbering will become apparent as the proof goes on.) Estimating the rest of the integrals requires more work. The first step in estimating the integral over I" - In_, is to observe
(a") There is a constant C such that if r > 1/2 and n < N, (1 - r)ke(r) < C4-" for all 0eI,, - In_1.
Proof From (4), it follows that
(1 - r)ka(r) = (l + r) 1 + 2r(1 - cos 0)1 (1 - r) Now if n < N and 0 E I" -
then
1 - cos0 > I - cos(2"-ln(1 - r)),
and we have from calculus that
inf
XE(0,l)
1 - Zosx-e>0, x
7.3
Equivalence of BMO to a SubNpace of :N. w V
197
so
(1 - r)ke(r) < 2(1 + E(2"-17,)2)-1, proving (an). The estimate in (a") takes care of ke(r) on In - In_1. To estimate the rest of the integral fin (7), we let 1/2
IIfIII
=(
If 12 dO) JI"
and
a" =
f
J9 jel
In
F
.
Since I I I I n2 is a norm, II(p
- a0II2
IIp - anII2 + E Ilak - ale-1II2 k=1
The first term 1/2
((p
-a")2dO)
<-II(pII*IInV12'
r"
The second term n
= k=1i Iak - ale-1 I.I1/2 To estimate Iak -ale-l I, we observe thatI dO I) 1/2
Iak - ale-1I = ((ak - ak-1)2)1/2 = (
(ak
-ale-1)2
k -1
(f
k
((P 1
1)2 I
- ale
k-
Ik
+ (f
de1/2
(ale-w)2IIk)
<-II(PII*+(2
II
d61 I)1/2
(p)2 IId81
(ale -
-1 )1/2
<3II(pII*'
Summing the estimates above, we find that
II' - aoII < (3n +
1)II(pII*II"I112,
so
(co - a0)2 d6 < 16n2 II(p
*2nn(1
- r).
Combining this with the estimate (an) and recalling that ao = 0c0, shows that
forn
J n -I"-1
((p - 9(p)2ke(z)dO < C'II(pII2n2/2-n.
198
7
H' and IMO,. w' and L *C)
The last detail is to estimate the integral over [-n, n] - IN. To do this, we observe that IN D [ - n/2, n/2], so cos 9 < 0 on [ - n, n] - IN, and it follows that from the proof of (aN+l) if r > 1 /2 and 0 e [ - n, n] - IN, then
(1 - r)ke(r) < 2(1 +
(2Nn)2)-1
and repeating the arguments used to prove
proves :
(q - Ycp)'ko(z)dO < C'Il(pjI*NZ/2-N
(bN+t)
Adding up the estimates (bo) + (bl) +
+ (bN+l) and recalling that II0I* _
sup, w(z), proves (7).
With (7) proved, it follows immediately (for reasons given after the proof of (6)) that we have (8)
For all z e D, BII(pII*
? w(z).
Combining this result with (6) shows that if we let «(V>>* = sup w(z) Z
(= the M.,#C norm of the associated martingale, Ytp(BtAt)), then A IIq'II* < <
For some arguments in Section 7.4 and beyond, it is useful to write the definition of << >>* in a slightly different way. Let tp e L2(8D, n) and let u = ° p. Since u(z) = E,cp(B,), then w(z) = E,(u(Bj) - u(z))2 = E,u(B)2 - u(z)Z. Since U, = u(B,,,) e .,tl Z, it follows that
w(z) = EZ
so in view of the results of Section 1.11, the condition for q' e BMO can be written as (9)
sup =
f GD(z, W) I Vu(w)12 dw < oo, D
where GD(z, w) =
2
n
log
1 -wz z-w
is the Green's function for D.
7.4 The Duality Theorem for H', Feffermao-Stein DeoompoItion
199
The last result is very similar to a classical analytical characteriilltion of BMO. A positive measure .1 on D is said to be a Carleson measure if there is a constant c such that for every sector
S={re`B:1-h
(p e BMO if and only if the measure defined by 1Vu(z)Izlog111dxdY,
z
where u = 99, is a Carleson measure and, furthermore, the constant K in the definition can be chosen so that Clll(pll* < K((p) <- C211911*
For an analytic proof see Garnett (1980), Chapter 6, Section 3. Even though (9) and (10) are very similar, I do not know how to get the second result from the first.
7.4 The Duality Theorem for H', Fefferman-Stein Decomposition In this section, we will identify (H,)* as BMO. There is a standard way to recover a complex linear functional A from the real part of the corresponding functional on the associated real Banach space : A(f) = Re A(f) - i Re A(if ), so, to make the transition to martingales easier, we consider only real linear functionals. The identification of (HO)* is a four-step procedure. (1)
If A e (Ho)*, then there are functions g, and g2 in L°° (8D, n) such that if f = u + iv c-H', then
A(f)= J
u91+vg2dn.
By results in Sections 7.2 and 7.3, the mapping f --> (u, v)lao identifies Ho with a subspace K' of L' (8D) x L' (8D) in such a way that
Proof of (1)
1If IIHI and 11uIILI + 11v1IL, are equivalent norms. If A is a continuous linear
functional on HO, then it gives rise in an obvious way to a linear functional on K', which by the Hahn-Banach theorem extends to a continuous linear
functional on all of L' x L'. Since (L' x L1)* = L°° x L°°, it follows that there exist g, and 92 E L°° such that (1) holds. The next step is to show:
. #' rnd 940
200
7
(2)
If fl , f2 e Ho and u;, v; are the boundary limits of Refs, Imf , then
H' rnd DMO,
fuiu2dlr = n
an
Proof Let U,'= u1(B) for t < T. Ito's formula implies
u;(B,) = 5 Vu,(Bs) dB, 0
so the formula for the covariance of two stochastic integrals gives
A similar argument shows
and the Cauchy-Riemann equations imply Vv;
_1
=C0
Vu;.
0)
so we have Vu1 Due = Vv1 Vv2 and
EoUa1U.2=Eo
(2) allowsjui us to write
A(f) =
+ u2 d7r, n
where g2 has the obvious meaning: it is the boundary limit of the conjugate function of .g2. The rest of the proof is very easy. Recalling that conjugation is a martingale transform, we see that (3)
91 + 62 a BMO.
Proof Since g1, 92 cL°°, they are in BMO. Let g3. = g2, u; _ Yg;. By results in Section 7.3, g; is in BMO if and only if T
sup E.
I V u; (B) 12 ds < oo.
Z
fo The Cauchy-Riemann equations imply that I Vu2l = I Vu3 1, so it follows that
g3 =g2eBMO.
7.4 The Duality Theorem for II', Fefferman-Stelo Decompwltlon
201
Given the martingale duality theorem and the results of the last Iwo sections, the last step is trivial. (4)
If h c- BMO and we let
A(f) = faD uhdn for all f e H02, then A has an extension that is a continuous linear functional
on H. Ut = u(BtA t). Since Proof Let h E BMO, fE Ho , u = Ref, and let Ht = Uo = 0 and BMO c H2, the usual domination argument shows
uhdrz=EU,,,, H..
Using the martingale duality theorem and the equivalence of the two norms shows EUUHH < 6IIUII1IIHII* < ClI f 11111h1j.
and completes the proof of (4).
Remark: The theorem above makes no mention of Carleson measures, but it is otherwise the same as the one given by Fefferman and Stein (1972), on page 145. The only difference is that we have carried out the two main steps, (3) and (4), using the correspondences developed in the last two sections to reduce these steps to results about martingales. As a corollary of the duality theorem, we get the following result from BMO. (5)
If cp e BMO, then qp can be written as g 1 + 92 + c, where c is a constant and g1, 92cL°°, and furthermore, this representation can be done in such a way that 119111.+119211.
Proof The first conclusion is trivial. If q E BMO, then
f-
f (Ref)cpdn ,Jao
defines a continuous linear functional on Ho, so it follows from the proof of the duality theorem that these are g,, g2 E L°° with
f (Ref )cpdn = f (Ref )(91 +92)dn .Jan
.Jan
for all f E H02, and taking test functions of the form YO, where 0 E L2 and
= 0, shows that cp - (g1 + g2) is constant.
do
202
7
ll' and BMO, . N' and :9 #C
To prove the second conclusion, we observe that, given a cp E BMO, the first equation in this proof defines a continuous linear functional A on Ho with II A II 1 = sup { I AJ I II f II 1 < 11 < C II w II * Now Ho is equivalent to a subspace of L' x L', so it follows from the Hahn-Banach theorem that we can extend A to be a linear functional I, on L1 x L' in such a way that :
I1"II1x1 < C'IITII*1
where IIF 1.1 = sup{Ir(u1,u2)I : IIu1II1 + IIu2II1 < 1} (the constant changes only because we change from the L1 norm to the L' x L' norm). Now any I' E (L' x L1)* comes from a pair (q1, g2) E L°° with IIg111., IIg2Ii. < IIFIIL'xL'
(consider functions of the form (u1, 0), (0, u2)), so the proof is complete. It is easy to generalize the last proof to prove a result about martingales.
To do this, we observe that if A1, ... , Am are matrices without a common eigenvector in R°, then Janson's theorem ((1) in Section 6.7) implies that if we let A° be the identity matrix, then the mapping X - (Ao * X, ... , Am * X)
embeds #' in 21 x
x .1 in such a way that
IIXIII
(6)
(recall that if YE 2', then II YIIy = II YIIy') If tp(X) = E<X, Y> defines a continuous linear function on .,#', then (6) implies that
O(A°*X, ...,Am*X) _ T(X) defines a continuous linear functional on a subspace of 2'' x . . . x Y', which the Hahn-Banach theorem asserts can be extended to the whole space.
Since 21 is isomorphic to
(2' x .. x T')* = ii
x
... , Ym c-.' such that if X °,
we have (2'1)* = #' and, hence, .. x r, that is, there are martingales Y°, .
.
.
,
Xm E 21, then
m
(X°,
,Xm)
m
E(XXY.) _ i=o
r=o
E<X`, Y`),,o,
and hence we have m
E<X, Y> _ (p(X) = (Ao * X, ... ,Am * X) = Y E
If XX= O'H.,-dBs
and
Y, =
f 0
0
then
(A HS, Ks) ds
=
(H.,, AT KK) ds = <X, AT * Y`> J0
,
7.4
The Durllty,rheorem for 11', F'effermrn-Stein DecompoNltlon
20.1
so we have m
E<X, Y>c = Y E<X,AT * Y`). t=o
and, since this equality holds for all X e .,#' ,
Y= I AT * Y` i=O
Now in the arguments above we started with an arbitrary Ye. -V.#0 so it follows from the last equation that we have : (7)
If A1, ..., A. do not have a common eigenvector in R', then any YeR,#O can be written as
Y°+
m
AT*Y`,
with Y`efor i = 0, 1, ... , in, and furthermore this can be done in such a way that II Y`II.
CII Y11*-
Since the proof of (7) uses the Hahn-Banach theorem, it does not tell us how to find the martingales Y°, ... , Ym. In simple cases, this can be done "by hand." Let d = 2,
Example 1
A = (_ 0
\
1
11
0 '
and let Xt = B,'AL where
i=
inf{t : I B, 01 = 11. If we let Y, = BfAt 0 and pick a, b so that (1, 0) = a0 + bAO,
then
Xt=af
tAT
( A S
A0 dB,
,Jo
,10
= (a Y,) + A * (b YY).
Finding a decomposition in general, however, seems to be quite a difficult problem. A. Uchiyama succeeded in doing this recently for d-adic martingales (1982a) and has generalized the construction to BMO functions (1982b), but
I do not know how to prove the result for RMO, and I invite the reader to consider this problem.
The key to deducing the duality theorem for Hp from its probabilistic analogue was the fact that the mapping M (mnemonic for Martingale) defined by M: p
Y p(B,j
t > 0
has I M(p 11 * < C II (p I * . For some applications, we will need to know that the inverse N, defined by I
N:X-+cp(e'e)=E(XXIBT=e`°), is also continuous.
204
7
H' and BMU, .N' and s ,#O
(8)
There is a constant C such that, for all X c- -4,#0,
INXII* <
CIIXII*.
Proof From the duality theorem for Ho and the open mapping theorem (see Dunford and Schwarz (1957), Chapter II, Section 2), it follows that
f fgp dn
If II* < C sup
.
J
OCHO
Ihll,si
Since Bt has a uniform distribution, J(NX)co drz = EE(XXcp(B,)IBL) = E(X,gp(BL))
(all the expectations above exist, since X e . #2 and cP e H2). If we let U = ecp(B,A,:), then it follows from the martingale duality theorem that IE(X,cp(B,))I < CIIXII*IIUIII <-
C'IIXII*.
From (8) and duality, it follows that if N is extended in the obvious way it maps Al' continuously into H'. Combining this result with a trivial result for p > 1 gives the following : (9)
If X e ,# 1 and NX = cp, then there is a unique analytic function f with Ref = Ycp and Imf(O) = 0. If p > 1 and we denote this function by FX, then there is a constant C < oe that depends only on p such that IFXIIp< CIIXIIP.
Proof For p > 1, this result is trivial, since IIFXIIp < CIIReFXIILP(aD) and
Re FX IIp(aD) = E I E(X.I B,) JP
<EE(IX.rIB,) = EIX,IP < EIX*IP = IIXIIp
To prove the result for p = 1, we observe that if Xe.llo, then FXeHO, so the duality result and the Hahn-Banach theorem imply that IIf111 _< C sup IIw
J(Ref)P d ir.
7.5
Examples of Mardnades In tt 4 ')
205
If we let u = 9ap, then we have
J(Re FX )cpdn
E(E(X,IB)u(B)) = EE(XCO U. I B)
= E(X, U.) < CIIXII1IIUII*
..#2.
Since .jf 2 is dense in .,# 1,
the desired result follows.
Remark: The last result is false when p < 1. It is known that there are nontrivial continuous linear functionals on H° for p < 1 (see Duren, Romberg, and Shields (1969)). In Section 7.8, we will show that there are no nontrivial continuous linear functionals on .,lf°, p < 1, so X -> FX cannot be continuous. Note: We learned the results on M and N, (8) and (9) above, from Varopoulos (1979), who attributes the results to Maurey. Exercise 1 Sharpen (3) by showing that if qp a BMO and Cp is the conjugate function, then Cp e BMO and
«w»* _ «w»*, IIwII* <- CIIgII* As a corollary of the last result, we see that if cp e L°°, then CP e BMO, a result first proved independently by Spanne (1966) and Stein (1967). The function f(z) = i log z shows that we may have tp e L°° but Cp 0 L°°.
7.5 Examples of Martingales in R ,#C In Section 7.1, we defined the -4.,#C norm II X II * of a martingale to be the smallest number c such that for all stopping times T (1)
E(I X, - XTI) 5 cP(T < oo).
In this section, we will try to explain what kind of martingales are in -4.#6 by giving four examples and describing some of their properties. Example 1
If X c- . e l f °° = {X : X* e L°° }, then it is immediate that X e RJie2
and I I X II * _< 2II X II
.
With a little thought, we can replace the 2 by a 1:
(EIXc -XTI)2 <E (X. -XT)2 EX .2 -EXr2. <- IIXIIW.
206
7
H' and HMO, . N' and .4.4#0
There are also unbounded martingales in M..#O. Perhaps the simplest example is the following. Example 2 Let Xt = Bt 1. If T is a stopping time, then Xm - XT = B1 - BT 1
is independent of . and has a normal distribution with mean 0 and variance (1 - T)+. Since Bc c1/2B1 and EIB1I =
2JrO
(27r)- 1/2xe-x2/2dx = (2/7C)112,
it follows that E(I X. - XTII
(2/7r)112P(T < oo)
and, taking T - 0, that IIXII * =
(2/it)112
Once you get started, it is easy to use Brownian motion to construct
numerous examples of unbounded martingales in 1f0. An obvious modification of the last example is to let S be a random variable that is independent of B, t > 0, and has an exponential distribution P(S > t) = e-xt. If we let Xt = Bs, then the lack-of-memory property, P(S > t + ul S > t) = P(S > u), of the exponential implies that X is in .4.,610. Another possibility is to introduce
the stopping times R" = inf{t > Rn_1 : I B(t) - B(Rn_1)I > 1} and let Xt = B(t A RN), where N is an independent random variable with a geometric distribution P(N > n) = (1 - p)", n = 0, 1, 2.... The next example is a variation
of this-N depends on B, t < TN, and is chosen to try to produce a large maximum. Example 3 Let Ro = 0, and for n >- 1, let R" = inf{t > Rn_1 : I B(t) - B(Rn_1)I
> 11. Let N = inf{n : B(R") - B(Rn_1) = -1 } and let Xt = B(t A RN). See Figure 7.2 for a sample path. It is easy to get a bound on the 8.610 norm of X. If T is a stopping time, then on {Rk < T < Rk+1, N > k}, XT = k + a for some ae(-1, 1), and P(N > k + jI.FT) <- (1/2)j-1 (the worst case is a - 1). On the other hand, a look at Figure 7.2 shows that I X. - XTI < N - k (the worst case is a -> - 1), so combining the estimates gives that E(I X. - XTI I.FT) < E(N -
2,
and we have IIXII* < 2. A look at the parenthetical remarks above shows that this is not the right answer, but we are not far from it. It is easy to compute f(a) = E(I X. - XTI I XT = a, T < R1) and maximize to find IIXII* = 9/8 (details of this computation are sketched in Exercise 1 at the end of this section). We have taken the trouble to compute the -4A'(9 norm of the last example so that we can check the accuracy of some of the inequalities below. By abstracting the construction above, you can produce many examples of martingales in .4.610. The next construction is so general that we can think of it as giving the typical element of -4,NO (see the proofs of (1) and (5) in the next section).
7.5
207
Examples of Msrtlogsles 1n .W.NP"
N=5
R,
T
R,
R,
R,
R,
Figure 7.2
Example 4 Let T" be an increasing sequence of stopping times with To = 0
and I B(t) - B(Tn_1)I < A for t e [T"_1, T"], n >- 1. Let N be a stopping time
for .F(T"), that is, {N < n} E.F(T"), and suppose that for all n > 1, P(N > (Tn_1)) < 0 < 1 on {N > n - 1}. If we let X, = B(t A TN), then X E R.,W0 and as the reader can easily show
nI.
IIXII* < A(1 + (1- 0)-1). The four examples above should give you an idea of what type of martingales are in -4.Af(9 and, hopefully, make the results in the next two sections more obvious. To lead you in the direction of the John-Nirenberg inequality ((1) in the next section), we will now compute P(X* > A) for the three unbounded examples. Example 2.
P(sup X, > ..) = 2P(B1 > i±) =
P(X * > 2) < 2P(sup, Xt > 2) =
4
2
e k2 /2 r
e-v212 dy.
exze X212 dx
21T
4
e x2/22-1. As 2 -+ o o.
2n
Example 3.
If n is a positive integer, then
P(X*>n)_(l/2)".
2011
7 Wand IMO.. N' and :4..WE)
E x a m p l e 4.
I f n is a positive integer, then
P(X * > nA) < P(N > n) < (1 - 0)". Exercise 1
Compute the -4.,lfC norm of the martingale in Example 3.
(a) By the strong Markov property and independent increments of Brownian motion, it suffices to consider the case XT = a e (- 1, 1) a.s. on IT < co}. (b) When XT = a, the distribution of IXX - XTI is given by the following table :
X.
I XX -XTI
-1
a+1
Probability (1-a) 2
( a +1 ) 2 /I)n+i
2
n
n-a
(a 2 1)
I\
Summing over the possibilities, we get
E(IX.-XTI)- (a+1)(2-a)/2 ifa>-0 P(T
1(a+1)(1-a)
if a<0.
The maximum value occurs for a = 1/2, and here the value is 9/8.
7.6 The John-Nirenberg Inequality In this section, we will prove a classical result on BMO functions due to John and Nirenberg (1961), which was discovered almost a decade before it was known that (H1)* = BMO. Following our usual inclination, we will prove the probabilistic analogue first and then use this result to prove the analytical result. (1)
There is a constant Ce(0, oc) such that if IIXII* < 1, then
P(X* > A)
The first step is to prove :
If Ye.4,06 and S > T are stopping times, then EI YS - YTI < II YII*P(T < oc).
Proof If T is a stopping time, then Z, = YY - Y, AT is a martingale and Iz l is a submartingale which is dominated by an integrable r.v. (recall that Ye.,A'1),
209
The John-Nlrenberg Inequality
7.6
so the optional stopping theorem implies that
El Ys- YTI =EIZSI <EIZmi =EIY. - YTI, proving the result, since, by definition of II Y II *
EIY. - YTI <-IIYII*P(T
P(Ri < c0)
EIXR.+l - XR.I > aP(Ri+1 < oc).
Therefore, P(Ri+1 < co IRi < oo) < 1/a and, by induction, P(R,, < oo) < (1/a)"P(R0 < oo). Let n be an integer and let an < 2 < a(n + 1). Then
P(X* > )) < P(X* > an) < P(R" < oo) < (1/a)"P(Ro < oo)
< a(l/a)'I° = ae-'(""I'. Setting a = e, to maximize (log a)/a, gives (1).
Remark: To summarize the proof in loose language : The amount of wiggling a martingale in -4,1f0 can do has a geometric upper bound because of (2),
and, hence, even if all the movement occurs in the same direction, a large maximum will not be produced. To translate (1) into a result about BMO functions, we recall that in Section
7.3 we showed that if q E BMO and we let u = 9vp and U, = u(B,A,), then UE. J1(!9 and there are constants c, Ce(0, oo) such that cII9II* < II UII* < CII(P
(*)
Combining the last observation with (1) gives: (3)
The John-Nirenberg Inequality. There are constants C, y e (0, oo) such that if * < 1 and $ cp do = 0, then {0 : I4 (ete)l > Al I < Ce-".
II
II
Proof (*) implies that II U II * < C, so 110: Iq(e`e)l > All < 21tP(U* > A).-5 C'e-'Ice.
It is easy to improve (3) to the usual result. (4)
Let C, y be the constants in (3). If II'II* < 1, then for all intervals I, {O E I :
I q(e`e) - cprl > All <
Ce-yxlI1.
Proof If! = (a - hn, a + hit), then >V (eie) _ cp(exp(i(a + Oh))) - gyp is in BMO with I IV/ II * < 1 and 10 = 0, so (4) follows from applying (3) to 0. The conclusion of (1) can be improved in a similar way.
210
7 W .ndBMo,.M' snd91_
(5)
Let C be the constant in (1). If IIXII* S 1, then for all stopping times T, P (sup I tzT
- XTI > A) :9 Ce--"P(T < oc). /l
Proof This could be proved in the same manner as (4), but it seems easier to observe that repeating the proof of (1) with Ro = T proves (5). From (5), we see that the definition of -4,#C as sup E(I
XT I
I T < oo) < oo
T
implies the much stronger conclusion that (6)
For all a < (eIIXII*)-1, supE(ealxw -xTIIT < oo) < o0 T
(in both cases, the supremum is over all stopping times). From (6), it follows that ...,df0 can be defined as the set of all martingales with
supE(IXc,, -XTI"IT
There is a constant C < oo such that IIXII* <- <X»* <- CII X II*
Proof We proved the left-hand inequality in Section 7.2. To prove the righthand inequality, we observe that it follows from (5) that if IIXII* 5 1, then
(oo),
E(X. - XT)2 < fO'O 22P (sup I Xt - XTI > .1 t>_ T
<
J
proving (7).
As a corollary of the John-Nirenberg inequality, we get the following striking result due to Zygmund (1929): (8)
If (p is continuous on 8D, then its conjugate function Cp has
JexP()(o)I)do
< oo
for all A < oo .
Proof Since every continuous (p is the uniform limit of trigonometric polynomials, we can, for any e > 0, write qp = gyp, + P2, where gyp, is a trigonometric
7.7
211
The G.rnett-Jones Theorem
polynomial (and hence has a p , e L°°) and 92 has 1 1P 21 1 , < II cp2 IIm ` r, u i I he desired result follows from (6). On page 253 of Zygmund (1959), you can find a continuous cp whose conjugate ap is unbounded: cp(e")
Exercise 1
°°
Y
=
sinnO
nlogn
cp(e
_
°°
cos n9
= n=2 Nogn
Calderon's Proof of the John-Nirenberg Inequality. Let
a(x) = sup{P(X* > x) : IIXII* < 1}.
By stopping at the first time, XXI > x, we see that a(x) < 1/x and a(x)a(y) > a(x + y). Taking logs gives log a(x) + log a(y) > log a(x + y), and one easily concludes from this that lim sup 1 log a(n) = inf 1 log a(m) < 0, R- D
n
m>1 m
proving the result with a constant that is not explicit, but is trivially the best possible.
7.7 The Garnett-Jones Theorem In this section, we will determine which martingales in -4.lf(9 are almost bounded, or, to be precise, we will find the closure of .,#°° in
The solution
to this problem is again the probabilistic analogue of a previously proven analytical result, and the analytical result can be recovered from the probabilistic one. To state the result, we first need some notation. Let ao(X) be the supremum of the set of all a so that supE(ealXw_XTIIT < oc) < oc
T
where the sup is taken over all stopping times (by (6) in Section 7.6, ao(X) > (eII X II*) ' > 0). Intuitively, ao(X) is the exponential rate at which P(I XX - XTI > .) goes to zero for the worst choice of T. Given this interpretation, it should not be surprising that ao(X) can be used to measure how well an X in M MO can be approximated by a Ye.#°°. (1)
There are constants c, Ce(0, oo) such that c
ao(X)
< inf IIX - YII* < Y.M-
C ao(X)
and, consequently, the closure of .,lf°° in R.&(9 is {X : ao(X) = oc }.
212
7
H' and BMO.. #' and 4U.wtV
Remark: This result and the proof we will give below are due to Varopoulos (1979).
Proof The inequality on the left is an easy consequence of the John-Nirenberg
inequality. If YeA' and IIX - Y I* = a, it follows from (5) in Section 7.6 that if Z = X - Y and a > 1/e, then E(e°ho_ZTlla) < C,,P(T < oc), so
P(I Za, -ZTI > 2) < Let Ao = II YII, If K > 2 and A > K.,, Then
P(IX(-XTI>2)
e-Qa(K-2)/KaP(T
< Ca
< oc).
At this point, we have shown that ao(X) > a(K - 2)/Ka. Letting a 1/e and K-> oo gives that ao(X) > Ilea, so the left-hand inequality in (1) holds with c = 1/e (the constant is inherited from the John-Nirenberg inequality). To prove the other inequality is more difficult. Given a large a and a martin-
gale Xe.A.J1 with ao(X) > a, we need to construct Ze.#' with IIx - Z Ih < C/a. This construction is accomplished in two stages. First, we construct an approximating martingale of the form described in Example 4 of Section 7.5, but we are forced to take 2 large to make 0 < 1, and we end up with a M'#6 norm that is too large. Then we use an ingenious construction, due to Varopoulos, to introduce a sequence of stopping times that smooths the transitions
between the times constructed in our first attempt and reduces the M.# C norm to the right size. The first part of the construction is straightforward. We let 2 be fixed and define R. inductively by letting Ro = 0 and Rn = inf{t > Rn_1 : I X(t) - X(Rn_1)I
> Al for n >- 1. If we let n = X(R,,) - X(Rn-1) on {Rn < oo} and Z =
Y-, S. 1(Rn
).
Z, will differ from X, by a martingale that is bounded. To see this, observe that when Rn < t < Rn-1,
X,-Z,=X,-E(> m1(Rm
= X, - X(Rn) +
E(
Y_
\m=n+1
Sm1(Rm
From the definition of the R, it follows that I X, - X(R,,) I < A. To get a bound on the other term, observe that if supE(e2 a_XTII T < oo) < K, T
7.7
213
The Gernett-Jones Theorem
then P(Rm+1 < oo I Rm < oc) < K/eau Therefore, if Ke -a l < 1, then (*)
E( m=n+1 "0
Sm1(Rm<W)I/
< 1(1 - Ke
ax)-1
JF
and we have Y, = X, - Z, c-,#'. Y, = X, - Z, is not the martingale in .,lf°° that we want. From Example 4 in Section 7.5, we see that the 8..110 norm of Z satisfies IIZII* < 2(1 + (1 Ke-az)-1), but this is too big because we have no control over K and, hence, over the choice of 2. To circumvent this difficulty, we use the construction referred to above. Pick 0 < 1 and then 2 so large that Ke-az < e-ax° and (for convenience) 2a = M is an integer. Let y = e-e, let So = 0, and, for n > 0,
1<j<M,let
SfM+j = inf{t : P(Rn+1 < co l.wt) > .
Since P(Rn+, < oo I F(Rn)) <
M-j}.
Ke-a l < yM, we have
R. < SnM+1 < ... < S(n+1)M :!
Rn+1
Let U"
_
I
M
M
Y 1(SnM+j
U" is a staircase that allows us to climb from { U. = 0} = {S"M+1 = oo } to {U" = 1} = {S(n+1)M < oo}. Let 00
M=1
Z, = E(Z' I If Rn < t < Rn+, , then
14 - z;I = E( m=n
m(1(Rm
\m=n+1
<2+2(1
-
Keal)-1
by the arguments used to prove (*) (observe that I m(1(Rm
Z' _
°°
Y
Ym(k)
k=1
1
M
1(Sk<00)'
214
7
H' sod lAW..N' and M.r Or
and we see from Example 4 in Section 7.5 that for all 0 < 1,
IIZ'II* < M(1 +
Letting 0T 1 proves the desired result with C = (2e - 1)/(e - 1) = 2.42. Example 1
Let X, be the martingale in Example 3 of Section 7.5. It is easy to see that ao(X) = log2 = .693, so (1) gives 1
.5307 =
e(.693)
2.42 < inf IIX- Z II* <= 3.49 693 ZEN
The upper bound is hardly informative, since
II X II
s = 9/8 = 1.125, but the
computation suggests the following question I could not answer. In Example 3, do we have
Problem
inf IIX- ZII* = IIXII*?
ZE.R'0D
An even simpler question is: Can you construct an example with this property? With (1) established, we turn our attention to proving the analytical result.
Since this is a thankless job and requires more work than a direct analytical proof (see Garnett and Jones (1982)), we just sketch the details. (2)
The Garnett-Jones Theorem. Let ao(cp) be the supremum of the set of a that has
sup1
exp(alrp-q,iI)d9
where rpr =
f 9 dB. r
There are constants c, Cc(0, oo) such that
< C) ` inf IIq'-'II*a ao((P) C
'P
Proof As in the proof of (1), the left-hand inequality follows easily from the John-Nirenberg inequality, so we leave the details as an exercise for the reader and turn now to the proof of the more difficult right-hand inequality. From the developments above, the plan of attack should be clear. We pass from 9 to U, = Yrp(B,A) = M(p, decompose U, = XX' + X,' according to the construction in the proof of (1), and then let tij(e`B) = E(XX I Bz = e`B) = NXj. From results about the maps M and N defined in Section 7.4, we see that to prove (2), it suffices to show that (3)
ao(T) < Cao(Mtp).
7.8
215
A Disappointing Look at (..N') When p < I
for then it will follow that 11(p - 0III*
CII U - X`II * < ao(U)
` «o(p) The proof of (3), however, is tedious and very similar to the proof of (7) in Section 7.3, so the reader is referred to Varopoulos (1979) for the details. It is inevitable when we use the correspondence between BMO and R.#O (or H" and .,#P) that the resulting inequalities will not be very precise. In the case of (2), an analytical proof gives a much stronger result. As Garnett and Jones (1978) remark, if we norm BMO by is constant}
(which we know from Section 7.4 to be an equivalent norm), then it follows from results of Helson-Szego (1960) and Hunt, Muckenhoupt, and Wheeden (1973) that n
gcLf
III-9II**-2ao((p)
(see Garnett (1980), Section 6). Varopoulos (1980) has shown (see his Theorem 4.2) that there is an analogous result for R.,# !2,
«X-Y>>*<2 «o
1
r inf and that the constant is the best possible. The reader is referred to his paper for the details of this and many other interesting developments (including a probabilistic proof of the Corona theorem).
7.8 A Disappointing Look at (J )) * When p < 1 In this section, we will consider the martingale spaces .,#P, 0
sponding HP space, which in the case 0 < p < 1 was first found by Duren, Romberg, and Shields (1969) (see Duren (1970), pages 115-118). In this section, we will show that this approach is not feasible-there are no nontrivial continuous linear functionals on .,#P, 0 < p < 1. This result, attributed to P. A. Meyer in a footnote in Getoor and Sharpe (1972), is in sharp contrast with the results for dyadic martingales obtained by Herz (1974b). In the dyadic case, a martingale is in (.,'P)* if and only if there is a constant C such that, for all stopping times T, E(I Y,,,, - YTI I FT) < CP(T < oo)1/P. In this section, we use the techniques of Section 7.1 to show that the same result is true for the
Brownian filtration and that, unfortunately, no nonconstant martingale has this property.
216
7
H' ind BMO..N' mod W Vr)
The first step in our study of the martingales in . KP is to obtain an atomic decomposition. The arguments given in Section 7.1 generalize with very little work to the case p < 1. (1)
A martingale A e d#P is said to be an atom if there is a stopping time T such that
(i) A,=Oift
IA,IP
If we let dd(X) = E(X *)P, then it follows from the definition that (2)
If A is an atom, dd(A) = E(A*)P < 1.
(3)
If A" is a sequence of atoms and Yn I cnIP < oo, then as N - oo, XN = Y _N cnAn converges to a limit Xedi", and dd(X) < En IcnIP.
The proof of (4) in Section 7.1 leads to the following : (4)
For all X e #P, there is a sequence of atoms A", n e Z, and a sequence of constants cn, n e Z, such that
(i) asN -oo, _n and (ii) En IcnIP < 2(21/P + 1)Pdd(X).
The proof is the same as the proof in Section 7.1. Let
T"=inf{t:IXI>2"/P} A" _ (X(t A Tn+1) - X(t A T"))/c" cn = (21/P + 1)2' "P(T < oc)1/P.
It follows from computations in Section 7.1 that (i) holds. To check (ii), we observe that dP(>cnA"J < ICnIP
_ (21/P + I)PY2"P(T < oc)
"
_
(21/P
+ 1)P Y 2"P((X *)P > 2") n
< 2(21/P + 1)E(X*)P.
Remark: In the argument above, the fact that p < 1 is important. If p > 1, the triangle inequality gives that cnA"I m-
P
<_ F.lcnl = n
(2"P
+ 1)Y(2"P((X*)P >
2"))1/P,
and the right-hand side > (21/P + 1) (yn 2"P((X *)P > 2"))1/P. Since the spaces .diP, p > 1, are easy to handle directly, we leave it to the reader to figure out how to define atoms for p > 1.
7.8
217,
A Dlwppolnting Look of (A")* Whenp < I
With the atomic decomposition established, it is easy to find the droll of .#P. From results in Section 7.2, if 6p is a continuous linear functional on . NP,
then there is a Ye.,lf2 such that if Xe.,N2, (p(X) = EX. Y.. To see what properties Y must have, we follow (7) of Section 7.1. Let T be a stopping time,
let Z. = sgn(Y. - YT), and let Z, =
Since Z,i 5 1, X, = (Z, ZTAr)/2P(T < oo)l/P is an atom. By a computation in Section 7.1, E I Y. - Y1.I =
E(Z.(Y. - YT)) = E((Z,0 - ZT) Y.), SO
EIY - YTI
E(X,
2P(T < 00)11P'
Taking the supremum over all stopping times, we see that there must be a constant c such that for all T, (5)
EI Y - YTI < cP(T < oo)'IP.
If p = 1, this inequality reduces to (6) of Section 7.1, so on the surface everything is fine. We have a family of spaces, which we might call MP, 0 < p < 1, that generalize R..#D and have (.,NP)* c MP. Unfortunately condition (5) is too strong when p < 1: (6)
If p < 1, then the only martingales Ye 4'2 that satisfy (5) are constant.
Proof The idea of the proof is simple. Suppose that C = 1. If we can find a T such that 0
and if we can pick an > 0 and define another stopping time
T' =
(T 00
if E(I Y.,, - YTI I FT) > E
otherwise
with P(T' < oo) = S, a small number chosen to have PIP-' < E, then we are done, since if (5) holds with c = 1, we have
P(T' < oo)1/P >_ E(I Y. - YT.I) > eP(T' < 00), which implies that 6'IP-1 > e, a contradiction. The technical problems we now face are (a) to find a stopping time with E(I Y. - YTII JFT) * 0 (easy) and (b) to shrink P(T < oo) to an appropriate size. To solve these problems, we observe that since every continuous local martingale is a time change of Brownian motion, it suffices to prove the result when Y, = B,A, and z is a stopping time. Let T. = inf{t : IB,I > a). Since Ye.,N2 and I Y. - Y,I < 2Y*, it follows from Theorem 9.5.4 in Chung (1974) that as E(I Y. - Y(Ta)II _F(Ta)) -'E(IYYII n a>o
.
(Ta)).
The right-hand side is E I Y. I (exercise), but we do not need this. The expectation
218
7
H' ud DM0, .N' rnd A,#&
of the limit is I Y,,,1, which is > 0 if Y * 0 (recall that YY = E(YY IA,)), so it must be positive with positive probability, and hence, if a is small enough and T = Ta, then E(I Y. - YTII.`FT) # 0, and we have accomplished (a). Since T' has a continuous distribution, the solution of (b) is trivial. We let I
T, - IT 100
if E(I Y. - YTII FT) otherwise
and 0 < T < A
and vary A to make P(T' < oo) the right size.
Remark: The last part of the proof obviously breaks down for dyadic martingales. In that setting, if you want a fixed value for the stopping time, say T = 1, then the probability of taking on that value cannot be arbitrarily small. It is this curiosity that allows nontrivial examples on the dyadic filtration.
8 PDE's That Can Be Solved by Running a Brownian Motion
A Parabolic Equations In the first half of this chapter, we will show how Brownian motion can be used to construct (classical) solutions of the following equations: U, = 20u
ut=2Au+g u,=ZOu+cu in (0, oo) x R" subject to the boundary condition: u is continuous in [0, oo) x Rd and u(0, x) = f(x) for x e Rd. The solutions to these equations are (under suitable assumptions) given by Ex(f(B,))
J(t E(f(B)+-s,BS)ds) \
E. (f(B) exp (f o
c
(BS) ds)
-
b(BS) dB, fo I b(BS) 2 dso /I \J 2 In words, the solutions may be described as follows : E. .f(Br) eXp
(i) To solve the heat equation, run a Brownian motion and let u(t, x) _ Ex.f (B,)
(ii) To solve the inhomogeneous equation u, - ZAu = g, add the integral of g along the path. 219
221)
8
PDE's That Can Be Solved by Running a Brownian Motion
(iii) To introduce cu, multiply f(B,) by exp(f o c(B5) ds) before taking expected values. In more picturesque terms, we think of the Brownian particle as having mass 1 at time 0 and changing size according to m' = c(B,) m, and when we take expected values, we take the particle's weight into account. (iv) To introduce b Vu, we multiply f(B,) by what may now seem to be a very strange-looking factor. By the end of Section 8.4, this factor will look very natural, that is, it will be clear that this is the only factor that can do the job. In the first four sections of this chapter, we will say more about why the expressions we have written above solve the indicated equations. In order to bring out the similarities and differences in these equations, we have adopted a
rather robotic style. Formulas (2) through (6) and their proofs have been developed in parallel in the four sections, and at the end of each section we discuss what happens when something becomes unbounded.
8.1
The Heat Equation In this section, we will consider the following equation:
(1)
(a) u, = 'Au in (0, oo) x Rd (b) u is continuous in [0, oo) x Rd and u(0, x) = f(x). The equation derives its name from the fact that if the units of measurement are chosen suitably and if we let u(t, x) be the temperature at the point x c Rd at time t when the temperatures at time 0 were given by f(x), then u satisfies (1).
The first step in solving (1), as it will be many times below, is to prove : (2)
If u satisfies (a), then M, = u(t - s, B.) is a local martingale on [0, t).
Proof Applying Ito's formula gives s
- u,(t - r, B,) dr
u(t - s, BS) - u(t, Bo) = 0
+ JVu(t - r,B,)dB, 0
Du(t - r, B) dr,
+2 0
which proves (2), since - u, + i Au = 0 and the second term is a local martingale.
If we now assume that u is bounded, then M, 0 < s < t, is a bounded martingale. The martingale convergence theorem implies that as s T t, M, converges to a limit. If u satisfies (b), this limit must bef(B), and since MS is uniformly integrable, it follows that
Ms = E,,(f(B,)I3 ) Taking s = 0 in the last equation gives us a uniqueness theorem.
8.1
(3)
221
The Nest Equation
If there is a solution of (1) that is bounded, it must be v(t, x) = Exf(B,).
Now that (3) has told us what the solution must be, the next logical step is to find conditions under which v is a solution. It is (and always will be) casy to show that v is a "generalized solution," that is, we have (4)
Suppose f is bounded. If v is smooth (i.e., it has enough continuous derivatives so that we can apply Ito's formula in the form given at the end of Section 2.9), then it satisfies (a).
Proof /The Markov property implies that Ex(f(B,)IJ1s) = EB(S)(f(Bt-s)) = v(t - s, B.).
Since the left-hand side is a martingale, v(t - s, BS) is also. If v is smooth, then repeating the calculation in the proof of (2) shows that
v(t-s,B.)-v(t,B0)= fo (-vt+ZAv) (t- r, B,) dr 5
+ a local martingale, so it follows that the integral on the right-hand side is a local martingale. Since this process is continuous and locally of bounded variation, it must be = 0, and hence, - vt + zAv = 0 in (0, oo) x Rd (vt and Ov are continuous, so if - vt + Av 0 at some point (t, x), then it is 0 on an open neighborhood of that point, and, hence, with positive probability the integral is * 0, a contradiction). It is easy to give conditions that imply that v satisfies (b). In order to keep the exposition simple, we first consider the situation when f is bounded. In this situation, the following condition is necessary and sufficient: (5)
If f is bounded and continuous, then v satisfies (b).
Proof (Bt - Bo) d= t1"2N, where N has a normal distribution with mean 0 and variance 1, so if t 0 and x -+ x, the bounded convergence theorem implies that
v(t,
Ef(x + t, 2N) -f(x). The last step in showing that v is a solution is to find conditions that guarantee that it is smooth. In this case, the computations are not very difficult. (6)
If f is bounded, then v e C°° and hence satisfies (a).
Proof We will show only that v c C Z, since that is all we need to apply Ito's formula. By definition, v(t, x) = Exf(B) =
$(2xt)_42e_1x_Y1212tf(Y)dY.
222
8
PDE'. flit ('an Be Solved by Running a Brownian Motion
A little calculus gives Die-Ix-yl2/2t =
-(xi -
Yi)e-Ix-y12/2tt
- Yi)2 Dve-Ix-yl2/2r = (xi - Yi)(x, Dite-Ix-yl2/2r = ((xi
t)e-Ix-yl2/2tt2 Y,)e-Ix-yl2/2rt2
i 5 j.
If f is bounded, then it is easy to see that for a = i or ij,
f
IDe-Ix-yl2/2tl
I f(Y)I dy < o0
and is continuous in R°, so the result follows from our result on differentiating under the integral sign (an exercise at the end of Section 1.10). For some applications, the assumption that f is bounded is too restrictive. To see what type of unbounded f we can allow, we observe that, at the bare minimum, we need Ex I f(B) I < oo for all t. Since I E.If(Br)I = J (2nt)
wee-Ix-yl2/2tIf(y)Idy,
a condition that guarantees this is (*)
IxI 2log+If(x)I -.0
as x- oo.
By repeating the proofs of (5) and (6) above and doing the estimates more carefully, it is not hard to show: (7)
If f is continuous and satisfies (*), then v satisfies (1).
Note: All we have done in this section is rewrite well-known results in a different language. An analyst (see, for example, Folland (1976), Section 4A) would write the first equation as
Btu-Du=0 u(0, x) = f(x)
and take Fourier transforms to get
t) -
t) = 0 =f(e)e-4,M21CI2r
t)
Now KK(x) = (47rt) -d/2 e- 1-"1'/4' has e-4n21C12t,
so it follows that u(t, x) = JKt(x - Y)f(Y) dY,
and we have derived the result without reference to Brownian motion.
8.2
The Inhomogeneous Equstton
223
Given the simplicity of the derivation above, we would be foolish to chiim that Brownian motion is the best way to study the heat equation in (0, cx)) x R. The situation changes, however, if we turn our attention to (0, oo) x G, where G is an open set (e.g., G = {z : IzI < 1}), and try to solve (1')
(a) u, = 'Au in (0, oo) x G (b) u is continuous in [0, oc) x G and u(0, x) = f(x) x e G u(t, x) = 0 t > 0, x e 8G. In this context, the analyst (for example, Folland (1976), Sections 4B and 7E) must look for solutions of the form f (x) exp(A,jt), "separation of variables," and show that the initial condition can be written as
f(x) _
a; f (x).
Proving this even in the special case G = {z: Izi < 1} requires a lot more work than when G = Rd, but for Brownian motion the amount of work is almost the same in both cases. We let T = inf{t : B, t G } and let
v(t, x) = EX(f(B); t < T). Repeating the proofs above shows
(2)
If u satisfies (a), then MS = u(t - s, Bs) is a local martingale on [0, T A t).
(3')
If there is a solution of (1) that is bounded, it must be v(t, x).
(4)
If v is smooth, then it satisfies (a).
(5')
If f is bounded and continuous, then u is continuous in [0, oo) x G and u(0, x)
f(x). If the reader is patient, he or she can also show that (6')
If f is bounded, then v, Div, and Djv, 1 < i, j < d, all exist and are continuous, so v satisfies (a).
Note: v will not necessarily satisfy the other boundary condition u(t, y) = 0 for y e G. We will discuss this point when we consider the Dirichlet problem in Section 8.5.
8.2 The Inhomogeneous Equation In this section, we will consider what happens when we add a function g(t, x) to the equation we considered in the last section, that is, we will study (1)
(a) u, = 'Au + g in (0, oo) x Rd (b) u is continuous in [0, oo) x Rd and u(0, x) = f(x).
224
8 PDE'e Thai ('en Be Sowed by RunIa a Brownlen Motion
The first step is to observe that we know how to solve the equation when g = 0, so we can restrict our attention to the case f 0. Having made this simplification, we will now solve the equation above by blindly following the procedure used in the last section. The first step is to prove (2)
If u satisfies (a), then
M, = u(t - s, Bs) + f sg(t - r, B,) dr 0
is a local martingale on [0, t).
Proof Applying Ito's formula gives
u(t - s, Bs) - u(t, Bo) = fos (-ut +
ZAu)(t - r, B,) dr + fos Vu(t
- r, B,) dB,,
wh ich proves (2), since - ur + Au = -g and the second term is a local martingale.
If g is bounded and u is bounded on [0, t] x R" and satisfies (a), then M 0 S s < t, is a bounded martingale. By the argument in the last section, if u satisfies (b), then
lim M, = fo g(t - s, B,) ds ,Tr
and
g(t - s, Bs) dsl).
M, = Ex o
Taking s = 0 gives Suppose g is bounded. If there is a solution of (1) that is bounded on [0, t] x Rd, it must be //
v(t,x) = ExI
g(t - s,Bs)ds). r,B,)drl.°Fs/
\ fo Again, it is easy to show (4)
Suppose g is bounded. If v is smooth, then it satisfies (a) ax, in (0, oo) x Rd.
Proof The Markov property implies that rg(t -
Ex\J
s
t-s
J(t-r,B,)dr+EB(.,)(fo
l
225
&2 7% I.homoQsn.ou. Equatlon
Since the left-hand side is a martingale, it follows that J3
v(t - s, B) +
g(t - r, B,) dr 0
is also. If v is smooth, then repeating the calculation in the proof of (2) shows that s
g(t - r, B,) dr
v(t - s, Bs) - v(t, Bo) + J0
= f (-vt+ZAv+g)(t-r,B,)dr 0
+ a local martingale,
Again, we conclude that the integral on the right-hand side is a local martingale and, hence, must be = 0, so we have (- vt + 4v + g) = 0 a.e. The next step is to give a condition that guarantees that v satisfies (b). As in the last section, we will begin by considering what happens when everything is bounded. (5)
If g is bounded, then v satisfies (b).
Proof If I g I < M, then rt
g(t - s, Bs) ds
E.
<Mt- 0.
J0
The last step in showing that v is a solution is to check that it is smooth enough. In this case,
v(t,x) = J'dr J(2xs)_2e__2/2sg(t - s,Y)dy, 0
and the normal density (2irs)-d/2e-Iv-XI2/2s
PS(x,Y) _
- oo if x = y and s -+ 0, so things are not as simple as they were in the last section. The expression we have written for v above is what Friedman (1964) would call a volume potential and would write as V(x, t) =
f f
Z(x, t ; , T)
i) d di.
J To D
To translate between notations, set To = 0, D = Rd, Z(x, t ; , T) = Pt-,(x, ),
and change variables s = t - r, y = . Because of their importance for the
226
$
PDE'e That (:nn Be Solved by Ruuing . Brownlen Motion
parametrix method, the differentiability properties of volume potentials are well known. Since the calculations necessary to establish these properties are quite tedious, we will content ourselves to state what the results are and indicate why they are true. The results we will state are just Theorems 2 to 5 of Friedman (1964), so the reader who is interested in knowing the whole story can find the missing details there. (6a)
If g is a bounded measurable function, then v(t, x) is continuous on (0, oo) x R°.
Proof This result follows easily from the bounded convergence theorem. (6b)
If g is bounded and measurable, then the partial derivatives Div = avlax; are continuous, and Div = J' JDP(xY)(t - s,Y)dyds.
Proof The right-hand side is
-
f f ,J
0
(2ns)-d/z (Xi - Yi) e- Ix-vl2/2sg(t S
,J
- s,Y) dY ds
= - JEx[(xi oS
-
B
s/g(t - S, B)]
Although the last formula looks suspicious because we are integrating s-1 near 0, everything is really all right. If lgl < M, then Exl(x1 - B5)g(t - s, BS)l < MEEI x; - BsI = CMs1'2, so
- B8)g(t - S, BS)l < 00
f
0
and it follows from the exercise in Section 1.10 that the partial derivatives Div exist and are continuous. The computations above are not hard to make rigorous, but you should save your strength. Things get very nasty when we take second derivatives. (6c)
Suppose that g is a bounded continuous function and that for any N < oo there are constants C, a e (0, oc) such that l g(t, x) - g(t, y)l < CJx - yla whenever
lxl, lyl, and t < N. Then the partial derivatives Dijv = a2v1ax;axx are continuous, and
Dv =
JDiiPs(xY)(t - s, y) dy ds. Ef
Proof Suppose for simplicity that i = j. In this case, the right-hand side is
227
8.2 The Inhomo$sneoa. Equ.tlo.
ft
f
(2ns)-az
((xi -Z Y,Z - S
)
`
Jo,J
e- i=-yi2/zsg(t - s,Y) dy ds
=
fo'
-sg(t-.s,B.)]tIS,
ExL(xi-BZ)
S
but this time, however, ExI(xt - Bs )2 - SI = sEoI(Bi)2 - 1I so
(x;-B:)2-S
= 00.
S2
We can overcome this problem if f is Holder continuous at x, because we can write [((x;-s2t2- sg(t
E.
- sB)
S
s
=Exl(x`-B2) S
s(g(t-s,Bj -g(t-s,x)))f
The second expression < Cs-1+a, so its integral from s = 0 to t converges absolutely, and with a little work (6c) follows. (See Friedman (1964), pages 10-12, for more details.) The last detail now is : (6d)
Let g be as in (6c). Then 8v/8t exists, and
x) + at (t, x) = g(t,
f'drf
A-,(x,Y)g(r,Y) dY
o
Proof To take the derivative w.r.t. t, we rewrite v as
v(t,x) = JJ'Ptr(xY)frY)'1Y. 0
Differentiating the left-hand side w.r.t. t gives two terms. Differentiating the upper limit of the integral gives g(t, x). Differentiating the integrand gives
fai--
pt-.(X y)g(r, y) dy dr
0
tr lJ
2d(2n(t
-
2
r))-(d+2)12
exp (
c
+
fj
r
+ IX (2n(t - r))-dl2 12(t
2(t
-) I g(r,Y) dy dr
X-Y
- r)I2) exp ( 2I(t -
2d :tdrr -Exg(r,Bt_r)dr+
'
dr
) 2) g(r,Y) dy dr
2N
PDE's'Ib.t Can Ile Solved by Rwtnlni a Brownimn Motion
In the second integral, we can use the fact that Ex(Ix - Bt_,I2) = C(t - r) to cancel one of the t - is and make the second expression like the first, but even if we do this,
' dr
J0t-r This is the difficulty that we experienced in the proof of (6c), and the remedy is the same : We can save the day if g is locally uniformly Holder continuous in x. For further details, see pages 12-13 of Friedman (1964). As in the last section we can generalize our results to unbounded g's. To see what type of unbounded g's can be allowed, we will restrict our attention to the homogeneous case g(t, x) = f(x). At the bare minimum, we need f(B,) ds < oo,
Ex J0
and if we want (b) to hold, we need to know that if t 10 and x
x, then
f0,"
f(B5) ds -* 0.
Ex
If we strengthen the last result to uniform convergence for x c R", then we get a definition that is essentially due to Kato (1973). A function f is said to be in Kd if (*) lim sup Ex (fot I f(B5) I ds) = 0. t4 o
x
By Fubini's theorem, we can write the above as t m sup J kt(x, y) I .f(y) I dy = 0,
where J(27s)_d/2e_k_Y12/2sds.
kt(x,y) = 0
By considering the asymptotic behavior of kt(x, y) as t -+ 0 and we can cast this condition in a more analytical form as
(**) limsupJ a. o
x
w(Ix-yl).f(y)dy=0, Ix-Y1
where
-log Izl
d>3 d=2
1
d= 1.
IZI
(P(z) _
(d z)
This is Theorem 4.5 of Aizenman and Simon (1982). The details of the proof,
8.3
The Feynman-Koc Formula
229
though simple, are a little tedious, so they are left to the reader. We will have more to say about these spaces at the end of the next section. For the developments there, we will also need the space K,"', which is defined in the obvious way: fcKK°° if for every R < co, fl(lxl
8.3 The Feynman-Kac Formula In this section, we will consider what happens when we add cu to the right-hand side of the equation we considered in Section 8.1, that is, we will study (1)
(a) u,='Au+cu in (0,co) x Rd (b) u is continuous in [0, co) x R' and u(0, x) = f(x).
(2)
If c(x) < 0, then this equation describes heat flow with cooling. As in Section 8.1, the solution u(t, x) gives the temperature at the point x c Rd at time t, but here we do not assume that there is perfect conduction of heat. Instead, we assume that heat at x dissipates at the rate k(x) = -c(x). We will see below that this corresponds to Brownian motion with killing at rate k, that is, the probability that a particle survives until time t is exp(-$ok(B,)ds). The first step in solving (1) is to prove: If u satisfies (a), then S
Ms=u(t-s,Bs)exp( c(B,)dr) \\ o
ff
is a local martingale on [0, t).
Proof Let c, = f o c (Bs) ds. Applying Ito's formula gives that u(t - s, BS) exp(cs) - u(t, Bo)
- ur(t - r, B,) exp(cr) dr +
Jexp(c,)Vu(t - r, B,) dB,
_ f s0
+
Ju(t - r, Br) exp(c,) dCr + 2
JAu(t - r, Br) exp(cr) dr,
0
which proves (2), since -u, + cu +
ZAu = 0 and the second term is a local
martingale.
If c is bounded and u is bounded on [0, t] x Rd and satisfies (a), then M, 0 < s < t, is a bounded martingale, so by an argument we have used in the last two sections lim MS = f(B,) exp(c,) STt
and
00
8 PDE's That ('en Be Solved by Rwol" Brownien Motion
M. = so taking s = 0 gives (3)
Suppose that c is bounded. If there is a solution of (1) that is bounded on [0, t] x Rd, it must be v(t, x) = EX(f(B,) exp(ct)).
As before, it is easy to show (4)
Suppose that c is bounded. If v is smooth, then it satisfies (a) a.e. in (0, oo) x Rd.
Proof The Markov property implies that EE(.f(Bt) exp(c) l.9s) = exp(cs)EB(S)(f(Bf-s) exp(ct-s)),
so if we let v(t, x) = EE(f(B,) exp(c)), then the last equality shows that v(t - s, Bs) is a martingale. If v is smooth, then repeating the calculation in the proof of (2) shows that
v(t - s, B) exp(cs)- v(t, Bo)
=
s
(- v, + cv + Z Av) (t - r, B,) exp (c,) dr + a local martingale.
J0
so again, we conclude that the integral on the right-hand side is a local martingale and, hence, must be - 0, so we have that - vt + cv + 1 Av = 0 a.e. The next step is to give a condition that guarantees that v satisfies (b). As before, we begin by considering what happens when everything is bounded. (5)
If c is bounded and f is bounded and continuous, then v satisfies (b).
Proof If Icl < M, then e_Mt < exp(c) < e", so exp(ct) --*I as t - 0. Since f is bounded, this result implies that E, exp(ct)f(B) - EE.f(Bt) - 0, and so the desired result follows from (5) in Section 8.1. This brings us to the problem of determining when v is smooth enough to be a solution. To solve the problem in this case, we use a trick to reduce our result to the previous case. We observe that exp
(f c (Bs) ds) = 1 + Jc(Bs)exP(,fc(Br)dr)ds fJ
so taking expected values gives that
v(t, x) = I + JEx [c(Bs)exp(jtc(Br)dr)f(Bt)]ds. Conditioning on .mss and using the Markov property, we can write the equation above as
8.3
231
The Feyameo-Kec Formula
v(t, x) = 1 +
JExc(Bs)v(tsBs)ds. 0
The second term on the right-hand side is of the form considered in the list section. If we start with the trivial observation that if c and f are bounded, then v is bounded on [0, t] x Rd, and if we apply (6a) and (6b) from the last section,
we see that v is continuous and the derivatives 8v/8x; are continuous. This implies that, for each N, I v(t, x) - v(t, y)I < CI x - yI whenever IxI, IyI, and t < N. If we assume that c is bounded and locally Holder continuous, then it follows from (6c) and (6d) that we have (6)
Suppose that f is bounded. If c is bounded and locally Holder continuous, then v is smooth and, hence, satisfies (a). As in the last two sections, we can generalize the results above to unbounded
c's. Given the formula above, which expresses v as a volume potential, it is perhaps not too surprising that the appropriate assumption is c E Kd. The key to working in this generality is what Simon (1982) calls Khasmin'skii's lemma: (7)
Let f >_ 0 be a function on Rd with
a = sup Ex (f,f(BS) dsl < 1. J
x
T hen sup E. exp
(f f(BS) dsl < (1 - a)-1.
x
Proof The Markov property and nonnegativity off imply that sup Ex X
f ... j' dsl ...
f(Bsn) < an,
so the desired result follows by noticing that
J...J
Jo...
nl
Jo
and summing on n. From the last result, it should be clear why assuming c c Kd is natural in this context. This condition guarantees that sup E. ( Jt I c (B.) Ids) X
/
o
-0
and, hence, that sup Ex exp x
\ fo, I c (
BS) I
ds) J
- 1.
232
s
PDE'r Tbat Can B. 9olwd by Ruing Brownian Motion
With these two results in hand, we can proceed with developing the theory much as we did in the case of bounded coefficients. Since Simon (1982) has written a lengthy and very readable account of how to develop the theory in this generality, we will content ourselves just to briefly describe a few of the results given in his paper. Part of our motivation for doing this is to establish the connection
between the notation we use and the way in which mathematical physicists write things. To make it easy for the reader to find the results in Simon's paper, we have used his theorem numbers below.
Let Ho = - A and V = - c z
H=Ho+V=-A+V, and define a linear operator e- 1H by setting
(e-`Hf) (x) = E. (exp
(_
, V (BS) ds).f(B,) I .
J
THEOREM B.1.1 Let V- E Kd and V E KK'0c. Then for every t> 0 and p < q < co, e-`H is bounded from L° to L.
Proof Since the proof relies on some things that we have not explained above, we simply sketch the proof and refer the reader to Simon (1982) for details. Let Ile " 111,9 denote the norm of e-`H as a map from L" to L. Step 1: p = oo, q = eo. The Feynman-Kac formula shows that if t is small, then
e 1H A. < Cll.fll., so the semigroup property e-(s+t)H = e-sHe-H imples that lle-t"11.0 m < CeAt,
where A = T -'In C. Step 2: p = 2, q = co. Using the Cauchy-Schwarz inequality in the FeynmanKac formula gives e-tH fl < (e-t(Ho+2V)1)1/2(e-uHol fI2)1/2.
Applying Step 1 to Ho + 2 V gives e-t(Ho+2v)Ih.0 < C'e' `,
and an easy estimate shows that e-tH09lh = sup J(2xt)_2pt(x,y)g(y)dy z
< (2nt)-112II9ll1,
8.3
233
The Feynmra-Kac Formula
so
IIe-tH fhlm < C"IIf112
Step 3: p = 1, q = 2. Since
1,2 =
Ile-1H11
2.o0
Step 4: p = 1, q = oc. By the semigroup property, lie-
W111'. G
Ile-IH12II1,2Ile-1H/2II2..
Step 5: Steps 1 and 4 show that a-` is bounded from L°° to L°° and from L' to L. The result now follows by "duality and interpolation." The next result gives another reason why the spaces Kd are well suited for studying Shrodinger semigroups. PROPOSITION B.1.4
Ile-`"II,,.,, = 1, then VEKd.
If V :!g 0 and
Proof This is an easy consequence of Jensen's inequality and is left as an exercise for the reader. Theorem B.1.1. shows a-`" maps L°° into L. With a little work this can be improved considerably. THEOREM B.3.1
Let V- e Kd, V+ e Kd10'. If f E L°°, then a-tH f is a continuous
function. THEOREM B.7.1
e-nHf(x)
Let V- e Kd, V+ E Kd ' . Then
=
,
where a-"(x, y) is jointly continuous in x, y, and t in the region t > 0. Let V- eKd, V+ e Kd"'. If f e L°°, then, for any t > 0, e-tH f has a distributional gradient in L'10; THEOREM B.3.4
To get more smoothness, one has to assume more boundedness. Suppose for simplicity that d > 2 and, for 0 < a < 2, let Kd be the set of all functions that satisfy :
(i) sup JIX X
< oo if a # 1
1)
Ix -
(ii) limsup r4o
yI-(d-2+a)I f(y)I dy
x
D(x,r)
yI-(d-2+a)I f(y)I dy
= 0 if a = 1.
234
$
PDE'e 111.t ('Mn Be Solved by Running is Brownian Motion
Let a < 2. Let V E Kd, V+ E Kd'°°. Suppose that the restric-
THEOREM 18.3.5
tion of' V to some hounded open set G lies in Kd. If feL°°, then for each t > 0, e- "`feC°(G) = the set of functions whose derivatives of order [a] are Holder continuous of order a - [a].
Remarks: The fact that V is not supposed to be smooth restricts the last
result to a < 2. The reader should also observe that by writing e-` = e-`12 e-`12 and applying Theorem B.1.1, we can conclude that the results above hold if f E L°° is replaced by f E UP1 L°.
Inspired by the work of Feynman (1948), Kac (1949) proved the first version of what is now known as the Feynman-Kac formula. He proved his result in d = 1 for potentials V = - c, which are bounded below, by discretizing time, passing to the limit, and ignoring a few details along the way. Rosenblatt (1951) extended Kac's work to d>- 2 and filled in the missing details (e.g., Note :
Holder continuity is needed if one wants the solution to be C 2). Since that time,
there have been a number of papers extending the result to more general processes and potentials. The results we have mentioned above are only a small sample of what is known. If the reader would like to see more examples of how
probability can be used to study these problems, he should look at McKean (1977), Berthier and Gaveau (1978), and at recent work of Carmona and Simon. Perhaps the best place to begin is Simon's (1982) survey paper.
8.4 The Cameron-Martin Transformation In this section, we will consider what happens when we add b . Vu to the righthand side of the equation considered in Section 8.1, that is, we will study: (1)
(a) u
x Rd and u(0, x) = f(x).
in
Physically, the extra term corresponds to a force field. In this section, we will see that the probabilistic effect is to add an infinitesimal drift b to our Brownian motion. The first step in solving (1) is to prove: (2)
If u satisfies (a), then MS = u(t - s, BS) exp
\Jo (f
S
1
b(B,) dB, - 2 JI1'(8r)I2th)
is a local martingale on [0, t).
Proof/ Let ZS =f o b(B,) dBr - i f o I b(Br) 12 dr. Applying Ito's formula to J (x0 , x 1 , ... , xd+ 1) = U (t - x0 , X1, ... , xd) exp (xd+1)
X°=s,X. =Bsfor1
8.4
The Cameron-Martin Trandormitloo
239
gives:
u(t - s, B) exp(Z9) - u(t, BO) = f - u, (t - r, B,) exp(Z,) dr o
+ J exp(Z,)Vu(t - r, B,) dB, 0
+ f u(t - r, B,) exp(Z,) dZ, 0 s
+2
J0
Au(t - r, B,) exp(Z,) dr
Jsiut - r,B,)exp(Z,)d,
+ i=1('
+ 2 f s u(t - r, B,) exp(Z,) d
To check this result, observe that the first three lines are the terms involving
first derivatives of f. The last three lines are the terms with D;jf where (a)
l
d + 1, respectively. The terms with i or j = 0 vanish, because X° is locally b.v. Applying the associative law and the formula for the covariance of two stochastic integrals to the mess above gives - u, (I - r, B,) exp(Z,) dr + 2 local martingales 0 fs
+ Ju(1 - r,B,)exp(Z.)(-zI b(B,)I2)dr °
Liu(t - r, B,) exp(Z,) dr
+2 0
s
jDtu(t_rB,)exp(Zr)bi(B,)dr
+ i=1
0 s
+
u (t - r, B,) exp (Z,) I b (B,) 12 dr. 0
2
The third and sixth terms cancel, and if u satisfies (a), the sum of the first, fourth, and fifth is 0, proving (2). At this point, the reader can probably anticipate the next step. (3)
Suppose that b is bounded. If there is a solution of (1) that is bounded on [0, t] x Rd, it must be
v(t, x) = E.(f(B,) exp(Z)).
2M
N
PDE's That Con N. Solved by Rundni
Brownian Motion
This time, however, we cannot simply define our problems away, because Xs = f o b(B,) dB, may be unbounded. Let T. = inf{t : IXtI > n}. The exponenis a local martingale. So tial formula implies that exp(ZZ) - exp(Xt i<X>) observing that <X>, z 0 and stopping at s A T. gives
Eexp(ZSAT) = Eexp(Z0) = 1.
Letting n - oc now and using Fatou's lemma shows Eexp(Z5) < 1, that is, if we let Y = exp(Z5), then Y is an Ll bounded martingale. Applying the last result with b replaced by 2b gives 1
>Eexp(2 f Sb(B,)-dB,-4.1 f lb(B,)I2dr) \\
/
2 ,J o
,10
> exp(-sb*)EY2, where b* = suplb(x)I, and we can use the martingale convergence theorem to conclude that if u satisfies (1), then lim M, = f(B) exp(ZZ) stt
and
M. = EE(f(B)eXp(Z,)I.), so taking s = 0 proves (3). There are many other ways to prove (3). Let
Exercise
Xs= 0
and observe that <X>s = J S I b (B,)12 dr < Cs, 0
so it follows from Levy's theorem (see Section 2.11) that
E (sup exp (Xs)) < E (exp ( sup B.) o<s
l
0<s5b't
1
and the right-hand side is finite since we have (see Section 1.5) Po (B* > a) = 2Po (Bt > a)
for a > 0
Note: This proof was originally in the text. I would like to thank Tom Liggett for pointing out the simpler proof used above. As before, it is easy to show the following: (4)
If v is smooth, it satisfies (a) a.e. in (0, oc) x Rd.
Proof The Markov property implies that Ex(.f(B) exp(ZZ)I. ) = exp(ZS)EB(s)(.f(Bt-s)exp(ZZ-s)),
8.4
237
The Cameron-Martin Transformation
so if we let v(t, x) = Ex(f(B,) exp(ZZ)), then the last equality shows that v(t - s, Bs) is a martingale. If v is smooth, then repeating the calculation in the proof of (2) shows that v(t - s, Bs) exp(ZS) - v(t, Bo)
=
fS
0
(- v, + b - Vv + Av) (t - r, B,) exp(Z,) dr + a local martingale. ?
Again, we conclude that the integral on the right-hand side is a local martingale and, hence, must be -0, so we have -v, + b - Vv + zAv = 0 a.e. The next step is to give a condition that guarantees that v satisfies (ii). As before, we begin by considering what happens when everything is bounded. (5)
If b is bounded and f is bounded and continuous, then v satisfies (b).
Proof Let Y = exp(Z,). As t - 0, Y,-+ 1 almost surely, and we have Ex Y* < cc. Since f is bounded, it follows from the dominated convergence theorem that
EE.f(B) exp(Z,) - Exf(B,) - 0, and so the desired result follows from (5) in Section 8.1. This brings us last, but not least, to the problem of determining when v is smooth enough to be a solution. As you might guess by extrapolating from the last three sections, this is a very difficult problem. The reader is invited to think about how he might try to solve this problem. We do not know a very simple way of doing this, so we will put off consideration of this point until Chapter 9, when we will confront the problem in a more general situation. Having skipped smoothness, the last item on our outline is: What happens if b is unbounded? To answer this question and to prepare for developments in Chapter 9, we will look at our solution through the eyes of Cameron and Martin (1949). Let P. be the measure on (C, ') that makes the coordinate maps B,(co) = ws a Brownian motion starting at x, and define a new measure on (C, by setting
Q.(A) = f
dPx
for A e
,
JA A
where Ib(BS)12ds.
2 fo
f o b(B5) dB, is a local martingale with <X> = fo Ib(B,)I2 ds, it follows from the exponential formula that exp(Z,) is a local martingale/Ps. Let T = inf{t : IZ,I > n}. By stopping at T. A t and letting n -+ oo, we see that Since X,
Q. (C) = E,, exp Z, < 1.
In some situations (e.g., b bounded), we will have Q.,(C) = 1. The next result shows that when this occurs, Q,, makes the coordinate maps behave like a Brownian motion plus a drift.
2,3$
B
(6)
If QS(C) = I, then under Qx
PDE's That ('en Be Solved by Running a Brownian Motion
W, = B, - Jb(B5)th 0
is a Brownian motion starting at x.
Proof Let W' be the jth component of W, Our first goal is to prove that W' is a one-dimensional Brownian motion starting at x;. The first step in doing this is to observe that it suffices to show that if for all 0, U = exp(iO W' + 02t/2) is a local martingale under QX, because then EQX(exp(i0(W' + 02t/2))I.) = exp(iOW' + 02s/2), and
exp(-02(t - s)/2).
EQx(exp(i0(W' -
In other words, W' - WS' is independent of J and has a normal distribution with mean 0 and variance t - s. By (2) from Section 2.13, U is a local martingale/Qx if and only if U exp(Z,) is a local martingale/Px. Unscrambling the definitions gives
U exp(Z,) = exp (i0 (B?
- f" b'(BS) ds) + 02 t 1 o
,1
- 12
exp\fotb(B.) dBs
J
fo
where
C, = ioBi + f b(B.,) dB,, 0
and
D, =
- 02t
2
+ i0
b'(BS) ds + 0
1
2
`
I b(BS)I2 ds.
fo
Now, if we let X, = fo b(BS) dB, then
, = t, , = Jbi(Bs)ds ,
and <X >=
' t 0
,
so D, = i
8.4
239
The Cameron-Martin' 1'ranaformatlon
The computation above shows that each component WJ is u local martingale. To complete the proof, we observe that (5) in Section 2.8 implics that <W', W'5, is the same under Q as it was under P,,, so <W`, WJ>, = t1Ut. The desired result follows from the characterization of multidimensional 1rownian motion given in Section 2.12. (6) shows that under Qx, the coordinate maps, which we will now call X,((o) = co, satisfy
X,=W+ f
b(Xs)ds
,I0
where W is a Brownian motion, an equation that we can write formally as (*) dX, = dB, + b(X,) dt
(here, to facilitate comparison with the results above, we have replaced the W (for Wiener process) with the letter B, which we usually use to denote Brownian motion). There is an obvious connection between (*) and (1): (2')
If u satisfies (a) and X satisfies (*), then M, = u(t - s, XS) is a local martingale on [0, t).
Proof Applying Ito's formula gives s
s
-u,(t - r, X,) dr +
u(t - s, Xs) - u(0, Xo) =
Vu(t - r, X,) dX, J0
0 s
+2Y_
D;ju(t-r,X,)d<X`,X'>,.
ii fo
Since <X`, X'>, = 6,,r, it follows that the right-hand side
f
- u, (t - r, X,) dr + a local martingale
o S
d
+Y f i=1
D,u(t - r, X,)b`(X,) dr
0
+ 2 J Lu(t - r, X,) dr, 0
which proves (2') since - u, + b Vu + 'Au = 0. From the discussion above, it should be clear how to approach the problem of solving (*) and (1) when b is only locally bounded. Let b" = bl(jxjs"). Since b" is bounded, we can solve (*) and get a process X" that satisfies the original equation for t < T" = inf{t : IXI > n}. The measures µ" on (C, W) that give rise to X" have the property that if m < n, then it,, and It agree on .F (T.), so we can let n co and construct a process that solves the original equation for t < T. =
240
8
PDE'. flit ('en Be Solved by Running
Browden Motion
lim T". When 17:,, (x) } has positive probability, we say that the process explodes. The next result gives a simple condition that rules out explosion. (7)
If x b(x) < C(1 + lxl2), then the process does not explode starting from any
xeR°. Proof Let b", X", and T" be as above. Let g(x) = 1 + lx12 and, to ease the notation, let Y = X,". Applying Ito's formula ula gives
9(Y)-9(Yo)=Y J2YsdYs+1Y2d
0
= a l ocal martingale +
f 2 Y`b.'(Y) ds + td, 0
and our hypothesis implies that Y J2Ysib(Ys)ds < 2C f g(Y)ds, 0
so applying Ito's formula to exp(- (2C + d)t)g(Y) gives
exp(-(2C + d)t)g(Y) - g(Y0) = a local martingale
+Y f exp(-(2C +d)s)2Ysb;,(Y)ds o
i
+
f
-(2C + d)exp(-(2C + d)s)g(Y) ds
0
+ 2 i f exp(-(2C + d)s)2d
The inequalities above show that the sum of the last three integrals is :50, so S, =- exp(-(2C + d)t)g(Y) is a local supermartingale, and since S, > 0, it is a supermartingale. Applying the optional stopping theorem at time T. A t now shows that
Exexp(-(2C + d)t)9(YT.A) < 9(x), so
PP(T
t)
exp( l
+ n d)t) 9(x),
and the desired result follows immediately.
Remark: (7) implies that if b(x) = lxlav where v 0 is some fixed vector, then the process does not explode if S < 1. We will see in a minute that it does explode if S > 1, so the condition above is sharp. Perhaps the best way to get a feel for the properties of solutions of (*) dX, = dB, + b(X,) dt
8.4
The Cameron-Martin Trwformation
241
is to consider the one-dimensional case. In this case, the analysis is simple because we can find a function cp that makes cp(X,) a local martingale (thiN is the "natural scale"). To see what p to choose, use Ito's formula to conclude thin
w(X)-w(Xo)=fo 4'(Xs)dxs+2 Jco"(X5)ds = a local martingale + fo cp'(XS)b(X9) + (p"(X,) ds, 2 so
if we want qp(X,) to be a local martingale, we must have
gp'(x)b(x) + 2cp"(x) = 0,
that is,
9"(x) = -2b(x)tp'(x) 9'(y) = C exp
(f
y
- 2b (x) dx)
o
cp (z) = B +
exp (
f
Y
- 2b (x) dx dy.
\ o
JoOz
Taking B = 0 and C = 1 in the last expression, we get a function that is very useful in studying the behavior of X. We have used 9 to try to remind you of the
results in Sections 1.7 and 3.1. If it did, you should have no trouble with the following exercises.
Exercise 2 Let T, = inf{t : X, = c}. Then if a < x < b, PX(Tb < T.) = (P(x) - (P(a) (p(b) - cp (a)
Exercise 3 Since 9 is increasing, 9(oo) = limX1. p(x) and cp(- oo) = limsl-. rp(x) exist. Show that X is recurrent (i.e., PX(Ti, < oo) = 1 for all x and y) if and only if tp(oo) = oo and q,(- oo) oo. To see what this means in a concrete case, consider b(x) = Clxla for Ixj > 1, and b(x) = 0 otherwise. In this case, qp(z)=
JZexp(_ fi
so when S > -1, tp(oo) < oo, and when S < -1,
cp(z)-zexp(- f
\
as z -+ oo. In the critical case 6 = -1,
242
N
PDE's 71st ('rn Be Solved by Running s Brownian Motion
cp(z) = 1 +
JexP(_2cloY)dY , 1
sotp(oo)= oo if and only if2C< 1. The last exercise shows that X is recurrent if and only if (p(- oo, oo) _ (- oo, oc). If we think about the results of Section 2.11, this conclusion is obvious. 9(X,) is a time change of Brownian motion run for a random amount of time T. If qp(- oo, oo) # (- oo, oo), then this time must be finite, whereas if gyp(- co, oo) = (- oo, oo), it must be infinite. By looking at the scale function, we can also tell exactly when the process will explode.
Exercise 4 A Special Case of Feller's Test. Let T.,, be the explosion time defined above. Either PX(T. = oo) = 1 or PX(T.. < oo) = 1, depending on whether or not J
0
(1P(x) - (o(-cc))dco(x) = oo =
f
((p(oo) - 9 (x))dcp(x)
o
Solution:
The key to the proof of (7) was the fact that
S, = (1 + XX2)exp(-A,t) > 0 is a supermartingale if A is sufficiently large. To get the optimal result on explosions, we have to replace 1 + x2 by a function that is tailor-made for the process. A natural choice that "gives up nothing" is a function g >- 0 that makes e-rg(X,) a local martingale. To solve this problem, it is convenient to look at things on
the natural scale, that is, let Y, = 9(X,) and find f = g 9-1, for then when we apply Ito's formula, the second term on the right is a local martingale:
e-f(Y)ds +
e-`.f(Ii) -f(Y0)
s
fo
+ Je8f"(}c)d
To evaluate the third term on the right, we observe that
2
Y - Y. = J gp'(XS) dX + bounded variation, 0
so
Combining this result with other observations above, we see that e-`f(Y) _ e-`g(XX) is a local martingale if and only if
zf"(x)gp'(x)2 -f(x) = 0.
8.4
The Cameron-Martin Transformation
243
To solve this equation, we iterate. Let .fo
.f"(x) =
f s d(p(x)
f'd(p(y)f._jy)
0
0
and let 00
f= n=o I.f" It is easy to show that f" < (f1)"/n !, so the series converges and has 1 + f1 <_ f :r'
exp(f1). If the solution f that we construct -oo as we approach either end of cp(- oo, oc), then we are done, because then we can let [a", b"] T tp(- oo, oo), let
T. = inf{t : X, 0 [a", b"] }, and apply the optional stopping theorem at time T" A t to conclude that if x c [a", b"], then
P.(T"
etf(x) .f(a") n f(b")
-0
as n - co (generalizing the inequality in the proof of (7)), so there is no explosion. On the other hand, iff(x) stays bounded as, say, x T 9(oo), we are also done, for then we let t = inf{t : XX = 0}, let 0 < x < oo, and apply the optional stopping theorem at time t A T" to conclude that
1 < g(x) =
1+
=1+ Letting n
< t) < t).
g(oo)EX(e-TAT"; T" g(oo)EX(e-T"; T"
oo, now have
EX(e-T"; T.
< t)1
Ex(e-T. ; T.
< t),
which is a contradiction, unless PX(T.,, < t) > 0, so X explodes in this case.
The last detail now is to relate the behavior of g to the behavior of the integrals above. Going back to natural scale and using the inequality 1 + f1 < f < exp(fl), we see that X does not explode if and only if f1 - oo at both ends of q,(- oo, oo), which is the condition given in the theorem.*
Note: This exercise is from McKean (1969). The result is due to Feller. Now that the one-dimensional case has been discussed in general, the last step is to consider two concrete examples.
Example l The d-Dimensional Bessel Process. In Section 2.10 (see (8) and (9)), we showed that if R, = IB,I, where B, is a d-dimensional Brownian motion, then
Rj-2
r
r
d 0
2
IR51ds= fRs1B5.dB., o
244
8
PDH's 71n1 Can He Solved by Nwainit a Hrownien Motion
and
B;=
Since the drill coefficient here is of the form b(x) = Cx-1, by applying Exercise 3 we see that R, is recurrent if and only if
2C= d - I < 1, that is, d!9 2. The Ornstein-Uhlenbeck Process.
Example 2
b (x) = - ax. In this case there is an amusing way of solving
(*) dX = - aX dt + dB by purely formal calculations: dX _
_ ax +
dt
dB dt
= (_ aeat)X + eat dB dt dt d (eatX) = eat dB dt dt eat dX
eatXX - Xo =
f
eas dBs
0
Xt = e-atX0 + eat
eas dBs. 0
All the calculations above are formal, but it doesn't really matter. Once we know what to guess, it is easy to check that the formula defines an OrnsteinUhlenbeck process
E,,(Xt - X0) = (e-at - 1)x - - axt as t - 0 t
E.(Xt - X0) 2 = ((1 - e-at)x)2 + E.
(e-at
=0(t2)+t+o(t) ast->0.
2
eas dBs 0
so it follows that Xt + f aXs ds is a local martingale and 0
<X>t = t.
The representation given above for the Ornstein-Uhlenbeck process is nice because it makes certain facts about the process obvious.
8.4
The Cameron-Martin Transformation
245
(a) If Xo = x, then XX has a normal distribution with mean a-°'x and vnrinnce (1 - e-z"')/2a, and hence (b) As t -+ oo, X, converges in distribution to normal with mean 0 and variance 1/2a.
Note: The discussion above of Example 2 is based on Section 16.1 of Brciman (1968). Although the first steps in developing the theory in this section were taken by Cameron and Martin (1944b), the formulation given above is due to Girsanov (1960). In our development above, we have basically followed Meyer (1976), but we have also incorporated some material from Friedman (1975) and Stroock and Varadhan (1979).
B
Elliptic Equations In the next three sections of this chapter, we show how Brownian motion can be used to construct (classical) solutions of the following equations:
0 ='Au
0=2Au+g
0='Au+cu in an open set G subject to the following boundary condition: u is continuous
inGandu=fonBG. The solutions to these equations are (under suitable assumptions) given by EF.f(BT)
E.(f(BT) + J g (BS) ds \\ EX (.f(B,) exp
o
\ fo
c(B.)
dsll
where r = inf{t : B, 0 G }. Comparing the solutions above to the solutions given
for the equations in Part A of this chapter shows that (except for a minor modification of the second solution) all we have done is replace t by T. From this viewpoint, the changes made above may seem ad hoc, but if we reverse our perspective and rewrite the first solutions in terms of space-time Brownian motions h, = (t - s, BS) run until time r = inf{s : B9 0 (0, oc) x Ra}, we see that the recipes are exactly the same.
24
N
PDE's 71n1('a.11r Solved by Running s Brownian Motion
8.5 The Dirichlet Problem In this section, we will consider the most classical form of the Dirichlet problem, that is, we will study: (1)
(a) Au=0in(1
_ (b) u is continuous on G and u = f on G. If we let h(t, x) = u(x), then h satisfies the heat equation with h(0, x) = u(x), that is, u is an equilibrium distribution for the heat equation in which the 8G is held at a fixed temperature f (which may vary from point to point). As in the first half of this chapter, the first step in solving (1) is to show:
(2)
Let T = inf{t : Bt 0 G }. If u satisfies (a), then M, = u(B,) is a local martingale on [0,,r).
Proof Applying Ito's formula gives ('
u(B,) - u(B0) = f Vu(BS) . dB., + 1 J Au(BS) ds, 2
o
which proves (2), since Au =_ 0 and the first term is a local martingale on [0, T). If u is bounded and satisfies (a), then MS, 0 < s < T, is a bounded local mar-
tingale, so M., converges to a limit as s T T. If G is bounded, T < oo a.s. (see Exercise 1 in Section 1.7), so if u satisfies (ii), the limit must be f(B,) and since Ms, S < T, is bounded, it follows that MM = E.(f(B,) I A).
Taking s = 0 in the last equation gives (3)
Suppose that G is bounded. If there is a solution of (1) that is bounded, it must be
v(x) = Exf(Bz) As in the first half of the chapter, it is easy to show: (4)
Suppose that G and f are bounded. If v is smooth, then it satisfies (i).
Proof The Markov property implies that on {T > s} EE(f(BjI.9;) = v(Bs) Since the left-hand side is a local martingale on [0, T), it follows that v(B.,) is also.
If v is smooth, then repeating the calculation in the proof of (2) shows that, for S E [0, T),
v(B.) - v(B0) = 2
Av(B,) dr + a local martingale, fo
so it follows that the integral on the right-hand side is a local martingale and, hence, must be 0, so we have Av = 0 in D.
Up to this point, everything has been the same as that in Section 8.1.
8.5
247
The Dirichlet Problem
Differences appear when we consider the boundary condition (b), since it is no
longer sufficient for f to be bounded and continuous. The open set (i Must satisfy a regularity condition. A point y c- 8G is said to be a regular point if Py(T = 0) = 1. (5)
Let G be any open set. Suppose that f is bounded and continuous and y is it v(y). regular point of 8G. If x e G and x -+ y, then
Proof The first step is to show (5a)
If t > 0, then x - Px(T < t) is lower semicontinuous. Proof Px(XSEG` for some se(E,t]) = f pE(x, y)PP(T < t - s). Since y - (T < t - s) is bounded and measurable and
p.(x,y) = it follows from the dominated convergence theorem that (27rE)-12e-Is-yI2/2E
x -. PX (XS e G' for some s E (c, t] )
is continuous for each E > 0. Letting E 10 shows that x --> Px(T < t) is an increas-
ing limit of continuous functions and, hence, by a standard argument, that if
x - y, then lim inf Ps (T < t) > Py(T < t).
If y is regular for G and t > 0, then Py(T < t) = 1, so it follows from (5a) that if x -+ y, then lim inf P,t.(T <- t) >- 1. n-OD
With this established, it is easy to complete the proof. Since f is bounded and continuous, it suffices to show (5b)
If y is regular for G and x,, --> y, then for all S > 0 Ps,, (BL E D(y, S)) -+ 1.
Proof Let E > 0 and pick t so small that PO
(sup
\0 <s
IBSI > 2J
<
Since Px,,(T < t) - 1 as x - y, it follows from the choices above that lim inf Px.(Bt a D(y, S)) > lim inf Px. (T < t, sup I Bt - x
\
o<Ss(
2)
> liminfP, (T < t) - PO sup IBtI >
n-
(O<S!5,
>1-E. Since r was arbitrary, this result proves (5b) and, hence, (5).
Sl 2
24$
x
PDE's That Can He Solved by Ruaaini a Brownies Motion
(5) shows that if G is regular (i.e., every point of 8G is regular), then v(x) = Exf(BT) will satisfy the boundary condition (b) for any bounded continuous f. It is easy to see that there is a converse to this result. Exercise 1 Let G be an open set and let y e 8G have Py(r = 0) < 1 (and, hence, by Blumenthal's zero-one law, Py(r = 0) = 0). Let f be a continuous function on 8G withf(y) = 1 andf(z) < 1 for all other z e 8G. Show that there is a sequence of points x" - y such that lim info-.,) v(xn) < 1.
Hint: To find these points, start a Brownian motion at y and run it until it exits D(y, 1/n). From the discussion above, we see that for v to satisfy (b) it is sufficient (and almost necessary) that each point of 8G be a regular point. This situation raises two questions :
Do irregular points exist? What are sufficient conditions for a point to be regular? In order to answer the first question we will give two examples.
Let d> 2 and let G = D - {0}, where D = {x : Ixj < 1}. If we let To = inf{t > 0: B, = 01, then PO(To = oo) = 1, so 0 is not a regular point of OR Example 1 (trivial)
Example 2 Lebesgue's Thorn. Let d>3 3 and let G= (-1, 1)d - U-n=1 [2-", 2-n-1] x [-an, (See Figure 8.1 for a look at G (1 {x : x3 = ... = Xd = 0}. Younger readers will notice that G is a cubistic cousin of Pac-Man with infinitely many very small teeth.) I claim that if a" j 0 sufficiently fast, then 0 is an]d-1
not a regular point of 8G. To prove this result, we observe that since threedimensional Brownian motion is transient and P0((B, , Bi) = (0, 0) for some t > 0) = 1, then with probability 1, a Brownian motion B, starting at 0 will not hit In = {x : x1 e [2-", 2-n-1], X2 = X3 = = Xd = 01, and furthermore, for a.e. w the distance between {B.: 0 < s < oo } and In is positive. From the last observation, it follows immediately that if we let Tn = inf{t : B, c- [2-", 2-n-1] x
[a.,an]d-1} and pick an small enough, then P0(Tn < oo) < 3-". Now Y' l 3-"= 3-1(3/2) = 1/2, so if we let i =inf{t > 0 : BOG} and u =inf{t > 0 : B, (-1, 1)d}, then we have PO(T< a) :5
PO
n=1
(7'n
so
Po(i > 0) > PO(T = o-) >
2
and 0 is an irregular point. The last two examples show that if G` is too small near y, then y may be
8.5
The Dirlchlet Problem
249
Figure 8.1
irregular. The next result shows that if G` is not too small near y, then y is regular. (5c)
Poincare's Cone Condition. If there is a cone V having vertex y such that v fl D (y, r) c G `, then y is a regular point.
Proof The first thing we have to do is explain what we mean by a cone. In Section 4.1, we defined Va = {(x,y)c- H: I x - 01 < ay,y,
< 1},
an object that might be called a cone with opening a, vertex (0, 0), and direction ed = (0, ... , 0, 1). Generalizing this definition, we define a cone with opening a, vertex z 1, and direction z2 as follows :
Va(zl,z2)={z:z=z1+y(z2+w)where w 122 andllwll
P-,(B,EV(z1,22))=E>0, where a is a constant that depends only on a, so an easy argument shows that if VV(z, z') fl D(z, r) c G` for some r > 0, then
2011
8
PDE'a That Can Be Solved by Running a Brownian Motion
lim inf PZ(B, E G`) >_ e.
40
Combining the last conclusion with the trivial inequality P.(-r < t) > PZ(B, e G`) shows that
PZ(T=0)=limPZ(T
and it follows from Blumenthal's zero-one law that PZ(r = 0) = 1.
The last result, called Poincare's cone condition, is sufficient for most examples (e.g., if G is a region with a smooth boundary). The ultimate result on
regularity is a necessary and sufficient condition due to Wiener (1924). To describe Wiener's test, we would have to define and explain the notion of capacity, so we will content ourselves to state what Wiener's test says about Example 2 above and refer the reader to Ito and McKean (1964), page 259, or Port and Stone (1978), page 68, for details. (5d)
In d = 3, Po(T = 0) = 0 if and only if - oo < Y log (2"a") < oo. n=1
In d > 4, Po(T = 0) = 0 if and only if Y (2"an)d- 3 < oo. n=1
In contrast, Poincare's cone condition implies that Po(T = 0) = 1 if lim inf 2"a" > 0. n_ao
The last result completes our discussion of the boundary condition, so we now turn our attention to determining when v is smooth. As in Section 8.1, this is true under minimal assumptions on f. (6)
Let G be any open set. If f is bounded, then v is smooth and, hence, satisfies (a).
Proof Let x e G and pick 6 > 0 so that D(x, 6) c G. If we let a = inf {t : B, D(x, 6)}, then the strong Markov property implies that (*) v(x) = EE.f(B,) = EX[EB(Q)f(Bt)] v(Y) dn(Y), D(x,a)
where it is surface measure on D(x, 6) normalized to be a probability measure, so it follows from (7) in Section 2.10 that v c- C°°. As in the last four sections, our last topic is to discuss what happens when something becomes unbounded. This time we will focus on G and ignore f. By repeating the arguments above, we can easily show the following:
8.6
(7a)
Polison's Equation
251
Suppose that f is bounded and continuous and that each point of 8G is regular. If for all xeG, Px(T < oo) = 1, then v satisfies (1) and is the unique solution. Conversely, we have
(7b)
Suppose that f is bounded and continuous and that each point of 8G is regular. If for some x e G, PX(T < oc) < 1, then the solution of (1) is not unique. Proof of (7b) Since h(x) = PX(T = oo) has the averaging property given in (*), it is C°° and has Ah = 0. Since each point of 8G is regular, P,(T = oo) < Px(T > 1)
- 0 as x - y e 8G. The last two observations show that h is a solution of (1) with f - 0, proving (7b). By working a little harder, one can show that adding aPP(T = oo) is the only way to produce new bounded solutions. (7c)
Suppose that f is bounded and continuous. If u is bounded and satisfies (1), then there is a constant C such that u(x) = EXf(BT) + CPP(T = ao).
Proof See Port and Stone (1978), Theorem 4.2.12.
8.6 Poisson's Equation In this section, we will see what happens when we add a function of x to the equation considered in the last section, that is, we will study (1)
(a) -Au = -gin G
_
(b) u is continuous in G and u = 0 on 8G.
If G = R" and the boundary condition is ignored, then any solution of (1) is called a potential of the charge distribution, because (if the units are chosen correctly) the gradient of the solution gives the force field associated with electrical charges distributed according to g. As in the first five sections of this chapter, the first step in solving (1) is to show (2)
Let T = inf{t : B, 0 G }. If u satisfies (a), then
M= u(B) + J(Bs)ds 0
is a local martingale on [0, T).
Proof Applying Ito's formula as we did in the last section gives u(B,) - u(B0) = fo Vu(BS) dBs + J Du(BS) ds, 2
0
252
8
PDB's not ('rn N. Solved by Rannin a Brownlan Motion
N and the first term on the right-hand side is a local martinpalc on 0, r). If G is hounded, then /s.,r (x) for all xeG (see Exercise 1 in Section 1.7), so if g is hounded and u is hounded and satisfies (a), then for s < T which proven (2). sinec JtAu -
5 1fu11z. + tIIgII.,
so M,, N < T, is a uniformly integrable martingale, and if u satisfies (b), then lim Al, = fo g(B,) dt
.fr
M,
Ex\Jog(B,)dtIFs//
.
Taking s = 0 now gives (3)
Suppose that G and g are bounded. If there is a solution of (1) that is bounded, it must be
v(x) = E. (fo g(B,) dt.
/
Again, it is easy to show (4)
Suppose that G and g are bounded. If v is smooth, then it satisfies (a) a.e. in G.
Proof The Markov property implies that on {T > s}, Ex\Jog(B,)dtl#s/
=
J g(B)di + En(s)(J(Bf)dt).
Since the left-hand side is a local martingale on [0, T), it follows that g(B,)dt + v(B.,)
is also. If v is smooth, then repeating the calculation in the proof of (2) shows that for s e [0, T), s
s
v(B,) - v(B0) + f g(B,) dr = f (ZOu + g) (B,) dr + a local martingale, 0
0
so again, we conclude that the integral on the right-hand side is a local martingale and, hence, must be - 0, so we have 'Au + g = 0 a.e. in D. After the discussion in the last section, the conditions needed to guarantee that the boundary conditions hold should come as no surprise. (5)
Suppose that G and g are bounded. Let y be a regular point of G. If x,, e G and y, then 0.
x
8.6
253
PolMon'. Equation
Proof We begin by observing: (i) In the last section we showed that if r - 0. P,,.(r > E) -, 0, and (ii) if G is bounded, then we have (see Exercise I of Section 1.7) C = sup Exr < oo. X
Combining the last two observations with the Markov property shows that for any r; > 0, Iv(x0I < EIIgII. +
E),
which proves (5). Last, but not least, we come to the question of smoothness. We begin with the case d > 3, because in this case,
dt < oo,
w(x) = E. f Joo00
so the strong Markov property implies that
w(x) = Ex f
dt + Exw(BL),
o
and we have and
(*) v(x) = w(x) - Exw(BT).
The last equation, allows us to verify that v is smooth by proving that w is, a task that is made simple by the fact that w(x) = c J I x - ylz-dg(Y) dy.
The first derivative is easy: (6a)
If g is bounded and has compact support, then w is C'. Proof As before, we will content ourselves to show that the expression we get by differentiating under the integral sign converges and leave it to the reader to apply Exercise 1 of Section 1.10 to make the argument rigorous.
Ix so
(x. _
ylz-d
C
\(z y)zJ1
/Z
2 - d,-Yi)z D,Ix_yIz-d=()((x 2
and we have
Diw(x) = c J (2 - d)iX`_ yig(Y)dY,
2(x,-Yi),
LOA
8
PD6's Thai ('en He Solved by Ruudni a Brownian Motion
the integral on the right-hand side being convergent, since Yld g(.v) I dy <- 11911.
J
lx
dyld-1
< 00.
L>O)
As in Section 8.2, trouble starts when we consider second derivatives. If
i#j, then ylz-d = (2 - d)(-d)lx - YI-d-2(x1 - yi)(x, - Y) Dijl x In this case, the estimate used above leads to JDiilx
-y12_dl < Ix- yl-d,
which is (just barely) not locally integrable. As in Section 8.2, if g is Holder continuous of order a, we can get an extra Ix - yla to save the day. The details are tedious, so we will content ourselves to state the result : (6b)
If g is Holder continuous, then w is C2.
The reader can find a proof either in Port and Stone (1978), pages 116-117, or in Gilbarg and Trudinger (1977), pages 53-55. Combining (*) with (6a) and (6b) gives (6)
Suppose that G is bounded. If g is Holder continuous, then v is smooth and hence satisfies (a).
Proof (6b) implies that w is CZ. Since w is bounded, it follows from (6) in the last section that x -p Esw(BT) is C. The last result settles the question of smoothness in d > 3. To settle the question in d = 1 and d = 2, we need to find a substitute for (*). To do this, we let
w(x) = JG(xY)g(Y)dY where G is the potential kernel defined in Section 1.8, that is, I
n
G(x,y) _
log(lx - yl) d=2
- llx - yl
d = 1.
G was defined as
p, (x, y) - a, dt 0 foo
where the a, were chosen to make the integral converge, so if $ gdx = 0, we see that
T0
J G(x,y)9(Y)dy = li m Ex
9(Bt)dt.
N.7
255
The Schrlidinger Equation
Using this interpretation of w, we can easily show that (*) holds, so tlgnin our problem is reduced to proving that w is C 2, which is a problem in calculuN. (nice all the computations are done, we find that (6) holds in d:5 2 and that in d I ,
it is sufficient to assume that g is continuous. The reader can find dctuils in either of the sources given above. On the basis of what we have done in the last five sections, our next step should be to consider what happens when something becomes unbounded. For the sake of variety, however, we will not do this, but instead, we will show how (1) can be used to study Brownian motion. Example 1 Let d = 1, G = (-1, 1), and g = 1. In this case, formulas (3) through (6) imply that v(x) = EX2 is the unique solution of (1)
(a) ?u"(x) _ -1 in (- 1, 1) (b) u is continuous on [-1, 1] and u(- 1) = u(1) = 0, so u(x) = 1 - x2. Once you see the one-dimensional case, it is easy to do the general case. Exercise 2
Let d > 2, D = {x: I x < 11, T = inf {t : B, 0 D}. Then
Exr = d(1 - Ix12). Remark: This result can also be proved by observing that IB,I2 - dt is a martingale, so Ix12 = 1 - dExr.
8.7 The Schrodinger Equation In this section, we will consider what happens when we add cu to the left-hand side of the equation considered in Section 8.5, that is, we will study: (1)
(a) 2'Au+cu=0inG (b) u is continuous in G and u = f on aG. We will explain the physical significance of this equation in the next section. For the moment, you should consider it simply as the inevitable next step in the progression established in Sections 8.1, 8.2, 8.3, 8.5, and 8.6. As in the first six sections of this chapter, the first step in solving (1) is to show
(2)
Let i = inf{t : Bt G }. If u satisfies (a) then
Mt = u(B,) exp (I c(B) ds) ff
is a local martingale on [0, r).
26
8
PDE'e That ('em Be Solved by fling a Brownian Motion
Proof Let c, =
c(B,)dLs. Applying Ito's formula gives
u(B,) exp(c,) - u(B0) =
fr
exp(c,)Vu(B5) dB, +
u(BS) exp(cs) dos o
0 e
+
Au(B,) exp(c,) ds, 0
2
which proves (2), since dc, = c(B,) ds, IAu + cu = 0, and the first term on the right-hand side is a local martingale on [0, r). At this point, the reader might expect that the next step, as it has been six times before, is to assume that everything is bounded and conclude that if there is a solution of (1) that is bounded, it must be v(x) = EE(.f(Bt) exp(cz))
We will not do this, however, because the following simple example shows that this result is false. Example 1 Let d = 1, G = (-n/2, n/2), and c = 1/2. If u(x) = cos x, then u' (x) = - sin x and u"(x) _ - cos x, so z u" + cu = 0 and u = 0 on 0G. But
there is obviously another solution: u = 0. We will see below that the trouble with the last example is that c = 1/2 is too large or, to be precise, if we let t
w(x) = Exexp(
c(B,)ds ),
then w = oo. The rest of this section is devoted to showing that if w * oo, then "everything is fine." The development will require several stages. The first step is to show (3a)
If w # oc and G is connected, then w(x) < oo for all x e G.
Proof Let c* = suplc(x)I. By Exercise 1 in Section 1.7, we can pick ro so small that if T, = inf{t : JB, - Bol > r} and r:5 ro, then Exexp(c*T,) < 2 for all xeG. If D(x, r) c G, then the strong Markov property implies that w(x) = E.[exp(c(T,))w(B(T,))] < Ex[exp(c*T,)w(('B(T,))]
= EX[exp(c*T)] J
w(y)dir(y), 8D(z,r)
since the exit time T, and location are independent (here TC is surface measure on D(x, r) normalized to be a probability measure). If S < ro and D(x, S) C G, multiplying the last equality by rd-1 and integrating from 0 to b gives
w(x) < 2 S
f
w(z) dz
D(z,6)
where C is a constant (that depends only on d).
8.7
257
The Schrtldinger Equation
Repeating the argument above and using c(T,) > -c*T,, givca it lower bound of w(x) > 2-'-C d
f",(X,6) w(Y) dy. ,a)
Combining the last two bounds gives: (3b)
Let S < ro. If D(x, 2S) c G and y e D(x, S), then w(x) > 2-(d+2)w(Y)
Proof w(x) > 2-1(26)d fD(x,26) w(z) dz
2-1 (26) d
w(z) dz
J d
>2 1( S)d 2 1Cw(Y)_ 2-(d+2)w(y). From (3b), we see that if w(x) < co, 26 < ro, and D(x, 26) < G, then w < oo
on D(x, S). From this result, it follows that Go = {x: w(x) < oo} is an open subset of G. It is easy to see that Go is also closed (if x - y e G and D(y, 36) c G, then for n sufficiently large, D(x,,, 26) c G and y e D(x,,, S), so w(y) < oc). From the last results, it follows (if G is connected) that Go = G, so (3a) holds. With (3a) established, we are ready to prove our uniqueness result. (3)
Suppose that f and c are bounded and that w # oc. If there is a solution of (1) that is bounded, it must be
v(x) = E.(f(B,) exp(ct)). Proof If u satisfies (a), then (2) implies that X3 = u(BS,,,)exp(cSA,) is a local martingale on [0, T). If u, f, and c are bounded and u satisfies (b), then letting s T T gives
u(x) = Ex(f(BT) exp(cT) ; T < t) + Ex(u(B,) exp(c,); T > t).
Since f is bounded and w(x) = Ex exp(ct) < oo, the dominated convergence theorem implies that as t --> co, the first term converges to
Ex(f(Bj To show that the second term -0, we observe that EX[u(B,) exp(c,); T > t] = Ex[Ex(u(B,)
T > t]
= Ex[u(B,) eXp(c,)w(B,); T > t]
and use the trivial inequality
w(x) > exp(-c*)PP(T < 1)
2
8
PDE'u 7ba1 Van Be Solvod by Rrrta/ Brownian Motion
to conclude that inf w(x) = is > 0. XEo
Replacing w(B,) by e,
EE[u(B,) exp(c,); T > t] < e-'EE[u(B,) exp(c.); r > t] < E 1IIuIImEx[exp(ci); T > t] - 0 as t
oo, since w(x) = E., exp(cT) < oo.
This completes our consideration of uniqueness. The next stage in our program, fortunately, is as easy as it always has been. (4)
Suppose that f and c are bounded and that w * oo. If v is smooth, then it satisfies (a) a.e. in G.
Proof The Markov property implies that on {T > s}, Ex(exp(c)f(BT)IS') = exp(cs)Ea(s)(eXp(c,)f(BT)) Since the left-hand side is a local martingale on [0, T), it follows that exp(c,)v(B,)
is also. If v is smooth, then repeating the calculation in the proof of (2) shows that for s e [0, T),
v(B) exp(c,) - v(B0) =
f(f& + cv) (B,) exp(c,) dr + a local martingale,
so it follows that the integral on the right-hand side is a local martingale and, hence, must be = 0, so we have ZAv + cv = 0 a.e. in G. Having proved (4), the next step is to consider the boundary condition. As in the last two sections, we need the boundary to be regular. (5)
Suppose that f and c are bounded and that w # oo. If f is continuous, then v satisfies (b) at each regular point of 8G. Proof Let y be a regular point of 8G. We showed in (5a) and (5b) of Section 8.5 that if S > 0 and x - y, then Px,,(r < S) -> l and Px,,(BT a D(y, S)) - 1, so if cl < M, then Psn(e-"ra < exp(c3) < e"i") - 1 and, since f is bounded and continuous,
EXJexp(c)f(B.); T < S) -+f(x). To control the contribution from the rest of the space, we observe that
Exjexp(ct)f(Bj; T > S) < eMall f < eMaIIfIImjIwII.P..(r
r > 8) a)
0.
This brings us finally to the problem of determining when v is smooth enough to be a solution. To solve the problem in this case, we use the same
8.7
29
The Schr6dinger Equation
trick used in Section 8.3 to reduce our result to the previous case. We observe that exp
(f
T
T
c(BS) dsl
=1+
J
o
T
c(B) exp (
\
,)o
c(B,) dr) ds, s
/
so multiplying by f(BT) and taking expected values gives
v(x) = 1 +
Ex
(c(Bs)exP(f
c(B,) dr)1(S<)) ds.
Conditioning on .'F5 and using the Markov property, we can write the above as
v(x) = 1 + f Ex(c(BS)w(BS); r > s) A 0
The second term on the right-hand side is of the form considered in the last section, so if c and f are bounded and w * oo, then v is bounded, so it follows from results in the last section that v is C 1. If c is Holder continuous, then the right-hand side is Holder continuous, and we can use (6) from the last section to conclude (6)
Suppose that f and c are bounded and that w * oo. If c is Holder continuous, then v is smooth and, hence, satisfies (a). Combining (4) through (6), we see that if, in addition to the conditions in (6), we assume that f is continuous, then v satisfies (1). Just as in the last section, this fact can be used to study Brownian motion. Example 2 Let d = 1, G = (-1, 1), c = - fl < 0, and f = 1. In this case,
v(x) = Exexp(-flr) < 1, so v is the unique solution of (1), and we can find v by "guessing and verifying." Let u(x) = acosh(bx) and recall that cosh x =
ex +2 ex sinh x = ex _ ex
cosh'x = sinh x sinh'x = cosh x. so we have
2u" - flu =
z
a2 - #a cosh(bx).
From the last equation, we see that u(x) = a cosh(bx) satisfies (a) if and only if
b2/2 = fi, that is, b = (, so picking a to satisfy the boundary condition, we find that
Ex exp(-fir) =
cosh(x 2fl)
cosh()
260
$
PDE's That Cam He Solved by Rua.M* a Nrownian Motion
When x = 0, the expression reduces to
E0exp(-/1r) = cosh(.,/2/1) '.
The formula above is valid only for fi > 0, but if we let /1= -a, a > 0, then cosh(
2a) =
e` 2- + e-"1-2' 2
=cos( 2a).
The resulting expression, Eo exp(at) = cos( 2a)-1,
makes sense for a < n2/8 and Too as a T n2/8. We leave it to the reader to prove
that the formula above is correct. We will return to this example in the next section.
If you know something about Bessel functions, you can extend the last result to higher dimensions.
Example 3 Let d>2, 2, G= {z: Izl < 11, and c=- -fl < 0. Again, v(x) = Ex exp(- fir) < 1, so v is the unique solution of (1). This time, however, it is not so trivial to guess the answer, so we just state the result: v(x) = CIXII-112Id/2
1(
where I, is the modified Bessel function m
m2
(Zv+ 2m
/(m!I'(v + m + 1))
M=0
(see Ciesielski and Taylor (1962) or Knight (1981), pages 88-89, for details). It is one of the great mysteries of life that the distribution of z starting from 0 is the same as the total amount of time a (d + 2)-dimensional Brownian motion spends in {z : I z I > 1 } (which can also be computed using the methods of this section). For more on this phenomenon, see Getoor and Sharpe (1979).
tis
Example 4 G = Rd, c(x) = -a - flk(x), where a, fl e(0, oc), and k(x) > 0. Since G is unbounded, this example is, strictly speaking, not covered by the results above. However,
E. (exp (- f f 'k (Bs) ds)) < 1 o
JJJ
and we have supposed that a > 0, so v(x) = fo dt e-'Ex
(exp (_fljk(Bs)ds)) f
nicely convergent, and it is not hard to show (details are left to the reader) that v is the bounded solution of (1), the boundary condition (b) being regarded as vacuous.
8.7
'the SchrtMtnaer Iqustlon
261
The most famous instance of this solution occurs when d = 1 and
k(x) =
I
t0
x>0 x<0
(which again does not satisfy our hypothesis). In this case, Kac (1951) showed that a(a + A
v(0) = l/
so inverting the Laplace transform, e-a
a(a + )
o
1
`
7r
o
e -Ps
s(t - s)
ds dt,
and observing that under PO the distribution of t-' l {s e [0, t] : BS > 0} I is independent of t, we get Levy's arcsine law: Po (I Is c [0, t] : BS > 0} I < Bt) =
1
n
dr
r(1 - r)
= 2 aresin(O). it
The reader should note that t-' l {s E [0, t] : Bs > 0} I
1/2
as t oc. In fact, the distribution of this quantity is independent of t, and its density has a minimum at t = 1/2! Examples where (1) can be solved explicitly are rare. A second famous example, due to Cameron and Martin (1944b), is k(x) = x2. In this case, Eo (exp
(- f
f (B5)2 ds)) =
(sech((2p)1/2t))1i2.
ff
The reader is invited to try to derive this equation. The proof given on pages 10-11 of Kac (1949) is a beautiful example of Kac's computational ability. Up to now, we have focused our attention on the question of the existence of solutions to (1). The probabilistic formula for the solution can also be used to study properties of the solution. Perhaps the most basic result is (7)
Harnack's Inequality. Let u > 0 and satisfy 'Au + cu = 0 in D = {x : I xI < 1}. Then for any r < 1, there is a constant C (depending only on r, c) such that if x, y e D (0, r), then
u(x) < Cu(y).
Proof Pick ro so small that if Tr = inf { t : I B, - Bo I > r} and r < ro, then Exexp(c*Tr) < 2, where c* = supIc(x)I. Repeating the first computation in the proof of (3a), we see that if S < ro and D(x, S) c G, then
u(x) < 2 b° f D(x,S)
u(z) dz
762
8
PDE's 9181 Can Be Solved by Rpmly u Brownian Motion
Figure 8.2
and
u(x) > 2-'-C J
u(y) dy.
,(x,6) Therefore, since we have assumed that u >- 0, we can repeat the proof of (3b) to conclude that if S < ro, D(x, 26) c G, and y e D(x, S), then (*) u(x) 2-(d+2)u(y) The desired result now follows from a simple covering argument. Fix r < 1
and pick b < ro A (1 - r)/2. Given x, y e D(0, 1 - r), there is a sequence of points xo = x, x1, ... , xm = y (with m < 3/6), such that for all i, we have Ixi - xi_1I < 2(5/3 and D(xi, 26) c D (see Figure 8.2). It follows from (*) that the inequality holds with C = 2(a+2)316
Remarks: (a) The constant in the proof above grows like exp(C(1 - r)-1) ; by working harder, you can get (1 - r)-,' (exercise). (b) The result in (7) is true if D is replaced by any connected open set G and D(0, r) is replaced by K, a compact subset of G. The details of the covering argument are much more complicated in this case. See Chung (1982), Exercise 3 on page 205, for a proof, or spend an hour or two and find your own. (c) The result above is also true if we assume only that c e Kd°` (the space defined in Section 8.2). The original proof, due to Aizenman and Simon (1982), was given in this generality. The fact that c can be unbounded causes quite a bit of trouble, but by following the outline of the proof of Theorem B. 1.1 given in Section 8.3 and using a clever time reversal argument as a substitute for the
8.8
EIReev.l.m of A + c
263
self-adjointness used in Step 3, they succeed in proving the key eslimule: 11' S < 60, then u(x) < C
u(y)dn(y) Jan(z, a)
(their (1.13)), and once this result is established, things follow pretty much as before. The reader can find the details clearly explained in their paper. Exercise 1
Use the Poisson integral representation ((2) of Section 3.3) to show
that ifu>0andAu=0inD,then foranyr<1,wehave forallx,ycD(0,r) that u(x) <
1-r
C1 +r)d
u(Y)
Hint: The worst case is u(z) = (1 - zl2)/z - 1". Problem Let r = inf{t : B, D} and suppose that EZ exp(cL) < oo. If we let gy(z) = EZ(eXp(ct)IB= - Y),
then any solution of (1) satisfies
jY((Y)kY(z)tht(Y).
u(z) = n
If we had good estimates on gy(z), then we could give a proof of (7) that is similar to the proof of Harnack's inequality which is given in Exercise 1, but I do not know how to do this. Note: The applications to Brownian motion are based in part on Section 2.6 of Ito and McKean (1964), where the reader can find more details and other applications. The proofs of (3a), (3b), and Harnack's inequality follow Chung (1982), which is, in turn, based on his previous work with several coauthors. These results were also discovered independently by Aizenman and Simon (1982), who proved their results for c e K,"'. Harnack's inequality is just one of several properties of solutions of Schrodinger's equation that can be studied using probabilistic methods. A good place to start learning about these results is Carmona and Simon (1981). They use probabilistic methods to study the exponential decay of Schrodinger eigenfunctions, and they give numerous references to earlier work on related topics.
8.8 Eigenvalues of A + c In this section, we will break the pattern set down in the first seven sections of this chapter and study a new type of problem:
N4
8
(I)
(a) }Au4 ru - AuinG
PDR's flit Con Be Nolvod by Rtanlty w Brownlnn Motion
(b) u is continuous in G and it
0 on 0G.
A function is that satisfies (I) is said to be an eigenfunction of ZA + c (with Dirichict boundary conditions), and 1 is the corresponding eigenvalue. These functions are of interest because they correspond to the pure tones of a drum made in the shape of G. More generally, they are stationary states of the Schrodingcr equation
iu,=IAu+cu. A good way of getting a feel for (1) is to consider a simple example. Example 1
Let D = (0, 1), c = 0. In this case, if we let u(x) = sin(nnx), then z
z
2 u"(x) = - (n2) sin(nnx) _ - (n2) u(x),
2 , -2n , -92 z
so (a) has solutions for A = -
2
2
,
.... The next result, which
is well known (see Courant and Hilbert (1953), vol. I), shows that this example is typical.
(A) If c is bounded and smooth enough (e.g., Holder continuous), then there is an infinite sequence of (real) eigenvalues
00>10>Al>12> 13... such that A,, -> - oo as n -* oo. Some of the eigenvalues may be multiple, but the first is simple, and the corresponding eigenfunction can be chosen
to be >0. In this section, we will investigate the probabilistic meaning of to and the
corresponding eigenfunction uo, and derive some characterizations of lo, ending with a "variational formula" for A,,. Symbolically, we will show that
4
q'1 !! kP2=V1i :! (PI:l0.
We will define the symbols above as we need them. The first inequality involves the function w used in the last section (and is from Chung and Li (1983)). Let w f(x) = E. (exp Jf(B)
(2)
dtl
)
cpo= -sup{O:we+e *ooinG}. 1
8.8
EIaenvshm of e + c
?Au+cu+Ou=O
265
inG
u=O inOG has a unique solution u - 0, so - 0 is not an eigenvalue. To work with go, it is convenient to recast it in a more compututitinal form. Let at = supE.(exp(ct); i > t) x
where
Jc(B5) A
Ct =
0
The strong Markov property implies that as+t = E.x(eXp(cs+t); i > s + t) = E.(eXp(cs)EB( )(exp(ct) l(t>t)); T > s) < a,EE(exp(cs); T > s) = atas,
Taking logarithms gives log as+t < log as + log at,
that is, bt = log at is subadditive. From the last observation, it follows easily that
(*) lim b/t = inf bs/s. f-.O
S>0
Proof It is clear that lim inf b/t > inf bs/s. t-oo
s>o
To prove the other inequality, observe that if s > 0, ns < t < (n + 1)s, and r = t - ns, then subadditivity implies
bt
co gives
lim sup bt/t < bs/s, t- ao
which proves (*) since s is arbitrary.
Let p, = limt-., (1/t)logat. It is easy to see that (3)
go <
Proof Let 1 > Cpl. Then if t >- to, log at < It (i.e., at < ett), so if 0 < -1, then
E. (exp (fo (0 + c) (BS) ds) ;T > t) = eetat <
f
f
2"
$
PDE's hut ('in no Solved by Rum doll a Brownlnn Motion
from which it follows easily that E.,
(exp (j
(0 + c) (B.) ds)) < 00.
Since this holds whenever -0 > cpl, it follows that
cpo= in fl-0:wc+0 *oc in G} 5 cp1. Having proved that 2o < cpl, we are now in a situation where we can use results of Donsker and Varadhan (1974-1976). Our next step (following their (1976) paper) is to show that (4)
cpi
=inf sup c + Au u
x
where the infimum is taken over all u e C+ = {u a C°° : u > 0 on G }.
Proof If 1 > >/i2, there is a ueC°° that is > 0 on G and has
c+Au
Ex(exp(f c(Bs)ds) ;T > t1
To prove this inequality, we observe that Au/2 + cu :!5; lu, so
vr- Av-cv=lue1`-
u+cule">0 inG x (0,00)
v(t, x) > 0
///
v(0, x) >- 1
on 8G x (0, oo) on G.
and using Ito's formula gives s
v(t - s, Bs) exp(c.,) - v(t, x) = f - V, (t - r, B,) exp(c,) dr 0
exp(c,)Vv(t - r, B,) dB,
+ 0
+ f v(t - r, B,) exp(c,)c(B,) dr 0
Ov(t - r, B,) exp(c,) dr.
+2 0
Since (VI -
A
v-
cv) >- 0, we see that
v(t - s, Bs) exp(cs)
8.8 Figeavdua of n + e
267
is a nonnegative local supermartingale on [0, T), so
Ev((t - T)+,
v(t, x)
Ex(eXp(c,); T > t),
proving the desired inequality and completing the proof of (4). Let >/il = sup inf
where the Jc + Adµ, 2u/
supremum is taken over all
probability measures on G and the infimum over all u c- C' 0. In this notation,
biz=infsuprrc+Audy µ
U
`J
2u
.
J
The next step is to show that, in our situation, the infimum and supremum can be interchanged. (5)
02 = 0i
Proof We only need to know that 0z < t/il, but since inequality in the other direction is trivial, we begin by proving it. Let F(x, y) be a function defined on some product space A x B. If (x,,, c A x B, then sup F(x,,, y) > F(x,,, x
y
y such that sup F(x,,, y) j inf sup F(x, y) x
y
inf F(x,
y
T sup inf F(x, y),
x
y
x
it follows that inf sup F(x, y) > sup inf F(x, y), x
y
y
x
proving iP2 > 01 .
The last inequality is valid for any function F, but as anyone who has heard of game theory can tell you,
(i) simple examples, for example, F(x, y) = lx - yl, show that we may have (here x - u, y - µ) inf sup F(x, y) > 0 = sup inf F(x, y) x
y
y
x
(ii) but if x -+ F(x, y) is convex for fixed y, y - F(x, y) is concave for fixed x, and F is continuous in a suitable sense, then inf sup F(x, y) = sup inf F(x, y). x
y
y
x
Theorems of the type described in (ii) are called min-max theorems. The one we use is due to Sion (1958).
26N
N
PDE's 71N ('rn 14 Solved by Rowing r Brownimn Motion
To apply this result, we write, with u = c-", Di(I" = (." l) D(je"
(."l),I/1 -+ ,I(I),h)2.
This change of variables makes (Let
+
J G(`µ,
l d= J(c++iXi2)d.
h) denote the right-hand side of the last equation. If we use the usual weak topology on p, then for fixed h e bC °°, p - G(p, h) is continuous and linear. On the other hand, if we use the C2 topology on C°° (i.e., h -+ h if and only if D,,h - Dah uniformly on G for all a with IaI < 2), then for a fixed probability measure, h
G(µ, h) is continuous and
G(p,0hl+(1-0)hz)= f cdp+0 f Mldµ+(1-0)f AZzdu +
JIOVh1 + (1 - 0)Vh2I z dµ.
2 To conclude that h - G(p, h) is convex, it suffices to show that for all a, b, and 0 e [0,1 ] (Oa + (1 - 0)b)2 < Oaz + (1 - 0)b z.
Proof Since (Oa + (1 - 0)b)2 < (OIal + (1 - 0)1b 1)2 and the result is trivial when a = 0, letting c = IbI/IaI it suffices to show that for all c > 0, (0 + (1 0)c)2 < 0 + (1 - 0)cz. This is true when c = 0. Differentiating the difference reveals that
aac = 2(1 - e)c - 2(1- 0)(0 + (1- 0)c) >0 ifc> 1
<0 ifc<1. Checking the value at c = 1 reveals that 0 + (1 - 0)cz - (0 + (1 - 0)c)2 = 0, so the inequality holds for all c. With the inequality above verified, we have shown that the hypotheses of Sion's theorem are satisfied and, hence, that (5) holds. This brings us to the fifth stage in the cycle: (6)
01 < (P l
1 = lim log sup E,,(exp(ct) ; r > t). rm t x
Proof We begin by making the simple observation that y f c dp is continuous and J(µ) = inf f A u/2u dp is upper semicontinuous, so there is a measure µo with 01 = JcdPo + J(µo).
8.8
269
EIQemduen of A + e
This result frees us from having to deal with supremum over µ in the dellnition of Vi, and prepares us for the main part of the proof. In order to focus attention
on the main steps of the proof, we will not prove first equality in detail. A complete proof (which involves discretizing time and passing to the limit) can be found on pages 601-602 of Donsker and Varadhan (1976). Let v(t, x) = Ex(exp(c); i t). 8t
j'lo(v(tx))d/Lo(x) =
> J v(t x) at v(t1
(t, x)
dµo(x)
x) (A + cl v(t, x) dµo(x)
J
_ f cdµo + J Av(t,x)dµo(x)
? Jcdµo+APO) =0i, so
f log(v(t, x)) dµo > 4i, t. Jensen's inequality implies that log Jdµo(x)v(t, x) > 0, t,
and we have
J dµo(x)v(t, x) > exp(4i, t), which proves (6). The last link in the chain is: (7)
4P 1 5 4.
Proof Let 1= q,1. We want to show that 1 is in the spectrum, that is, if we let R1f(x) = f
m
e-uEx(.f(B,)
exp(c)) dt,
0
then R, is not a bounded operator on C(G). If R, 1 is bounded, then we have 00 >
j'dzo(x)J e`v(t,x)dt o
f dte " f dµo(x)v(t,x), ,J
but it follows from (4), (5), and the proof of (6), that the last expression is
270
N
PDE's That Can Be Solved by Rumina
Brownian Motion
dt r "cxp(ot t) = oo.
Z 0
This is a contradiction, which proves (7) and completes the chain
A0
1. The equality of A0 and p0 in the case c > 0 is due to Khas'minskii (1959), who showed that )0 < 0 was equivalent to the existence of a solution
of2Au+cu =0that is > OonG. 2. The function u that minimizes sup c + x
Au 2u
is the eigenfunction go associated with 2o. The equality of 02 and A0 was first proved by Protter and Weinberger (1966). 3. The function
J (µ) = inf
f
0u 2u dµ
defined in the proof of (6) is -1 times what Donsker and Varadhan call I(µ). If dµ = fdx where f is smooth, then
J(µ) = -
f JVgl2dx
where g =f112 (see Donsker and Varadhan (1975-1976), I, Section 4), so substituting this equality in the definition of 01 gives
Ao = - inf
cg2 + I Vg I2 dx,
9 E Lz J
119112=1
the classical Rayleigh-Ritz variational formula for the first eigenvalue. The approach we have taken is certainly not the most direct way of proving this result; see Courant and Hilbert (1953), vol. I, Chapter 6. 4. The results we have proved above for Brownian motion are only a small part (and a relatively trivial one) of the theory of large deviations for Markov processes developed by Donsker and Varadhan. Their theory is one of the most important developments in probability theory in the last ten years, but you will have to learn about this from somebody else.
9 Stochastic Differential Equations
9.1 PDE's That Can Be Solved by Running an SDE Let Lf(x) = i >;, A,j(x)D,j f(x) + Y, b,(x)D, f(x), where the A,j(x) and b,(x) are
(for the moment) arbitrary. In this chapter, we will consider the following equation : (1)
(a) u, = Lu in (0, oo) x Rd (b) u is continuous in [0, co) x Rd and u(0, x) = f(x).
In Section 8.4, we solved (1) in the special case A(x) - I by first solving the stochastic differential equations
dX"=dB,+b(XX)dt Xo =x and then running the resulting processes to solve (1): u(t, x) = Ef(X, ).
On the basis of the results for A - I, it seems reasonable to try to solve the general case by solving
Xo = x
(*) dXX = o (XX) dB, + b(Xx) dt
(where o is a d x d matrix) and letting u(t, x) = Ef(XX ). To see which a to pick, we apply Ito's formula to get s
u(t - s, X) - u(t, Xo) =
- u, (t - r, X,) dr 0
+I D,u(t-r,X,)dX, '
oS s
+2
ij
Jo
D;,u(t - r, X,) d<X', X'>r 271
272
9
Stochutlc I)Ifferentlol Equatloou
(here we have taken the liberty of dropping the superscript x to simplify the formulas). Now dXs = bi(XS) ds + Y aij(Xs) dBs ,
j
so it follows from the formula for the covariance of two stochastic integrals that <Xi1XjX= Y_
aik(X,)ajk(Xs)dS
k
and we have 5
u(t - s, X5) - u(t, X0) = f - u, (t - r, X,) dr 0
+Y
Diu(t - r, X,)bi(Xr) dr
i
+ a local martingale 5
+ 2Y
Diju(t - r,Xr)(aaT)ij(X,)dr,
so if CUT = A, then we have
u(t - s, XS) - u(t, Xo) =
(- ur + Lu) (t - r, X,) dr + a local martingale. 0
The condition A = auT obviously restricts the set of A's we can consider, for if A = auT, then for each xERd"1 (i.e., Rd viewed as d x 1 matrices), XTAx = XTQQTX = Ior TXIZ > 0,
that is, A is nonnegative definite. If we assume (as we can without loss of generality) that A is symmetric, then this condition is also sufficient, because results in linear algebra tell us that any nonnegative definite symmetric matrix can be written as UT DU, where U is an orthogonal matrix (i.e., UTU = I) and D is a diagonal matrix. This observation allows us to define a by setting or = UTCU, where C > 0 is the diagonal matrix that has CZ = D. With this choice of a, we have: (2)
If u satisfies (i) and X satisfies (*), then M., = u(t - s, Xs) is a local martingale on [0, t).
Proof By computations above,
u(t - s, Xs) - u(t, Xo) = f (- ut + Lu) (t - r, X,) dr + a local martingale. o
If If u satisfies (i), the first term is = 0.
Remark: Looking back at the proof of (2), we see that after we used Ito's formula to conclude that
9.1 M's That Can Be Solved by Running n SDE
273
u(t - s, Xs) - u(t, Xo) = f - U, (t - r, X,) dr 0
+1: f5 Diu(t-r,X,)dX, i
2
0
+ 1Y_
D;ju(t - r, X,)d<X`, XJ) o S
all we did was work out what dX, and d<X`, X'>, were and plug in their values, so we have the same conclusion if X satisfies : (**)
(i) For each i, XX - f o b;(X) ds is a local martingale (ii) For each i, j,
A;;(X)ds.
<X`,X'>, = J0
We will see in the next section that there is essentially no difference between (*) and (**).
As has been the case many times before in Chapter 8, the last result leads immediately to a uniqueness theorem. (3)
Let X be a solution of (*) (or (**)), with X0 = x. If there is a solution of (1) that is bounded, it must be v(t, x) = Ef(X).
Proof If u satisfies (1), then MS = u(t - s, X) is a bounded local martingale on [0, t) that converges to f(X) as s r t and satisfies M, = E(f(X,) j A), so taking s = 0 proves (3).
Remark: (3')
The reader should note that (3) is also a uniqueness result for (*):
Suppose there is a solution of (1) that is bounded. If X, and X7 are two solutions of (*) with X0 = Xo = x (constructed, perhaps, on different probability spaces using different Brownian motions), then
Ef(X:) = Ef(X') (2) and (3) may look the same as the corresponding steps in Sections 8.1 through 8.7, but when we start to consider the existence of solutions, things become very different. We first have to construct solutions of (*) and then run them to produce solutions of (1). The first task will be accomplished in Sections 9.2 through 9.5 (since I am a probabilist, a neophyte, and a pedagogue,
we will spend some time investigating the countryside on the way to our destination-constructing "weak" solutions of (*)). In Sections 9.6 and 9.7, we turn our attention to v(t, x) = Ef(XX) and prove the analogues of the results
274
9
Stochutk 1)Iffercntrl Equatlou
that we called (4), (5), and (6) in Sections 8.1 through 8.7. The first two results, which are easy consequences of the Markov and Feller properties, are dispensed with in Section 9.6. In Section 9.7, we confront (but do not conquer) the problem of proving (6).
9.2 Existence of Solutions to SDE's with Continuous Coefficients In this section, we will describe Skorohod's approach to constructing solutions of stochastic differential equations. In order to focus our attention on o and not on b (which we have already considered in Section 8.4), we assume that b - 0. The same method works when there is a (bounded) continuous b * 0, but there are more estimates to do and, as parenthetical qualification suggests, many of the complications involve issues (e.g., explosions if b is too big) that we have considered earlier. Skorohod's idea for solving stochastic differential equations was to discretize time to get an equation that can be solved by induction, and then pass
to the limit and extract subsequential limits to solve the original equation. Given this approach, it is natural (and almost necessary) to assume that each A,3 is a continuous function of x, and, for the moment, we will suppose that A is bounded, that is, IAij(x)I < M for all i, j, x. For each n, define X"(t) by setting X"(0) = x and, for m2-" < t < (m + 1)2-",
X"(t) = X"(m2-") + o,(X"(m2-"))(B, - B(m2-")), where the second term is the matrix Q(X"(m2-")) times the Brownian increment
(B, - B(m2-n)). Since X" is a stochastic integral with respect to Brownian motion, the formula for covariance of stochastic integrals implies
<X", xi>, = k k
J0
Aij(Xn([2"s]/2")) us,
and it follows that if s < t, then
<M(t-s), so we have (see Section 6.3)
E sup I X,i(u) - XX(s)I° < CEI<XX>, - <Xi>slp12 uc[s.r]
so by taking subsequences, we can guarantee that for each i and rational t, converges weakly to a limit as k -+ oo. Taking p = 4, we see that
9.2
Existence of Solutions to SDE's with Continuous Coefficients
275
CM(t - s)2,
E sup I XX(u) u E[s, t]
so the processes X satisfy Kolmogorov's continuity condition uniformly in n. Combining the last observation with a standard result (Theorem 14.3 in Billingsley (1968)), we can conclude that the measures induced on (C,'6) by the X"(k) converge weakly to a limit QX on (C, W). I claim that under Q., the coordinate maps XX(co) = cot satisfy (*). To prove this, we will first show that if f e C 2 and we let Lf = 2 Y A;,Dijf then we have (1)
f(X) -f(Xo) -
Lf(XS)ds J is a local martingale/QX.
Proof It sufficesto show that if f, D; f, and D,j f are bounded, then the process above is a martingale. Ito's formula implies that
f(X"(t)) -f(X"(s)) _ >J iD1f(X"(r))dX.(r) s
t
+ 2Y
Dijf(X"(r))d<X,:,XI>,, s
and it follows from the definition oY" that
f 0
so if we let
L"f(r) _ Y A,J(X" ([2"r] 2-"))Djjf(X, ),
then f(X"(t)) - f(X"(s)) - f s L"f(r) dr is a local martingale. The Skorohod (1956) representation theorem implies that we can construct processes Yk = X"(k) on some probability space in such a way that with probabil-
ity 1 as k -- oc, Y, converges to a limit Y, uniformly on [0, T] for any T < oo. If s < t and g : C - R is a bounded continuous function that is measurable with respect to F,, then
E(9(Y) IftY) -
-f(Y) - JLf(Yr)dr}) = limE(9(Yk).{f(Yk) -f(YY) - f" L,f(r) drjI = 0,
which proves (1).
276
9
Stochautlc l)Ifferendal Equation
xixj, we see that under Q. the
Applying (1) to fi(x) = xi and coordinates X; are local martingales with
Aij(X)ds,
<X`, X'X = 0
that is, X solves the problem we called (**) in Section 9.1. The final step is to construct a Brownian motion B such that
(*) Xt-X0= ra(X.)dBs. J0
If a is invertible for each x, then the proof is trivial. We let Bt = fo u-1(X5) dX5.
The associative law implies that this process satisfies (*). To see that it is a Brownian motion, observe that each component B; is a local martingale and that
B',Bj>t-
aik1(Xs)Qji1(Xs)A kI (XD ds ki
-
o t
(o-'Au-1)ij(/Xs) ds = bijt,
0
since a is symmetric and
a-'Au-1 = a-1a2a-1 = I. When a is not invertible, for example, when a - 0, one has to first enlarge the space by adding independent Brownian motions and then use some linear algebra to get around the fact that a-1 does not exist. Since the details get a little messy, we leave it to the reader either to figure out how to do this or to look up the answer in Ikeda and Watanabe (1981) on pages 89-91.
The discussion above shows how to solve (*) when A is bounded and continuous. We now deal with a general continuous A. Let 0 < g(x) < 1 be a continuous function such that A(x) = g(x)A(x) is bounded. Let Q(x) _ g(x)1"2a(x'') and let Y5 be a solution of dYs = &(YY) dB,. Let
at=Jg(Ys)ds (
T=
g(Ys)ds, J
and for s < T, let
ys=inf{t:a,>s} Xs = Y(ys)
Since ys is an increasing family of stopping times, each component Xs is a local martingale. To compute <Xi, Xj>s, we observe that if t < T, then t
XXX? - JA()ds = 0
YY(t, YYit) - JAiJ(Y(s))ds. 0
9.2
Existence of Solutions to SDE's with Contlnuoue Coefficients
Changing variables s = o, in the integral above and observing that )'(n,) converts the right-hand side to
277
r
Ylt)
yr(nYv) -
fo
which is a local martingale, so we have .
JAj(Xs)ds.
<X' 1X'>1 =
0
As before, we can construct a Brownian motion B such that for t < T,
X, - X0 = J a(Xs) dB, If P(T = oo) = 1, then we have solved (*) for all times and we are done. If P(T < oo) > 0, then, as you might guess from the discussion in Section 8.4, we are also done, for the process has exploded, in other words, lim,1T Y = 00 a.s. on IT < oo }. It is easy to explain why this is true-for any R, there is a 6 such that if Ixl < R and S2R = inf{t : IXI > 2R}, then Px(S2R > S) > 6; therefore, with probability 1, X, cannot leave I x I < 2R and return to I x I < R infinitely
many times in a finite time interval. A full proof, however, requires the strong Markov property, which we have not established, and a number of unpleasant
details, so we leave the rest to the reader. Again, a complete proof can be found in Ikeda and Watanabe (1981), this time on pages 160-162. To get an idea of when explosions occur, consider the following: A;i(x) = (1 + Ix16)5;j. _ In these processes, if we let g(x) = (1 + lxl8)-1, then A(x) = I, so only the Example 1
second part of the construction is necessary: W
T= Jo ( 1+IBIb)-1ds. In d = 1, 2, Brownian motion is recurrent, so T = oo (and the process never explodes). In d > 3, Fubini's theorem implies that
ExT = E.
(1 + iBsl6)-1 ds 0
=C
lx
The integrand -
f
-
1
1
y1d-2 I + Iylb lyl2-a-d
dy.
as y -. oo, so if
r2-b-drd-1 dr < o0
(i.e ., S > 2), we have EXT < oo.
Conversely, we have (2)
If trace(A) < C(1 + lxl2), then QX(T = cc) = 1.
27K
9
Stochutlc Differential Equ.tlor
Proof This proof is the same as the proof of (7) in Section 8.4. Let p(x) _ 1 + 1x12. By Ito's formula, w(Xt)-(P
(x°)=Y
it,
i=1 Jo
o t
d
= a local martingale +
A;;(X5) ds. 0i=1
The last integral < C f o cp(X5) ds, so another application of Ito's formula shows
that e-ctcp(X) - tp(Xo) = a local martingale + Y f t e-cSAjj(Xs) ds 0
+ J(_C)eco(X)ds = a local supermartingale.
If we let T. = inf{t: IXI > n}, it follows from the optional stopping theorem that Note: Our treatment of the existence of solutions follows Section 4.2 of Ikeda and Watanabe (1981), who in turn got their proof from Skorohod (1965). The treatment of explosions here and in Section 8.4 is from Section 10.2 in Stroock and Varadhan (1979). They also give a more refined result due to Hasminskii.
9.3
Uniqueness of Solutions to SDE's with Lipschitz Coefficients The first and simplest existence and uniqueness result was proved by K. Ito (1946).
(1)
If for all i, j, x, and y we have I a (x) - aij(y) I < KI x - y I and I bi(x) - bi(y) I < KI x - y1, then the stochastic differential equation X= x+ fo t a(XS) dB., + fo t b(X) ds
has a unique solution.
Proof We construct the solution by successive approximation. Let X° = x and define :
X' = x + fo a(X: -1) dBs
+ J b(X: -1) ds for n > 1. 0
9.3
279
Uolqueo.M of Solution to SDE'. with Llp.chltz Codflclsnte
Let A,,(t) = E ( sup Xs - Xs -' j
.
I
Ossst
J
We estimate A. by induction. The first step is easy: XSl
= x + a(x)Bs + b(x)s,
so
IX91- X°I < Ia(x)Bsi + Ib(x)Is.
Squaring and using the fact that sup Ia(x)BSI ossst
a 1112
sup lo(x)BI,
o<s<1
gives
A1(t) < C1t + C2t3/2 + Ib1212 < C(t + t2)
where
C=C1+C2+lb12. To bound Am(t) form > 2, we recall that Ia + b12 < 2a2 + 2b2 and observe that this implies that for n >- 1,
2E sup
ta( s) -
a(Xs-1)dBs I 2
o5t5T fo t
+ 2E sup
05tsT
2
fb(Xn) - b(Xs -1) ds 0
To bound the second term, we observe that the Cauchy-Schwarz inequality implies that e b'(Xs) - b`(Xs 1) dsl2 <
\
\J
f t (b`(X:) - b`(X: -1))2 ds/
(fo
/, 1 dsJ
so we have
(a) 2E sup Jb(x:) - b(x:')ds O
2
< 2TE f T Ib(X:)
- b(Xs-1)I2ds
0
0
T
< 2dTK2E
I Xs - Xs -1 12 ds. 0
To bound the other term, let ai be the ith row of a and observe that Doob's inequality implies that 2
E sup (\fo (a,(X.,) - a,(Xs 1)) dBs) < 4E 0
J
(f
T
\0
l2
(ai(XS) -
fT
= 4E
I a,(Xs) - a,(X:-1)I2ds, 0
J
280
9 Stochudc Differential Equations
so we have 2
(b) 2E sup
fo, (a(XS) - a(Xs-1)).dBs
O
2
< 2EY sup
i=10
(ai (X") - ai (X"-')) - dB,
I f,0
<8EY JI1(x:) - o(X:-1)I2ds t=1
0
T
< 8d2K2E
IXs - Xs-1I2ds. 0
Combining the last two inequalities shows that T
(c) 0"+1(T) < BE
I Xs - Xs -1 12 ds 0
A. (s) ds 0
(where B = 2dTK2 + 8d2K2 depends on T). Since Al(t) <- C(t + t2), iterating (c) gives that if t < T, then
C(s + s2) ds = BC (+) t
A2(t) < B 0
A3(t)
and it follows by induction that if t <- T, then (d) O"(t) < Bn-1C
2t"+1
t"
With this estimate established, the rest of the proof is routine. Chebyshev's inequality shows that
P (Osup I X, -
X°-'l
> 2-n) < 22nD"(T)
5t5T
Since the right-hand side is summable, the Borel-Cantelli lemma implies that
P ( sup I Xn -
X°-'l
> 2-" i.o
0
lJ = 0,
so with probability 1, Xe -+ a limit Xt uniformly on [0, T], and it follows from the estimates above that
(e) for all m < n < oo,
E(sup IXm-Xsl2) <( n Ak(T)1/2 //I2. 05s
f
=m+1
If we let XX° = Xt, then the result also holds for n = oc.
9.3
Uniquenew of Solutions to SDE's with Lipechits Coeflkbete
281
Proof If n < oo, then it follows from the triangle inequality that
sup IX9 -XSI2
O<s
E Ok(7
sup IXm - XT-1I
05s
k=m+1
2
2
k=m+1
Letting n -> oc and using Fatou's lemma proves the second claim. At this point, we have assembled all the ingredients. The rest of the proof
consists of applying what we have learned to prove (1). To see that Xt is a solution, we observe that if we let
Y:=x+ J a(Y)dB5 +
,
th en it follows from the proof of (c) above that (c')
Ersup IYt-ZtI2) -BE
f
`0
T
IY-ZsI2ds.
o
Letting Yt = X' and Zt = Xt shows that
E sup IX°+1 0
fT
Xti2) < BE
f
J0
IXs - XXI2 ds
<;BTE (sup IX '- X, OSs
f
0
byO,e solimXtn+1=X
t
To prove uniqueness, observe that (e) implies that
E( sup Ix-X,I21 <(-
/
0<s
z
Am(T)1/2
\`m=1
so if Y is another solution with E supo<ss T I Ys I2 < oo and we let
4P(t)=E(sup IXs0<s
Y11
2),
/
then p(0) = 0, (c') implies that lp(t) < B
f t (p (s) ds, 0
and it follows from an easy argument that cp(t) < eeBt for any s > 0, so cp = 0 and, hence, X = Y. To remove the integrability condition on Y, observe that for any R < oe, we can modify a and b outside IxI > R in such a way that a and b still satisfy IQ;;(x) - a;;(y)I < KI x - yI and Ib;(x) - bt(y)I < KI x - yI, but
a and b are = 0 off some compact set. When we do this, any solution has E(supo<s
282
9
Stochwflc Differential Equations
Before the reader forgets the proof given above, we would like to observe
that the argument above gives a continuity result for x - Xx (the solution starting at X0 = x). Let X and Y denote solutions of (*) with X0 = x and Yo = y, and let X" and Y" be the sequence of processes generated by the construction above when X° = x and YO - y. If we let
0;,(t)=E(sup
IXS - YsI2
J
o<s
observe that Do(t) = I x - yI2, and iterate (c'), we see that if t < T, then
A (t)
Blx-yl2sds=B2lx-y12?
O'(t)
2
fo
It follows by induction that if t < T, then (d') 0;,(t) < B"I x - y1271,' so summing as we did in the proof of (e) gives
(e') E ( sup IX1 -
YYI2)
<
2
(
Ak(T)1/2/
<
Ix-y12CT,
where
= 00 ((BT)" \1/z CT
- k-o
n.
fI
< 00
and
BT = 2dK2T2 + 8d2K2T. Remark: If X0 = X and Y° = Y are random, then the same result holds with Ix - y12 replaced by El X - YI2. At this point, we have completed Ito's construction of solutions to (*). Before we proceed, however, the reader should note that we are not in the usual Markov process setup (cf. Section 1.1). We started with a Brownian motion defined on some probability space (S2, S, P), and by successive approximation we defined for each x c R° a process XX that satisfies
X" = x +
ft
Jo
a(X') dB., + f t b(X') ds. Jo
In other words, we have one probability measure (P) and a family of stochastic processes (X', x E Rd), rather than one set of random variables (the coordinate maps on (C, ')) and a family of measures (Px, x e Rd) (which was what we got
9.4 Some Examples
283
from Skorohod's construction in Section 9.2 and from the Camermi-Martin transformation in Section 8.4). Note: The material in this section has appeared previously in a number of places. The proof of (1) above is a hybrid of the proofs in Friedman (IV75), pages 98-102, and Stroock and Varadhan (1979), pages 124-126, with cmc small improvement : We estimate A(t) = E sup { X; - X; -1 I2 : 0 S .c' s t , rather than ElXt - x1 I2, and this simplifies the argument somewhat.
9.4 Some Examples Having solved our stochastic differential equation using two methods under two different sets of conditions, it is time to compare the results by looking at some examples. We begin with a "trivial" one. Example 1 Let A (x) - 0. Then the stochastic differential equations become deterministic : dXs = b(X,) ds.
Skorohod's approach allows us to construct solutions by successive approxima-
tion whenever b is continuous: In contrast, Ito's approach requires that b be Lipschitz continuous but implies that for such b, there is, for each x e R°, a unique (deterministic) process X, that has X0 = x and that solves (*). A simple family of examples shows that uniqueness need not hold if the assumption of Lipschitz continuity is replaced by Holder continuity of any order < 1.
Example2 Letd= 1,a-0,b(x)=Ixl6,0
so to make dXs = b(XX) ds, we first pick p so that p6 = p - 1 (i.e.,p = (1 - 6)-1) and then pick C such that Cp = Ca (i.e., p = C1-b or C = 1/(1 - 8)1-a)
Example 3 Let d = 1, a(x) = lxla, b - 0. Since a is continuous, Skorohod's approach allows us to construct solutions by successive approximation for any
S > 0. In contrast, Ito's approach requires that Q(x) = I X1612 be Lipschitz continuous, so we are restricted to S > 2, but in this range we can conclude that the solution is unique.
Although the condition required by Ito's approach is not sharp in this case, there is a good reason why Ito's approach, or any other approach that proves uniqueness, does not work for small 6-the solution of (*) is not unique
284
9
Stochutk Differential Equations
when 6 < 1. To prove this statement observe that by using the approach in the second part of Section 9.2, we can think of a(x) = 1/g(x) where g(x) = IxI-a and solve the equation by time-changing a Brownian motion. If 6 < 1, then
(a) Eo f IB,I ads= f s-612EoIBi ads 0
0
=
fos-112 dsl EoIB1I-6 < o0
(since the first factor is < oo for 6 < 2 and the second for 6 < 1), so we can construct a solution of dX, = I XSlaiz dB, that starts at 0 and does not stay there. Once we have two solutions starting from 0, it is easy to see that there are
two solutions starting from x 0 0. If x > 0
and
i = inf{t : B, = 0}, then by
results in n Section 1.9,
(b) E. J
xlyl-a(x AY)dy
TIBSI-a1(a,<x)ds= I z
Y1-ady < 00
(whenever 6 < 2), and we have I B,I a1(Bs>z)ds <
1-6
T < co
fol
(for any 6 < c0). Combining the last two results gives
T=
BsI -ads < oo
a.s.,
0
that is, starting at x, we hit 0 at a time T < c0. Once we hit zero, we have two choices : We can stop, or we can continue by using the time substitution (or, if you are more sophisticated, you can stop the first time the local time of B, at zero exceeds a fixed or exponentially distributed level, or ... ). At this point, we have settled the uniqueness question when S z 2 (unique) and when 6 < 1(not unique). Looking at the proofs of (a) and (b) more carefully reveals that you cannot escape from 0 when 6 >_ 1 and you cannot reach 0 when 6 >- 2 (see Exercises 1 and 2 below for converses). On the basis of this conclusion,
you might guess that the solution is unique when 1 < 6 < 2 and that all solutions stop when they hit 0. This conjecture is indeed true. The first fact is a consequence of a theorem of Yamada and Watanabe (1971). (1)
Suppose that (i) there is a strictly increasing function p with la(x) - a(y)I < p(I x - yl) that has p(0) = 0 and
JP2(u)du = c0 for all a> 0
9.4
285
Some I xemplee
(ii) there is an increasing and concave function ) with I b(x) - b(y)I c A (I % that has A (O) = 0 and
i'I )
A-1(u)du= oo forallE>0. fo,
Then (*) has a unique solution.
Proof See pages 168-170 of Ikeda and Watanabe (1981). (1) implies that our equation has a unique solution when 6 > 1, so we now have a complete picture of Example 3. unique for 6 > 1 The solution of (*) is not unique for 6 < 1.
-
While (1) helps us with Example 3; Example 2, on the other hand, helps us to understand the reason for the difference between the assumptions about a and b in (1). When b(x) = x1 and a = 0, (1) implies that
unique for 6 > 1 I not unique for 6 < 1,
solution of (*) jl
which is the same as the conclusion for Example 3, except for the fact that it pertains to a(x) = Ixlaiz To justify my remark that "you cannot escape when 6 > 1," show that for all e > 0, Exercise 1
IBSI-1 ds = oc P. a.s. fo
Proof By scaling and monotonicity, it suffices to prove the result where a is replaced by T1 = inf{t : IB,I - 1}, and to do this, it suffices to show that T1
lim sup nom
fo
I B.I 11(IBSI <2 no ds > 0 PO a.s.
To prove the last result, let R0 = 0, Sn = inf{t > Rn : IB,I = 2-"}, Rn+1 = inf{tr> Sn : B, = 0} for n -> 0, and N = sup{n : S. < T1}. Observe that N f S, N >Y 2"(Sm - Rm) 0Tl > Y M=O Rm
J
m=0
where EN > 2" and the S. - R. are i.i.d. with E(Sm - Rm) = C2 2n. A simple argument (compute variances) shows that the lim inf >- C a.s. Exercise 2 To justify my remark that "you cannot reach 0 when 6 > 2," show
that if r=inf{t:B,=0}, then for allx
JIBI_2ds= oo P., a.s. 0
0
286
9 Stochudc Dlfferontlil Equtlow
Proof The game is the same as in the last example, but the stopping times are different and don't work as well. Let S_1 = 0, R. = inf{t > Sn_1 : IB,I = 2"j, S. = inf{t > R" : IB,I = 2-"} for n >- 0, and N = sup{n : S" < T1}. Again, we have T, 0
N`
Sm
N
> Y 2Z"(S, - Rm),
L
m=0 Rm
m=0
but this time EN = 1 and E22"(Sm - Rm) = C, so an even simpler argument shows that the lim inf = oo a.s.
9.5 Solutions Weak and Strong, Uniqueness Questions Having solved our stochastic differential equation twice and seen some examples, our next step is to introduce some terminology that allows us to describe in technical terms what we have done. The solution constructed in Section 9.3 using Ito's method is called a strong solution. Given a Brownian motion A and an x e R°, we constructed a process X, on the same probability space in such a way that
(*) X = x +
f
A(X,) dB, +
0
f
b(XS) ds.
0
In contrast, the solution constructed in Section 9.2 by discretizing and taking limits is called a weak solution. Its weakness is that we first defined X, on some probability space and then constructed a Brownian motion such that (*) holds. With each concept of solution there is associated a concept of uniqueness. We say that pathwise uniqueness holds, or that there is a unique strong solution, if whenever B, is a Brownian motion (defined on some probability space (f2,.F, P)) and X and X are two strong solutions of (*), it follows that, with probability 1, X = Xt for all t >- 0.
We say that distributional uniqueness holds, or that there is a unique weak solution, if all solutions of (*) give rise to the same probability law on (C, ') when we map co e f2 --> X(w) E C.
Ito's theorem implies that pathwise uniqueness holds when the coefficients
are Lipschitz continuous. It is easy to show that in this case there is also distributional uniqueness. (1)
If the coefficients o and b are Lipschitz continuous, then distributional uniqueness holds.
Proof Let B, and B, be two Brownian motions, and let Xe and X, be the sequence of processes defined in the proof of Ito's theorem when X° = X° = x and the Brownian motions are B, and Bt, respectively. An easy induction shows that for each n, X" and X" have the same distribution, so letting n - oo proves M.
9.5
Solutlonn Week and Strong, Unlqueneaa Questions
287
With two notions of solution, it is natural, and almost inevitable, to risk about the relationship between the two concepts. A simple example due to Tanaka (see Yamada and Watanabe (1971)) shows that, contrary to the naive idea that it is easier to be weak than to be strong, you may have it unique weak solution but several strong solutions.
Let a(x) = 1 for x > 0 and = -1 for x < 0. Let W be a Brownian motion starting at 0 and let Example 1
B: = I a(W)dW Since B = e W is a local martingale with = t, B is a Brownian motion. The associative law implies that
a(W)-B=u(W)Z-W= W, so we have
W=
Ja()dBs.
Since a(- x)
c (x) for all x
0, we also have
-W = f a(-W)dBs. 0
The last two equations show that there is more than one strong solution of dXs = o (xs) dBs. To prove that there is a unique weak solution, we observe that if dXs = a(XX)dBs, then X is a local martingale with <X> = t and, hence, a Brownian motion.
In the other direction, we have the following result of Yamada and Watanabe (1971) : (2)
Pathwise uniqueness implies distributional uniqueness.
We will not prove this because we are lazy and this is not important for the developments below. The reader can find a discussion of this result in Williams
(1981) and a proof in either Ikeda and Watanabe (1981), pages 149-152, or Stroock and Varadhan (1979), Section 8.1. The last result and Tanaka's example are the basic facts about the differences between pathwise and distributional uniqueness. The best results about distributional uniqueness are due to Stroock and Varadhan (1969). To avoid the consideration of explosion, we will state their result only for bounded coefficients. (3)
Suppose that
(i) A is bounded, continuous, and positive definite at each point (ii) b is bounded and Borel measurable. Then there is a unique weak solution.
Much of Chapters 6 and 7 in Stroock and Varadhan (1979) is devoted to preparing for and proving (3), so we will not go into the details here. The key
20
9 Stochudc Differential Equador
step is to prove the result when b = 0 and I Aij(x) - 5,i1 5 s for all x E R°. The reader can find a nice exposition of this part in Ikeda and Watanabe (1981), pages 171-176.
In the material that follows, we will develop the theory of stochastic differential equations only for coefficients that are Lipschitz continuous. We do this not only because we have omitted the proof of (3), but also because in the developments in Section 9.6, the proofs of the Markov and Feller properties, we will use the fact that our processes are constructed by Ito's iteration scheme. You can also prove these results in the generality of (3) by knowing that there is a unique solution to the martingale problem, but for this you have to read Stroock and Varadhan (1979).
9.6 Markov and Feller Properties Having constructed solutions of (*) and having considered their nature and number at some length, we finally turn our attention to finding conditions that guarantee that v(t, x) = Ef(X") is a solution of (1) of Section 9.1. In Section 8.1, when we dealt with XX = B, a Brownian motion, the first step was to observe that if f is bounded, then the Markov property implies that
Ex(f(B)IJs) = v(t - s, B,), so v(t - s, B,) is a martingale. To generalize this proof to our new setting, we need to show that the X, have the Markov property, that is, (1)
If f is bounded and continuous and v(t, x) = Ef(Xx), then
E(f(Xt)I.)=v(t-s,X ). Proof Let X,.x(t) (the process starting at x at time s) be defined as the solution of
X, = x +
a(X,) dB,. + fS b(X,) dr g
for t >- s, and X,,x(t) = x for t < s. It follows from uniqueness that if s < I < u, then
X,,x(u) = X,,x,,x(,)(u)
a.s.
(recall that all the random variables X, x(u) are defined on the same pyobability space, (C, ', P)). From the last result, it follows immediately that if < s < t < u, if A 1, ... , A. are Borel sets, and if f is bounded, then 0< s1 < E(.f(XXx); Xx(s1)EA1, ... ,
X x(s1) E A1, ... , X x(s,,) c (recall that Xx = X0.x(t)). To prove (1), it is enough to prove the following:
9.6
289
Markov end Pallor Propertler
v(u - t, Xx).
(2)
To this end, observe that for any y e R°, Xr.v(u) a u(Br+s - Br, s Z 0)
and is, therefore, independent of ., so
E(f(Xr,v(u))I) = Ef(X,,,(u)) = v(u - t,y). Now if Y : S - R° is .yr measurable and takes on only a finite number of values yl, ... , y,,, then n
X"Y(u) _ Y l(r=vt)X y1(u)
a.s.
It follows from the last result that
E(f(XX,r(u))I.) = v(u - t, Y). To prove this equality, let A e., with A c { Y = y,} and observe that it follows from the first result that E(f(Xt,y(u)); A) = E(f(X1,v,(u)); A)
= v(u - t, y;)P(A) = E(v(u - t, Y); A). To extend our results to a general Ye., (and, hence, to prove (1')), pick a sequence Yn of random variables that take on only finitely many values and have Y. - Ya.s. and EI Yn - YI2 - 0 as n - oo. From the continuity result (e') in Section 9.3 (or, to be more precise, the remark afterwards), we have that XX,rn(u) - XX,r(u)
in L2
and, hence, E(f(Xr,r(u))I.°1t)
in L.
By the result above for simple Y, the left-hand side is v(u - t, Yn). To complete the proof of (2) (and, hence, of (1)), it suffices to prove (3)
Suppose that f is bounded and continuous. Then for fixed t, x -> v(t, x) = Ef(X x) is continuous.
Proof The continuity result (e') in Section 9.3 implies that
E(sup IXS -XXIZSlx-yl2CT, `OSsST
lJ
where CT is a constant whose value depends only on T. From this result, it follows immediately that if xn - x, X, n - Xt in probability and, hence, v(t,
Remark:
Ef(Xx^) -. Ef(Xx) = v(t, x).
The proof of (1) given above follows Stroock and Varadhan (1979), pages 128-130. 1 think it is a good example of the power of the idea from measure
290
9
Stochardc Dlfferoadal Equadooe
theory that to prove an equality for the general variable Y, it is enough to prove the result for an indicator function (for then you can extend by linearity and take limits to prove the result). With the Markov property established, it is now easy to prove (4)
Suppose that f is bounded. If v is smooth, then it satisfies part (a) of (1) in Section 9.1.
Proof The Markov property implies that E(.f(X, )I#;) = v(t - s, X. ,x).
Since the left-hand side is a martingale, v(t - s, Xs) is also a martingale. If v is smooth, then repeating the computation in the proof of (2) (at the beginning of this chapter) shows that s
v(t - s, XS) - v(t, x) =
(- v, + Lv) (t - r, X,) dr + a local martingale, 0
where
Lf= 1 YA`VD;;f+ Yb`Dif, 2 ii
so it follows that the integral on the right-hand side is a local martingale. Since
this process is continuous and locally of bounded variation, it must be = 0 and, hence, (- v, + Lv) = 0 in (0, ao) x Rd. With (4) established, the next step is to show the following: (5)
If f is bounded and continuous, then v satisfies part (b) of (1) of Section 9.1.
Proof From the continuity result used in the last proof, it follows that if
x - x and t - 0, then
x in probability and, hence,
v(tn, xn) = E.f(X',,(t,,)) - f(x)
9.7 Conditions for Smoothness In this section, we finally confront the problem of finding conditions that guarantee that v(t, x) = Ef(X) is smooth and, hence, satisfies (1)
(a) u, = Lu in (0, oo) x Rd (b) u is continuous in [0, oo) x Rd and u(0, x) = f(x).
In this category, the probabilistic approach has not been very successful. By purely analytical methods (i.e., the parametrix method; see Friedman (1964), Chapter 1), one can show (2)
Suppose that Aj(x) and b;(x) are bounded for each i and j and that
9.7
Condltlm for Smoottm u
291
(a) there is an a > 0 such that for all x, y e Rd, >Y,Aij(x)yj > alYl l
(b) there is a J > 0 and C < oo such that for all i, j, x, and y, l Aij(x) - A.,(Y)l <- Clx - ylP l bi(x) - b.(Y)l < Clx - ylP.
Then there is a positive function p,(x, y) that is jointly continuous in all of its variables and such that if f is a bounded continuous function, then v(t, x) =
, Y)f(Y)
dy.
Remark: This result is a combination of several theorems in Friedman (1964); see Friedman (1975), pages 141-142. For analysts, p,(x, y) is the fundamental solution with pole at x (i.e., p,(x, ) = Sx at t -+ 0) ; for probabilists, p,(x, y) is the transition probability
Pt(x,Y) = P(Xt = Y)
On the other hand, the best result I know of, which can be proved by purely probabilistic means, is on page 122 of Friedman (1975). (3)
Suppose that b, a, and f are C2 and that these functions and their derivatives of order < 2 are bounded by C(1 + lxl") for some C, y < oo. Then v is smooth and, hence, satisfies (1).
Remark: The reader should observe that although (3) requires more smoothness for the coefficients, it does not require the "strict ellipticity," (a) in (2), and hence can be applied in situations where a degenerates. In this context, the results obtained from (3) are almost the same as those obtained by Olenik (1966) using purely analytical methods. Probabilists (and anyone who does not read Italian) can find this result in Stroock and Varadhan (1979), Theorem 3.2.6.
Proof Since the proof is rather lengthy, we content ourselves with simply giving an idea of what is involved by assuming that everything is bounded and indicating why D,v exists. (In our defense, we would like to observe that not even Friedman (1975) spells out the details for the second derivatives; see page 123.)
To deal with derivatives with respect to x,, we will show that x -' X; is, in the "L2 sense," a differentiable function of x. To explain this statement, we need a definition. A function g(x, co) on Rd x i2 is said to have 8g/8x, =fin the LZ sense if
E (g (x + he,, co) - g (x, co) h
ash - 0.
-.l (x, (v) 2 -
0
M 9 Stochudc DIfferendel Equetloas
With this definition introduced, we can state our first differentiability result as (4)
If Djr and Deb exist for all j and are bounded and continuous, then 8Xx18x; exists in the LZ sense; furthermore, if we let e; (t) = (8; (t), ... , e°(t)) = 8X, 18x;, then 8i satisfies
e;(t) = e; +
f 8/ (s)D;b(XS) ds + Jo
f i
8/ (s)DDQ(XS) dBs,
,J o
where e; is the ith unit vector (i.e., (e;)3 = bij).
Proof There is only one way to start the proof of a result like this. Let h > 0 and write
h (Xi +hei - Xj) = e1 + J(b(xi) - b(XS )) ds
+ J(o(x:'i) - U(XS )) dB,,. 0
To change the first integral on the right-hand side into something that looks like the first integral in the desired answer, we write f- b(XS) ds 0
_
f ids f ld9 db(Xs
+O(Xs+hei-Xs))
('o
0
ds f 1
de Y D;b(Xs + 6(Xx +hei - X:)) Xs
+hei.
=J
.
i
0
j
XS .;
h
where XS - is the jth coordinate of X. ,x. The same trick works for the second term, with the result that h
r` a(XX +he,) - a(Xss) dB.'
= =
1
dBs
h
0
f r dBs 0
1
dB-
d
u(XS + e(X: +hei - XS))
0
f dO i> DQ(X: + B(X: +hei - Xs ))
+he,._
XS .i
Xs
gives a stochastic integral equation for Xs+hei - Xs Ah(s) =
s
h
s
in which the coefficients are almost the ones given in (4). The last step in the proof of (4), then, is to prove a result that says that if the coefficients of the equation converge in a suitable way, then so do the solutions. There is a large
293
Notes on Chapter 9
body of literature on this subject, which goes under the heading "Ntuhllity of solutions" (e.g., Jacod and Memin (1981)). A result that is sufficient For our purposes is given on pages 118-119 of Friedman (1975); the desired concluNilm follows immediately from that result. (4) shows that if the coefficients a and b are C1, then so is the solution XX when viewed as a function of x. Once this result is shown, it is not hard to show that v(t, x) = Ef(XX) is C'. One does this by proving the "chain rule": (5)
D1Ef(Xx) = YEDf(Xx)81(t). Proof The proof iss based on the trick used t*.'i prove (3). We write E(/'J (Xs +hei)
-J (`Ys ))
=E
f1
d6 def(Xs + O(Xs +he, - Xs ))
= D f 1 deY Df(X5 + e(xs Jo
i
+hei
- X x))
XS +he
,j
- X51i
h
and let h -* 0. Further details are left to the reader. A complete discussion of the results in this section can be found in Section 5.5 of Friedman (1975).
Notes on Chapter 9 To steal a line from somebody, this book ends "not with a bang, but with a whimper." The results in this chapter are but a small sample of the results known about SDE's and their relationships with PDE's, and even worse, in many cases we have thrown up our hands and referred the reader to Friedman (1975), or Stroock and Varadhan (1979), or Ikeda and Watanabe (1981) for the details. In our defense, we can only say that the book had to end somewhere and that the three sources of which we have referred are all good places to learn more about the subject.
Appendix A Primer of Probability Theory
A.1 Some Differences in the Language For an analyst, reading the probability literature must be like being an American in England. The language that is spoken is basically the same, but some of the
words are different or have slightly different meanings. My first task, then, is to explain some of the colloquialisms that probabilists use. For convenience of exposition, we will begin at the very beginning.
A probability space is a triple (0, F , P) where fl is a set, F is a a-field of subsets of S2, and P is a probability measure, that is, a nonnegative, countably
additive function on .F that has P(Q) = 1. Let JP be the set of all Borel subsets of R. A function X : S2 - R is said to be measurable if for each Be R we have that {oo : X(co) e B} e F. For convenience, the phrase "X is measurable with respect to .-" is often abbreviated Xe JF, and measurable functions are commonly referred to as random variables. In measure theory, one often talks about a sequence of functions f converging to a limit f "in measure" or "almost everywhere." These concepts are also used in probability, but they go by different names. A sequence of random variables X is said to converge in probability to a asn -.oo. limit Xif for all X is said to converge to X almost surely if P(co : X,, (to) - X(CO) as n -+ oo) _
1. The last conclusion is usually abbreviated as X - X a.s. The words almost surely and their abbreviation a.s. are used throughout
probability as substitutes for almost everywhere and a.e. For instance, if P(co : X(co) = Y(w)) = 1, then we say that X = Y a.s.
As in measure theory, X. - X a.s. implies that X. - X in probability, and the converse is false, but X - X in probability implies that there is a subsequence X a.s. We will prove the last statement, because the proof gives us an excuse to state some more definitions. 294
2"
Al .Some Difference. In the Language
The indicator of a set A is the function
_ IA (w)
1
wEA
0 w0A.
The notation is meant to suggest that this function is 1 on A. We do not uMr because it looks too much like X, our favorite letter for random variables, and we do not call this a characteristic function, because that term is reserved for something else (see Chapter 6 of Chung (1974)). If A. is a sequence of sets, then lim sup An = {w : lim sup lAn = 1 } n-m
m
m
=nUAn.
N=1n=N
The set defined above is usually referred to as {w : w e A. i.o. }, where i.o. is short for infinitely often. As the next result indicates, we often make {w : co e A i.o.} even shorter by dropping the w's.
Borel-Cantelli Lemma. If In P(An) < oc, then P(An i.o.) = 0.
Proof For any N, P(lim sup An) < P (U An < = P(An). n=N
f
n=N
Letting N oo gives the desired result. With (1) established, it is trivial to prove that X. - X in probability implies that there is a subsequence Xnk -> X a.s. Let Ek 10, pick nk - oo such that P(I Xnk - XI > Ek) < 2-k, and then apply (1). This proof is, of course, nothing more than the standard proof from measure theory translated into the language of probability theory. Up to this point, all of the changes have been semantic. When we turn our attention to integration, we encounter our first serious differences in notation. What an analyst would write as
XdP (assuming that f I X I dP < oc) n J ` a probabilist writes as EX (assuming that EIXI < oo) and calls the expected Jn
value of X, or the mean of X. One clear advantage of the probabilistic notation is indicated by the typography of the last sentence-EX consumes less space and does not have to be displayed. There is also one clear disadvantage. To steal a quip from Dynkin, "If you use E to denote expectation with respect to
P, then what do you use for expectation with respect to Q?" The obvious answer, F, is obviously unacceptable. Dynkin's remedy is to write PX instead of EX. Although this suggestion has considerable merit and would be useful at several points in the text, we will stick with the traditional notation.
2%
Appendix A Primer of Probability Theory
Extending the notation above to integration over sets, we will let E(X; A) = fA X dP.
Again, the notation is for typographical convenience and is motivated by the fact that the set A often has a complicated description. The proof of the next result illustrates the use of this notation. (2)
Chebyshev's Inequality. Let Y > 0 and let cp > 0 be a function that is increasing on [0, oo). Then
cp(a)P(Y> a) < Ecp(Y).
Proof Since p is increasing and >_ 0,
cp(a)P(Y> a) < E((p(Y); Y> a) < Ecp(Y). This result is trivial but useful. The following is a typical application : (3)
P(I XI > e) <
Ez X
.
A.2 Independence and Laws of Large Numbers I have heard it said that "probability is just measure theory plus the notion of independence." Although I think that this statement is about as accurate
as saying that "complex analysis is just real analysis plus," there is no doubt that independence is one of the most important concepts in probability. We begin with what is hopefully a familiar definition and then work our way up to a definition that is appropriate for our current setting. Two sets A and B are said to be independent if P(A fl B) = P(A)P(B).
Two random variables X and Y are said to be independent if for all Borel sets A and B,
P(XeA, YEB) = P(XEA)P(YEB). Two or-fields
and 5 are said to be independent if for all A E _5F and Be
P(A fl B) = P(A)P(B).
The third definition is a generalization of the second: Let F = a(X) = the a-field generated by X(= the smallest a-field .IF such that X E.F), let W = a(Y), and observe that A e a(X) if and only if A = {w : X((o) E C } where C E M. The second definition is, in turn, a generalization of the first : Let X = 'A, let Y = 1B,
A.2
Independence and lawn of large Numbers
297
and observe that if A and B are independent, then so are A` and B, A' iinI ,,e, A and SZ, A and 0 and so on. In view of the last two remarks, when we define what it means for several things to be independent, we take things in the opposite order.
a-fields F1, ... , S are said to be independent if whenever A, E.#, I'ur i= 1..... n we have that P rn
_ fl P(A1).
`t-1 Af i
i=1
Random variables X1, ... , X,, are said to be independent if whenever A, E. 9 for i = 1, ... , n we have that
P(n {X,EAi}l = JJ P(X,EA,).
/
\i-1
i=1
Sets Al, ... , A. are said to be independent if whenever I c { 1, ... , n} we have that
P (n IEI Ai//
= 1 1 P(Ai). iEI
If you think about it for a minute, you will see that the third definition is what we get when we specialize the second to X = lAi. It is important to note that the last definition is not equivalent to requiring that P(Ai fl Aj) = P(A,)P(AJ) whenever i j (this is called pairwise independence). Example 1 Let X1, X2 , and X3 be independent random variables that have P(X, = 1) = P(Xi = -1) = 1/2, and let Al = {X2 = X3}, A2 = {X3 = X1}, and
A3 = {X1 = X2}. These events are pairwise independent, since P(A, fl A) = P(X1 = X2 = X3) = 1/4 = P(A,)P(AJ), but they are not independent, since P(A1 flA2 flA3) = P(X1 = X2 = X3) = 1/4 1/8 = P(A1)P(A2)P(A3) Of the three definitions in the second list above, the first is the most important, so if it is unfamiliar, it would be a good idea to spend a minute and prove the next three results to get acquainted with the ideas involved. (1)
If
J are independent and X, E , have El X; I < oo, then n
n
E H Xi = n i=1
HEX,.
i=1
Note that EIII7=1 Xii < oc is part of the conclusion. The proof of this result follows a plan of attack that is standard in measure theory: Prove the result first for Xi = 1Ai, use linearity to extend the result to simple random variables, Ximonotone convergence to extend it to Xi > 0, and write Xi = X+ to prove the result in general. (2)
If X1, .. . , X. are independent random variables, then the a-fields S i = a(Xi)
are independent, and, consequently, if fl, ... , f are Borel functions, then f1(Xi), . . . ,fn(XX) are independent.
298
Appendix
(3)
Generalize the proof of (2) to conclude that if 1 < n1 < nz < the f : S2 R"i-ni-1 are Borel measurable, then
A Primer of Probability Theory
A (X1 , ... ,
Jk(Xnk-1 +1 ,
.
< nk = n and
, Xnk)
and independent.
Hint: Start with f that are of the form H .1,1.(X.) and then use the approach described for (1) to work your way up to theJ general result. Sequences of independent random variables are very important in probability theory, because they are (hopefully) what is generated when an experiment is repeated or a survey is taken. Motivated by this example, one of the main problems of the subject is to give conditions under which the "sample mean" (X1 + + Xn)/n approaches the "true mean" as n - oo, and to estimate the rate at which this occurs. Much of the first quarter of a graduate probability course is devoted to developing these results, but since they are not essential for what we will do, we will just prove one sample result to illustrate some of the concepts in this section and then state two more results in order to give you a taste of the theory. (4)
Let X1, Xz, ... be a sequence of independent and identically distributed random
variables (i.e., P(X; < x) is independent of i) that have EXz < oc. As n -, oo, 1(X1 n
+
EX1 in probability.
+ Xn)
Proof Letµ=EX;, Y=X; -p. Now n(X1+...
+XX)_P=n(Y1+... +Y),
so it suffices to show that the right-hand side - 0 in probability. The key to doing this is to observe that
)2(n n =E
E
YY
Y
i=1j=1
i=1 n
_
n
n
EY Y = > t=1j=1
(since if i
nEY1
i=1
j, EY Yj = EYEY, = 0). If we let S. = Y1 +
then
ES = Cn,
i.e.,
S" z
E
n
= C n
and Chebyshev's inequality implies that
/ Sn
PI
//
1\z
>E <E-LEIS"I =CE
\n- -
n
-c
n
->O.
+ Yn and C = EYz,
A.2
299
todepeadence and Lows of Large Numbers
The result above is a weak form of the weak law of large numhcrs. The strongest form of the strong law is (5)
Let X1, X2, ... be a sequence of i.i.d. random variables (for a translation, see (4)) with EI Xj I < co. As n
n (X1 + ... + Xn) -, EX1
oc,
a.s.
Analysts may recognize this result as being a consequence of Birkhoff's ergodic theorem. It is easy to show (see Exercise 2 below) that EI X; I < oo is necessary + X,)/n to converge to a finite limit so the condition in (5) is for (X1 + "sharp." The weak law holds in a little greater generality: (6)
Let X1, X2, ... be a sequence of i.i.d. random variables. There is a sequence of constants an such that 1
n
+ Xn) - a - 0 in probability
(Xl +
if and only if nP(I XiI > n) - 0. In this case, we can take an = E(Xi ; I Xi 1 < n).
Proof We leave the proof as an exercise for the reader. Once someone WIN you to look at Xi,n = Xi I(IXil,,n) and use Chebyshev's inequality, the rest is not hard. See Feller (1971), page 235, for a solution. The next two exercises are much easier, but the first is much more important.
The Second Borel-Cantelli Lemma. If A1, A2, ... are independent and >P(An) = co, then P(An i.o.) = 1. Exercise 1
Hint:
/
PI
If M:5 N < oo, then
\
A I = [1 (1 - P(An)),
n n/ =M
and, for any M, the right-hand side - 0 as N
co.
Exercise 2 Let Xl, X2, ... be a sequence of i.i.d. random variables with EI X,I = oo. Then
P(ISn-S.+1I>ni.o.)1, so Sn/n cannot converge to a finite limit on a set of positive probability. Hint : 00
n=0
W
P(IX;I>n)> fo P(IXI>x)dx=El XI.
300
Appendix A Primer of Probability Theory
A.3 Conditional Expectation Given a probability space (52, .to, P), a a-field F c .moo, and a random variable
Xe Fo (i.e., FO measurable) with EI XI < oo, we define E(XI.F) to be any random variable Y that has (1) Ye.F
(ii) for all A e .F, f
dP = fA YdP.
A
Y is said to be a version of E(XI fl. Any two versions of E(XI -F) are equal almost surely.
Interpretation: We think of .t as describing the information we have at our disposal : For each A e F, we know whether or not A has occurred. E(X I.F) is, then, our "best guess" of the value of X given the information at our disposal. Some examples should help to clarify this. In each case, you should check that the answer we have given satisfies (i) and (ii).
Example l If X e .F, then E(X I.F) = X, that is, if we know X, then our "best guess" is X itself. In general, the only thing that can keep X from being E(XIF) is condition (i). Example 2
Suppose that 521, ... , n. is a partition of S2 into disjoint sets, each
of which has positive probability, and that F = a(521, ... , 52 ), the a-field generated by these sets. Then on 52;,
E(XI F) = E(X ; f1i) P(52i)
In words, the information in F tells us the element of the partition which contains our outcome, and given this information, our best guess for X is the average value of X over 52;.
Example 3 Suppose that X is independent of F, that is, P({X e Al fl B) =
P(XeA)P(B) for all Ae. and Be.F. In this case, E(XI,fl = EX, that is, the information in .F is of no help in guessing the value of X. Let X > 0 and let Q be the measure that has density X with respect to P, that is, Q (A) = JA XdP. Let P' and Q' be the restrictions of P and Q to .F. Then Q' << P', and E(X I.) is the Radon-Nikodym derivative dQ'/dP'. This is, in fact, how we show that the conditional expectation exists. Conditional expectation has many of the same properties as ordinary Example 4
expectation :
(a) linearity E(aX + YI °F) = aE(XI F) + E(YI F )
A.3
Conditional Expectation
3111
(b) order preserving
if X < Y, then E(XI.F) < E(Y13) (c) monotone convergence if X T X, then E(X I.°F) T E(XI F)
(d) Jensen's inequality if cp is convex and EI XI, EI gp(X)I < oe, then
q(E(XI.F)) < (e) LP convergence
if X. - X in LP, p > 1, then E(XI _F) in LP
(f) dominated convergence if X -> X a. s. and Y with EY < co, then E(XXI.t) -), E(XI f) a.s.
These properties are not very hard to prove using the definition of conditional expectation. To prove (a), we simply check that the right-hand side a .F and has the same integral as aX + Y over all A e .F. To prove (b), we observe that if A e .F,
f fA XdP< SA and applying this result to A = {E(XIfl > JA
A
we conclude that A = 0,
that is, E(XI.F) < E(YI). With these two examples as a guide, you should be able to prove (c) through
(e), but, for the moment, I want to discourage you from doing this. It is more important to understand the following properties of conditional expectation, which have no analogue for ordinary expectation, and I leave the proofs of these properties as recommended exercises. E(E(Xj
)) = E(X).
If F c .F2, then (i) E(E(XI.°,Fi) I Fz) = E(X I.°wi) (ii) E(E(XI _O--z)I.Fj) = E(XI.F1).
In words, the smaller a-field always wins out. (3)
IfAe4andEIYI
By using linearity and taking limits, we can easily extend this to
302
Appendix A Primer of Probability Theory
(4)
If XE!N and El YI, El XYI < oo, then
E(XYI 9) = XE(YI 9).
From (4) (and (d)), we get a geometric interpretation of the conditional expectation. (5)
Suppose that EX 2 < oo. Then L2 (Fo)
y C .moo
: EY2 < oo } is a Hilbert space,
and L2(5F) is a closed subspace. In this case, E(XI.9F) is the projection of X
onto L2(3F), or, in statistical terms, E(XI.fl is the random variable YE.F that minimizes the mean square error E(X - Y)2.
A.4 Martingales Let ffl be an increasing sequence of a-fields. A sequence X. is said to be adapted for all and X E(Xn+1 I.° ) = X is said to be a martingale. If, in the last definition,
to J if
= is replaced by < or > , then X. is said to be a supermartingale or submartingale, respectively.
To give an example of a martingale, consider successive tosses of a fair coin, and let = 1 if the nth toss is heads and = -1 if the nth toss is tails.
Let S. = 1 +
+ . S is called the (symmetric) simple random walk. It
represents the amount of money a gambler has after the nth toss if each time the coin is tossed he bets one dollar on the coin coming up heads. Let # = As we mentioned before, we think of J as giving the information we have at time n, which in this case is the outcomes of the first n tosses. S To prove this, we observe that S. = 1 + is a martingale with respect to oo, and, by (3) in Section A.2, n+1 is independent of + we have
/ + Ee(Sn+1 I .F.) = E(S
KK
bn+1 l
E(SS I Win) +
I .F.)
If the successive tosses have P( = 1) < 1/2, then the computation above shows that
S corresponds to betting on an unfavorS is able game, we see that there is nothing "super" about a supermartingale. The
name comes, instead, from the fact that superharmonic functions, when composed with Brownian motion, give rise to supermartingales. For what follows, we will need a few simple facts about martingales, the proofs of which we leave as exercises for the reader.
A.5 Gambling Systems and the Martingele Convergence Theorem
303
(1)
If X is a martingale and m < n, then
(2)
If X is a martingale and T is a convex function with El (p(X,)l < rfj for nil n, then qp(X,) is a submartingale.
(3)
If X,, is a submartingale and qp is an increasing convex function with E(p(X,) < CZ)
for all n, then (p(X,) is a submartingale. (4)
Orthogonality of Martingale Increments. If X. is a martingale with EX,2 < 00 for all n and V!9 m 5 n, then
E((Xn-Xm)Xi)=0. Since the proof of (4) is a classic example of manipulating conditional expectations, we will give the proof and let the reader justify the steps
E((X, - Xm)Xi) = EE((X, - Xm)XiI&m = E[X,E(X, - XmI `fm)] = E[Xi(E(Xnl
Xm)]
=0. From the proof above, it is immediate that we have (5)
Under the hypotheses of (4), E((Xn - Xm)ZIFm) = E(X,
X.2.
A.5 Gambling Systems and the Martingale Convergence Theorem Let ? be an increasing sequence of o-fields. H. is said to be predictable if for all n > 1. In words, the value of the process at time n may be predicted (with certainty)
from the information available at time n - 1. You should think of H. as the amount of money a gambler bets at time n. This amount can be based on the outcomes at times 1, ... , n - 1, but it cannot depend on the outcome at time n. Once we start thinking of H. as a gambling system, it is natural to ask how much money we would win if we used it. For concreteness, let us suppose that the game consists of flipping a coin and that for each dollar bet we win one dollar when the coin comes up heads and lose one dollar when the coin comes up tails (most games in casinos reduce to this situation when you ignore all the ritual). Let S,, be the net amount of money we would have won at time n if we had bet one dollar each time. If we bet according to a gambling system H, then our net winnings at time n would be
304
Appendix
A Primer of Probability Theory
n
(H. S)n = Y Hm(Sm - Sm-1), m=1
since S. - Sm_1 = + I or -1 when the mth toss results in a win or loss, respectively.
-
The next result is the most basic fact about gambling systems and is, apparently, little known. It says that there is no system for beating an unfavorable game. (1)
Let Xn be a supermartingale. If H,, >- 0 is predictable and each H is bounded, then (H- X)n is a supermartingale.
Proof (H- X). + E(Hn+1(X,,+1 - Xn)I` n)
E((H.
= (H' X),, + Hn+1E(Xn+1 (1-1 -
since E(Xn+1 -
0 and Hn+, > 0.
Remark: The same result is obviously also valid for submartingales and for martingales (and in the second case without the restriction H. > 0). To keep from being repetitious, we will state our results for only one type of process and leave it to the reader to translate the result to the other two. In my remark preceding (1), I did not mean to dismiss gambling systems as worthless. There is one system that allows us to prove the martingale convergence theorem. Let h > 0, let No = 0, and for k >- 1 let N2k_1 = inf{m > N2k_2 : X. < a}
N2k=inf{m> N2,_1:Xm?a+h} if N2k_1 < m - 1 < N2k for some k otherwise U. = sup{k: N2k < n}.
Hm
11 0
Since X(N2k-1) < a and X(N2k) > a+ h, between times N2k_1 and N2k, X crosses from < a to > a + h. H. is a gambling system that tries to take advantage
of the "upcrossings." In stock-market terms, we buy when Xm < a and sell when Xm > a + h. In this way, every time an upcrossing is completed, we make
a profit
h. Last but not least, U. is the number of upcrossings completed
by time n. (2)
The Upcrossing Inequality. If X. is a submartingale, then
hEU, < E(X - a) + - E(Xo - a) Proof Let Y. = (Xn - a)+. Since Y is a submartingale that upcrosses [0,h] the same number of times that X,, upcrosses [a, a + h], it suffices to prove the result when a = 0 and Xn > 0. In this case, we have that
A.5 Gambling Systems and the Martingele Convergence Theorem
35
hU" S (H X)",
since a final incomplete upcrossing (if there is one) makes a nonncgolivc contribution to the right-hand side. Let Km = 1 - Hm.
and it follows from (1) that E(K X)" > E(K X)o = 0, so E(H X)" S E(X" Xo), proving (2) in the special case and, hence, in general.
Remark: We have proved the result above in its classical form even though this approach is a little misleading. The key fact is that E(K X)" > 0, that is, even by buying high and selling low, we cannot lose money on a submartingale,
or in other words, it is the reluctance of submartingales to go from above a + h to below a that limits the number of upcrossings. From the upcrossing inequality, we easily obtain (3)
The Martingale Convergence Theorem. If X. is a submartingale with sup,, EX < oo, then as n -> oo, X. converges almost surely to a limit X with El XI < oo.
Proof Since (X - a)+ < X + + IaI, (2) implies that EU" 5 (IaI + EIX"I)/h, so if we let U = lim U,, be the number of upcrossings of [a, a + h] by the whole sequence, then EU < co and, hence, U < oo a.s. Since this result holds for all rational a and h,
U {lim inf X" < a < a + h < lim sup X" a,hEQ
rm
n-m
has probability 0, and lim sup X" = lim inf X"
n-
n,ao
a.s.,
which implies X = limX" exists a.s. Fatou's lemma guarantees that EX+ S
lim inf EX < oo, so X < oc a.s. To see that X > - oo, we observe that EX,- = EX - EX" < EX,, - EXo (since X,, is a submartingale), and another application of Fatou's lemma shows that EX - < lim inf EX,,- < oo. From (3), it follows immediately that we have (4)
If X" < 0 is a submartingale, then as n -+ oo, X"
X a.s.
The last two results are easy to rationalize. Submartingales are like increasing sequences of real numbers-if they are bounded above, they must converge almost surely. The next example shows that they need not converge in L1.
Example 1 (Double or nothing) Suppose that we are betting on a symmetric simple random walk and we use the following system: H" _
2n-1
on S,,-, = n - 1
0
otherwise.
3%
Appendix
A Primer of ProbablUty Theory
In words, we start by betting one dollar on heads. If we win, we add our winnings to our original bet and bet everything again. When we lose, we lose everything
and quit playing. Let X" = 1 + (H- S)n. From (1), it follows that X is a martingale. The definition implies that X. > 0, so EI XnI = EX. = EXo = 1, but it is easy to see that P(X" > 0) = 2-", so X" -. 0 a.s. as n - oo. This is a very important example to keep in mind as you read the next three sections.
A.6
Doob's Inequality, Convergence in L", p > 1 A random variable N is said to be a stopping time if {N = n} e .F" for all n < oo.
If you think of N as the time a gambler stops gambling, then the condition above says that the decision to stop at time n must be measurable with respect to the information available at that time. The following is an important property of stopping times : (1)
If X. is a submartingale and N is a stopping time, then XfAN is a submartingale.
Proof Let Hn = 1(N> n). Since {N > n } = {N< n - 1 }` E and it follows from (1) of the last section that (H - X). = (2)
Hn is predictable, is a submartingale.
If X. is a submartingale and N is a stopping time with P(N < k) = 1, then EXo < EXN < EXk .
Proof Since XNnn is a submartingale, it follows that EXo = :!g EXNnk = EXN. To prove the right-hand inequality, let H. = 1(N<m). Now {N < m} = {N < m - 11 so (H X)n = Xn - XN,n is a submartingale and EXk EXN = E(H X)k > E(H X )O = 0. Remark: Let Xn be the martingale described in the last section and let N = inf In: X,, = 0}. Then EX, = 1 > 0 = EXN, so the first inequality need not hold for unbounded stopping times. In Section A.8, we will consider conditions that guarantee that EXo < EXN for unbounded N. From (2), we immediately get (3)
Doob's Inequality. If X. is a submartingale and A = 10<M
).P(A) < EX. lA < EX. .
Proof LetN=inf{m:Xm>Aorm>n}.XN>AonAandN=nonA`,so it follows from (2) that ).P(A) < E(XN 1A) < EX. lA
(observe that XN = Xn on A`). The second inequality in (3) is trivial.
A.7
Uniform IntearebWty and Convergence In L'
307
Integrating the inequality in (3) gives (4)
If X,, = max Xm, then for p > 1, OSmSn
p
E(X,°) <
P 1
E(X ) P.
p
Proof Fubini's theorem implies that EX,? = f 'pA'-1P(X,, > A) dl
<
=
0
JP2P' (2_i fn
X
J X,,2x}
X dPl di
/
(iox
n PAP- 2 d2 dP
f
P
P-1 fn If we let q = p/(p - 1) be the exponent conjugate to p and apply 11Older's inequality, we see that the above < q(EI X,. I P)1/P(EIXnIP)1/4
At this point, we would like to divide both sides of the inequality uhovc by (EIX,,IP)1/q to prove (4). Unfortunately, the laws of arithmetic do not allow us to divide by something that may be oo. To remedy this difficulty, we observe that P(X,, A N > A) < P(X,, > A), so repeating the proof above shows that (E(X, A N)") "P -< q(EI X,, I P) 1/p,
and letting N -+ oc proves (4). From (4), we get the following LP convergence theorem: (5)
If X,, is a martingale, then for p > 1, sup,, EI X. I P < oo implies that X,, - X in LP.
Proof From the martingale convergence theorem, it follows that X,, - X a.s. Since I X I is a submartingale, (4) implies that supra I X. I e LP, and it follows from the dominated convergence theorem that E I X. - X I P -+ 0.
Remark: Again, the martingale described at the end of the last section shows that this result is false for p = 1.
A.7 Uniform Integrability and Convergence in L1 In this section, we will give conditions that guarantee that a martingale converges
in L'. The key to this is the following definition:
30$
Appendix A Primer of Probability Theory
A collection of random variables {Xi, i e I } is said to be uniformly integrable if lim (supE(IX;I ;IX;I > M)I =0.
\ieI
M- O
J
Uniformly integrable families can be very large. (1)
If X e L', then {E(XI.fi) } is uniformly integrable.
0, then the dominated converProof If A. is a sequence of sets with gence theorem implies that E(I XI ; Aj 0. From the last result, it follows that if e > 0, we can pick S > 0 such that if P(A) < S, then E(I XI ; A) < e.
Pick M large enough so that EI XIIM < S. Jensen's inequality and the definition of conditional expectation imply that
E(JE(XI.F) JE(XI.F) > M) < E(I XI ; E(IXI I.F)
> M),
and we have that P(E(I XI I _F) > M)
< E(E(IMI I.f )) = EIXI < s,
so for this choice of M,
suupE(IE(XI.9)I ; E(XI.f)I > M) < e, and since a was arbitrary, the collection is uniformly integrable. Another common example is Exercise 1
Let (p be any function with 9(x)lx
oo as x - oo, for example,
qp(x)=x" or cp(x)=xlogx. If Eq(IXi1)
If X -. X a.s., then the following statements are equivalent: (a) {X,,, n > 0} is uniformly integrable
(b) X. - X in L' (c) EIXXI
Proof (a)
EIXI < oo.
(b) : Let
M 9M(x)
M<x
X -M<x<M
-M
x < -M
and observe that by patiently checking the nine possible cases,
<EI(pM(X.)-gM(X)1 As n
oo, the first term
M).
0. If e > 0 and M is large, then the second term < e.
A.8
:9
Optional Stopping Theorems
To bound the third term, we observe that uniform integrability implieN that that EIXI < sup. EIXI < oo, M the
that
sup El X
E.
- XI < 2e, proving (b).
the
(b)=(c): EIXI - EIXU :!9 El IX.1-IXI <EIX -XI. (c) =>(a): Let /IM(x) = x on [0, M - 1], OM = 0 on [M, oo), and let dIM he
linear on [M - 1, M]. If M is large, EIXI - Eom(I XI) < e/2. The bounded convergence theorem implies that E//M( 1 X I) - EI//M(X I ), so using (c) we get that if n > no, E(IXXI ; IXI > M):!5; EI X1 - E,M(Xfl) < E.
By choosing M larger, we can make E(I XI ; I XI > M) < r: for 0 < n < no, so X is uniformly integrable.
We are now ready to state the main theorems of this section. Since we have already done all the work, the proofs are short. (3)
For a submartingale, the following statements are equivalent:
(a) it is uniformly integrable (b) it converges in L'.
Proof (a) => (b): Uniform integrability implies that supElXl < ou, which by (3) of Section A.5 implies that X -> Xa.s., which by (2) above implies that X -> X in L'. The converse, (b) (a), is a corollary of (2). (4)
For a martingale, the following statements are equivalent:
(a) it is uniformly integrable (b) it converges in L' (c) there is an integrable random variable X such that X = E(XI..). Proof (a) (b) => (c) :
(b) : This result follows from the proof given in (3). Let X = lim X,,. If m > n, then E(Xm I-F) = X,,, so if A e .F,,,
E(X ; A) = E(Xm ; A). As m - oo, Xm IA --> X1A in L1, so we have that E(X ; A) _
E(X, A) for all A e 3, and it follows that X. = E(Xl S ). (c) (a): This result follows from (1) above.
A.8 Optional Stopping Theorems In this section, we will prove a number of results that allow us to conclude that if X. is a submartingale and M :!g N are stopping times, then EXM < EXN. The first step is to show (1)
If X. is a uniformly integrable submartingale, then for any stopping time N, XNAf is uniformly integrable.
310
Appendix A Primer of Probability Theory
Proof We begin by observing that XNnn is a submartingale with EXlvnn < EX
(since X is a submartingale), so sup EX,',,, < sup EX.' < oo n
(since/Xn - X. in L1), and it follows from the martingale convergence theorem that XNAn -+ XN a.s. and EI XNI < oo. With this result established, the rest is easy, since M) < E(I XNI ; I XNI > M) + E(IXnI ; IXnI > M)
E(IXNnnI ;
(2)
and Xn is uniformly integrable. From (1) it follows immediately that we have: If Xn is a uniformly integrable submartingale, then for any stopping time N< oo,
EXo < EXN < EX..
Proof (2) in Section A.6 implies that EXa < EXNnn 5 EX.. Letting n - oo and observing that (1) above and results in Section A.7 imply that XNAf - XN in L' and X. - X. in L1, gives the desired result. From (2), we get the following useful corollary: (3)
The Optional Stopping Theorem. If XNAf is a uniformly integrable submartingale,
then for any stopping time M < N, EXM < EXN.
We have given (3) a name since it is the basic result in this section and it is usually the result we are referring to when we use the words "optional stopping theorem" in the text. In applying (3), the following fact is useful: (4)
If Xn is a submartingale and N < co is a stopping time with (a) EIXNI < oo (b) E(I XXI ; N > n) - 0, then XNnn is uniformly integrable.
Proof E(IXNnnI; IXNnnI > M) <_ E(IXNI;IXNI > M) + E(IXXI;N> n, IXnI > M).
Let e > 0. If we pick no large enough, then E(IXnI ; N > n) < e/2 for all n > no. Having done this, we can pick M large enough so that E(I XNI ; I XNI > M) < E/2
and, for 0:!9 n :!g no, E(IXnI; IXnI > M) < e/2, so it follows from the first inequality that E(I XNnnI ; IXNnnI
> M) < E
for all n, and hence XN.,, is uniformly integrable.
Finally, there is one stopping theorem that does not require uniform integrability :
A.8
(5)
Options[ Stopping Theorem.
311
If Xn is a nonnegative supermartingale and N.-!5; oc is a stopping time, then EXo Z EXN, where X. = lim Xn (which exists by (4) of Section A.5). Proof EXo > EXNAf. The monotone convergence theorem implies that E(XN, N < oo) = lim E(XN ; N< n), n- w
and Fatou's lemma implies that E(XX ; N = oo) <_ lim inf E(XX ; N > n), n-m
so adding the last two lines gives the desired result.
References
In the list of references below, we have for convenience shortened the standard abbreviations for the sources referred to most often : AMS BAMS
CMP CPAM JFA LMS
LNM MAA PAMS
PJM Sem.
TAMS TPA
ZfW
American Mathematical Society Bulletin of the American Mathematical Society Communications on Mathematical Physics Communications on Pure and Applied Mathematics Journal of Functional Analysis London Mathematical Society Lecture Notes in Mathematics, Springer-Verlag, New York Mathematical Association of America Proceedings of the American Mathematical Society Pacific Journal of Mathematics Seminar de Probabilites (Strasbourg) Transactions of the American Mathematical Society Theory of Probability and Its Applications Zeitschrift fur Wahrscheinlichkeitstheorie and Verwandte Gebiete
Agmon, S. (1965). Lectures on elliptic boundary value problems. Van Nostrand, New York. (1982). Lectures on exponential decay of solutions of second-order elliptic equaliona, Mathematical Notes. Princeton Univ. Press, Princeton. , Ahlfors, L. (1966). Complex analysis. 2d ed. McGraw-Hill, New York.
Aizenman, M., and B. Simon (1982). Brownian motion and a Harnack inequality for Schrodinger operators. CPAM 35:209-273. Ash, J. M., ed. (1976). Studies in harmonic analysis. MAA Studies in Mathematics, vol, 13, MAA, Washington, D.C.
Austin, D. G. (1966). A sample function property of martingales. Ann. Math, Slat, 37:1396-1397. Bachelier, L. (1900). Theorie de las speculation. Ann. Sci. Ecole Norm. Sup. 17:21-86, Baernstein, A., II (1978). Some sharp inequalities for conjugate functions. Indiana Math, J, 27:833-852. 313
314
Reference.
Bary, N. K. (1964). A treatise on trigonometric series. Pergamon Press, New York. Bernard, A., and B. Maisoneuve (1977). Decomposition atomique de martingales de la classe H1. In Sem. XII, pp. 303-323. Springer LNM 649. Bers, L., F. John, and M. Schechter (1964). Partial differential equations. AMS, Providence, RI. Berthier, A., and B. Gaveau (1978). Critere de convergence des fonctionelles de Kac et applications en mechanique et en geometrie. JFA 29:416-424. Billingsley, P. (1968). Convergence of probability measures. John Wiley, New York. Blumenthal, R. M., and R. K. Getoor (1968). Markov processes and their potential theory. Academic Press, New York. Breiman, L. (1968). Probability. Addison-Wesley, New York. Brelot, M., and J. L. Doob (1963). Limites angulaires et limites fines. Ann. Inst. Fourier (Grenoble) 13:395-415. Brossard, J. (1975). These de troisieme cycle. (1976). Comportement nontangentiel et comportement Brownien des fonctions
harmonique dans un demi-espace: Demonstration probabiliste d'un theoreme de Calderon et Stein. In Sem. XII, pp. 378-397. Springer LNM 649. Burkholder, D. L. (1962). Successive conditional expectation of an integrable function. Ann. Math. Stat. 33:887-893. (1964). Maximal inequalities as necessary conditions for almost everywhere convergence. ZfW 3:75-88. (1966). Martingale transforms. Ann. Math. Stat. 37:1494-1505. (1970). Martingale inequalities. In Martingales, ed. H. Dinges, pp. 1-8. Springer LNM 190. (1973). Distribution function inequalities for martingales. Ann. Prob. 1: 19-42. (1975). One-sided maximal functions and H°. JFA 18:429-454. (1976). Harmonic analysis and probability. In Studies in harmonic analysis, ed. Ash (1976), pp. 136-149. (1977a). Brownian motion and classical analysis. Proc. Symp. Pure Math. 31:5-14. (1977b). Exit times of Brownian motion, harmonic majorization and Hardy spaces. Adv. in Math. 26:182-205. (1978). Boundary value estimation of the range of an analytic function. Michigan Math. J. 25:197-211. (1979a). Martingale theory and harmonic analysis in Euclidean spaces. In Harmonic analysis in Euclidian space, ed. Weiss and Wainger (1979). (1979b). A sharp inequality for martingale transforms. Ann. Prob. 7:858-863.
(1981). A geometrical characterization of Banach spaces in which martingale difference sequences are unconditional. Ann. Prob. 9:997-1011. (1982). A nonlinear partial differential equation and the unconditional constant of the Haar system in L. BAMS (New Series) 7:591-595. (in press). Boundary value problems and sharp inequalities for martingale transforms. Ann. Prob.
Burkholder, D. L., B. Davis, and R. F. Gundy (1972). Integral inequalities for convex functions of operators and martingales. In Proceedings of the Sixth Berkeley Symposium, vol. II, pp. 223-240. Burkholder, D. L., and R. F. Gundy (1970). Extrapolation and interpolation of quasilinear operators on martingales. Acta Math. 124:249-304.
(1972). Distribution function inequalities for the area integral. Studia Math. 44:527-544.
References
315
(1973). Boundary behavior of harmonic functions in a half space and Hriiwnion motion. Ann. Inst. Fourier (Grenoble) 23:195-212. Burkholder, D. L., R. F. Gundy, and M. L. Silverstein (1971). A maximal function characterization of the class H5. TAMS 157:137-153. Calderon, A. P. (1950a). On the behavior of harmonic functions at the boundary. 7'A MN 68:47-54. (1950b). On a theorem of Marcinkiewicz and Zygmund. TAMS 68:55.61. (1950c). On the theorems of M. Riesz and Zygmund. PAMS 1:533-535. (1966). Singular integrals. BAMS 72:426-465. Calderon, A. P., and A. Zygmund (1964). On higher gradients of harmonic functions, Studia Math. 24:211-226. Cameron, R. H., and W. T. Martin (1944a). Transformation of Wiener integrals under translations. Ann. Math. 45:386-396. (1944b). The Wiener measure of Hilbert neighborhoods in the space of real continuous functions. J. Math. Phys. 23:195-209. (1945a). Evaluations of various Wiener integrals by use of Sturm-Liouville differential equations. BAMS 51:73-90. (1945b). Transformation of Wiener integrals under a general class of linear Ironsformations. TAMS 58:184-219. (1949). The transformation of Wiener integrals by nonlinear trunsformntions. TAMS 66:253-283. Carleson, L. (1958). An interpretation problem for bounded analytic functions. Amer. J. Math. 80:921-930. (1961). On the existence of boundary values for harmonic functions in several variables. Archiv fur Math. 4:393-399. (1962). Interpolations by bounded analytic functions and the corona prohlem. Ann. Math. 76:547-559. Carmona, R. (1978). Pointwise bounds for Schrodinger eigenstates. CMP 62:97 106. (1979a). Processus de diffusion gouvern6 par la forms de Dirichlet de l'oporulcur tie Schrodinger. In Sem. XIII, pp. 557-569. Springer LNM 721.
(1979b). Regularity projects of Schrodinger and Dirichlet semigroups. JI"A 33:259-296. Carmona, R., and B. Simon (1981). Pointwise bounds on eigenfunctions and wave packots in n-body quantum systems V. CMP 80:59-98. Chao, J. A., and M. H. Taibleson (1973). A subregularity inequality for conjugate systems on local fields. Studia Math. 46:249-257. Chung, K. L. (1974). A course in probability theory. 2d ed. Academic Press, New York. (1976). Excursions in Brownian motion. Archiv fur Math. 14:155-177. (1979). On stopped Feynman-Kac functionals. In Sem. XIV, pp. 347-356. Springer LNM 784. (1981). Feynman-Kac functional and the Schrodinger equation. In Seminar on Stochastic Processes I, ed. E. Cinlar, K. L. Chung, and R. Getoor. Birkhauser, Boston. (1982). Lectures from Markov processes to Brownian motion. Springer-Verlag, New York. Chung, K. L., R. Durrett, and Z. Zhongxin (1983). Extension of domains with finite gauge. Math. Annalen 264:78-79. Chung, K. L., and P. Li (1983). Comparison of probability and eigenvalue methods for the Schrodinger equation. Adv. in Math. Chung, K. L., and K. M. Rao (1980). Sur la theorie du potentiel avec la fonctionelle de
316
Reference.
Feynman-Kac. C. R. Acad. Sci. Paris 29OA: 629-63 1. (1981). Potential theory with the Feynman-Kac functional. Prob. Math. Stat. 1. Chung, K. L., ahd S. R. S. Varadhan (1979). Positive solutions of the Schrodinger equation in one dimension. Studia Math. 68:249-260. Ciesielski, Z., and S. J. Taylor (1962). First passage times and sojurn times for Brownian motion in space and exact Hausdorff measure. TAMS 103: 434-450. Coifman, R. R. (1974). A real variable characterization of H". Studia Math. 51:269-274. Coifman, R. R., and C. Fefferman (1974). Weighted norm inequalities for maximal functions and singular integrals. Studia Math. 51:241-250.
Coifman, R. R., and R. Rochberg (1980). Another characterization of BMO. PAMS 79:249-254. Coifman, R. R., and G. Weiss (1977). Extensions of Hardy spaces and their use in analysis. BAMS 83:569-645. Conway, J. B. (1978). Functions of one complex variable. 2d ed. Springer-Verlag, New York. Courant, R., and D. Hilbert (1953, 1962). Methods of mathematical physics. Vols. I, II. Interscience, New York. Davis, B. (1968). Comparison tests for martingale convergence. Ann. Math. Stat. 39:21412144.
(1969). A comparison test for martingale inequalities. Ann. Math. Stat. 40:505-508.
(1970). On the integrability of the martingale square function. Israel J. Math. 8:187-190. (1973a). On the distributions of conjugate functions of nonnegative measures. Duke Math. J. 40:695-700. (1973b). An inequality for the distribution of the Brownian gradient function. PAMS 37:189-194. (1974). On the weak type (1, 1) inequality for conjugate functions. PAMS 44:307311.
(1975). Picard's theorem and Brownian motion. TAMS 213:353-362. (1976). On Kolmogorov's inequalities 11a < Coll f 111, 0 < p < 1. TAMS 222:179192.
(1979a). Applications of the conformal invariance of Brownian motion. In Harmonic analysis in Euclidean space, ed. Weiss and Wainger (1979). (1979b). Brownian motion and analytic functions. Ann. Prob. 7:913-932. (1980). Hardy spaces and rearrangements. TAMS 261:211-233. Deift, P., W. Hunziker, B. Simon, and E. Vock (1978). Pointwise bounds on eigenfunctions
and wave packets in n-body quantum systems IV. CMP 64:1-34. Dellacherie, C. (1980). Un survol de la theorie de l'integrale stochastique. Stoch. Proc. Appl. 10:115-144. Dellacherie, C., and P. A. Meyer (1978). Probabilities and potential. English translation of Probabilites et potentiel. North Holland, Amsterdam. Doleans-Dade, C. (1970). Quelques applications de la formule de changements de variables pour les semimartingales. ZJW 16:181-194.
Donsker, M. D., and M. Kac (1950). A sampling method for determining the lowest eigenvalue and the principal eigenfunction of Schrodinger's equation. J. Res. Nat. Bureau of Standards 44:551-557. Donsker, M. D., and S. R. S. Varadhan (1974). Asymptotic evaluation of certain Wiener integrals for large time. In Proceedings of the International Conference on Function Space Integration. Oxford Univ. Press. (1975a). Asymptotics for the Wiener sausage. CPAM 28:525-566; errata, p. 677.
Rd.resoes
317
(1975b). On a variational formula for the principal eigenvaluc fur upeuiturs with a maximum principle. Proc. Nat. Acad. Sci., USA 72:780-783. (1975-1976). Asymptotic evaluation of certain Markov process exhrrlntwns for large time. CPAM (I) 28:1-47; (II) 28:279-301; (III) 29: 389-461. (1976). On the principal eigenvalue of second-order elliptic differontlal operators. CPAM 29:595-622. Doob, J. L. (1953). Stochastic processes. John Wiley, New York. (1954). Semimartingales and subharmonic functions. TA MS 77:86 121 (1955a). Martingales and one-dimensional diffusion. TA MS 78:168 208. (1955b). A probability approach to the heat equation. TAMS 80:216 280.
(1956). Probability methods applied to the first boundary value problem. In Proceedings of the Third Berkeley Symposium, vol. II, pp. 49 NO.
(1957). Conditional Brownian motion and the boundary limits of harmonic functions. Bull. Soc. Math. France 85:431-458. (1958a). Boundary limit theorems for a half-space. J. Math, Pure-,% Appl. 37:385392.
(1958b). Probability theory and the first boundary value prohlciii Illinois J. Math.
2:19-36. (1959). A nonprobabilistic proof of the relative Fatou theorem ,4nn. Inst. FourIa (Grenoble) 9:293-300. (1960). Relative limit theorems in analysis. J. Analyse Math. 8 2H) 306. (1961). Conformally invariant cluster value theory. Illinois J. Moth. 5:521-549. (1962). Boundary properties of functions with finite Dirichict integrals. Ann. Inst, Fourier (Grenoble) 12: 573-622. (1963). One-sided cluster-value theorems. Proc. LMS 13:461 470. (1964). Some classical function theory theorems and their modern versions. Ann, Inst. Fourier (Grenoble) 15:113-136.
(1966). Remarks on the boundary limits of harmonic functions. J. SIAMNumerical Analysis 3:229-235. Dubins, L. E., and D. Gilat (1978). On the distribution of maxima of mariingales. PAMS 68:337-338. Dubins, L. E., and G. Schwarz (1965). On continuous martingales. Proc. Nat. Acad. Sci., USA 53:913-916.
Dunford, N., and J. T. Schwarz (1957). Linear operators, Part I: General theory. Interscience, New York. Duren, P. L. (1970). The theory of H" spaces. Academic Press, New York. Duren, P. L., B. W. Romberg, and A. L. Shields (1969). Linear functionals on HP spaces with 0 < p < 1. J. Reine Angew. Math. 238: 32-60.
Durrett, R. (1982). A new proof of Spitzer's result on the winding of two-dimensional Brownian motion. Ann. Prob. 10:244-246. Dvoretsky, A., P. Erdos, and S. Kakutani (1961). Nonincreasing everywhere of the Brownian motion process. In Proceedings of the Fourth Berkeley Symposium, vol. II, pp. 103-116. Dynkin, E. B. (1960). Markov processes and related problems of analysis. Russian Math. Surveys 15, no. 2, pp. 1-24. (1963). Markov processes and problems in analysis. In Proceedings of the International Congress, Stockholm, 1962. AMS Translations, series 2, vol. 31, pp. 1-24, (1981). Markov processes and related problems of analysis Selected papers. LMS Lecture Notes. Cambridge Univ. Press, Cambridge.
318
References
Einstein, A. (1905). On the movement of small particles suspended in a stationary liquid demanded by the molecular kinetic theory of heat. Ann. Phys. 17. (1926). Investigations on the theory of the Brownian movement. Reprinted by Dover Books, New York, 1956. Fatou, P. (1906). Series trigonometriques et series de Taylor. Acta Math. 30: 335-400. Fefferman, C. (1971). Characterization of bounded mean oscillation. GAMS 77:587-588. (1976). Harmonic analysis and H' spaces. In Studies in harmonic analysis, ed. Ash (1976), pp. 38-75. Fefferman, C., and E. M. Stein (1972). H° spaces in several variables. Acta Math. 129:137193.
Feller, W. (1971). An introduction to probability theory and its applications. Vol. II. John Wiley, New York. Feynman, R. J. (1948). Space-time approach to nonrelativistic quantum mechanics. Rev. Mod. Phys. 20:367-387. Folland, G. B. (1976). Introduction to partial differential equations. Princeton Univ. Press, Princeton. Freedman, D. (1970). Brownian motion and diffusion. Holden-Day, San Francisco. Friedman, A. (1964). Partial differential equations of parabolic type. Prentice-Hall, Englewood Cliffs, NJ. (1969). Partial differential equations. Holt, Rinehart & Winston, New York. (1975). Stochastic differential equations and applications. Academic Press, New York. Garnett, J. (1979). Two constructions in BMO. In Harmonic analysis in Euclidean space, ed. Weiss and Wainger (1979). (1980). Bounded analytic functions. Academic Press, New York. Garnett, J., and P. Jones (1978). The distance in BMO to L'. Ann. Math. 108:373-393. (1982). BMO from dyadic BMO. PJM99:351-371. Garsia, A. M. (1970). Topics in almost everywhere convergence. Markham, Chicago.
(1973a). The Burges-Davis inequality via Fefferman's inequality. Ark. Mat. 11:229-237. (1973b). On a convex function inequality for martingales. Ann. Prob. 1: 171-174. (1973c). Martingale inequalities: Seminar notes on recent progress. Benjamin, Reading, MA. Getoor, R. K., and M. J. Sharpe (1972). Conformal martingales. Invent. Math. 16: 271-308. (1979). Excursions of Brownian motion and Bessel processes. ZfW 47:83-106. Gikhman, I. I., and A. V. Skorohod (1973). Stochastic differential equations. SpringerVerlag, New York. Gilbarg, D., and N. S. Trudinger (1977). Elliptic partial differential equations of second order. Springer-Verlag, New York. Girsanov, I. V. (1960). On transforming a certain class of stochastic processes by absolutely continuous change of measure. TPA 5:285-301. (1961). On Ito's stochastic integral equation. Soviet Math. 2:506-509. (1962). An example of nonuniqueness of the solution of K. Ito's stochastic integral equation. TPA 7:325-331.
Gundy, R. F. (1968). A decomposition for L' bounded martingales. Ann. Math. Slat. 39:134-138.
(1969). On the class LlogL, martingales, and singular integrals. Studia Math. 33:109-118. (1980a). Inequalites pour martingales a un et deux indices: l'espace H'. L`cole d'ete
Reference.
319
de Probabilitcfs de Saint-Flour, VIII. Springer LNM 774. (1980b). Local convergence of a class of martingales in multidimcnsioniil time. Ann. Prob. 8:607-614. Gundy, R. F., and E. M. Stein (1979). H° theory for the polydisc. Proc. Nat. Acad. Sci., USA 76:1026-1029. Gundy, R. F., and N. Varopoulos (1976). A martingale that occurs in harmonic analysis.
Ark. Mat. 14:179-187. (1979). Les transformations de Riesz et les integrales stochastiques. C. R. Acad. Sci. Paris 289A: 13-16. Gustafson, K. E. (1980). Introduction to partial differential equations. John Wiley, New York. Hardy, G. H. (1915). The mean value of the modulus of an analytic function. Proc. LMS 14:269-277. (1928). Remarks on three recent notes in the Journal. J. LMS 3: 166 169. Hardy, G. H., and J. E. Littlewood (1926). Some new properties of Fourier constants. Math. Ann. 97:159-209.
(1930). A maximal theorem with function-theoretic applicutions. Acta Math. 54:81-116. (1931). Some properties of conjugate functions. J. Reine Anyewt-. Math. 167:405423.
Helson, H., and G. Szego (1960). A problem in prediction theory. Ann. Mat. Pura Appl. 51:107-138. Herz, C. (1974a). Bounded mean oscillation and regulated martingales. TAMS 193:199216.
(1974b). Hp spaces of martingales 0
Hoffman-Ostenhof, M. and T., and B. Simon (1980a). Brownian motion and a consequence of Harnack's inequality, nodes of quantum wave functions. PA MS 80: 301305.
(1980b). On the nodal structure of atomic eigenfunctions. J. Phys. A. 13:1131-1133. Hunt, G. A. (1956). Some theorems concerning Brownian motion. TAMS 81:294-391. Hunt, R. A., B. Muckenhoupt, and R. L. Wheeden (1973). Weighted norm inequalities for conjugate functions and Hilbert transforms. TAMS 176:227-251. Ikeda, N., and S. Watanabe (1981). Stochastic differential equations and diffusion processes. North Holland, Amsterdam. Ito, K. (1944). Stochastic integrals. Proc. Imp. Acad. Tokyo 20:519-524. (1946). On a stochastic integral equation. Proc. Japan Acad. 22:2, 32-35. (1950a). Brownian motions in a Lie group. Proc. Japan Acad. 26:8, 4-10. (1950b). Stochastic differential equations in a differentiable manifold. (I) Nagoya Math. J. 1:35-47; (1I) Mem. Coll. Sci. U. Kyoto, Series A, 28, 1, 81-85. (1951a). On a formula concerning stochastic differentials. Nagoya Math. J. 3:5565.
(1951b). On stochastic differential equations. Memoirs AMS, no. 4. (1960). Wiener integral and Feynman integral. In Proceedings of the Fourth Berkeley Symposium, vol. II, pp. 227-238. Ito, K., and H. McKean, Jr. (1964). Diffusion processes and their sample paths. SpringerVerlag, New York. Jacod, J. (1979). Calcul stochastique et problemes de martingales. Springer LNM 714.
320
References
Jacod, J., and J. Memin (1981). Weak and strong solutions of stochastic differential equations: Existence and stability. In Stochastic integrals, ed. Williams (1981), pp. 169-212.
Janson, S. (1977). Characterization of H' by singular integral transformations on martingales and R". Math. Scand. 41:140-152. (1979). Singular integrals on local fields and generalizations to martingales. In Harmonic analysis in Euclidean space, ed. Weiss and Wainger (1979).
(1981). BMO and commutators of martingale transforms. Ann. Inst. Fourier (Grenoble) 31:265-270. John, F. (1961). Rotation and strain. CPAM 14:391-413. (1982). Partial differential equations. 4th ed. Springer-Verlag, New York. John, F., and L. Nirenberg (1961). On functions of bounded mean oscillation. CPAM 14:415-426. Jones, P. (1980). Carleson measures and the Fefferman-Stein decomposition of BMO (R). Ann. Math. 111: 197-208. Kac, M. (1946). On the average of a certain Wiener functional and a related limit theorem in the calculus of probability. TAMS 59:401-414. (1949). On distributions of certain Wiener functionals. TAMS 65:1-13. (1951). On some connections between probability theory and differential and integral equations. In Proceedings of the Second Berkeley Symposium, pp. 189-215. (1953). An application of probability theory to the study of Laplace's equation. Ann. de Societe Math. Polonaise 25:122-130. (1959). Probability and related topics in physical science. Interscience, New York. (1966a). Can one hear the shape of a drum? Amer. Math. Monthly 73:1-23. (1966b). Wiener and integration in function spaces. BAMS 72:52-68. (1970). On some probabilistic aspects of classical analysis. Amer. Math. Monthly 77:586-597. (1972). On applying mathematics: Reflections and examples. Quart. J. Appl. Math. 30:17-29. (1974). A stochastic model related to the telegrapher's equation. Rocky Mountain
J.4:497-509. Kac, M., and J. M. Luttinger (1974). Probabilistic methods in scattering theory. Rocky Mountain J. 4:511-537. (1975). Scattering length and capacity. Ann. Inst. Fourier (Grenoble) 25:317-321. Kahane, J. P. (1976). Brownian motion and classical analysis. Bull. LMS 7:145-155. Kakutani, S. (1944a). On Brownian motion in n-space. Proc. Imp. Acad. Tokyo 20:648652.
(1944b). Two-dimensional Brownian motion and harmonic functions. Proc. Imp. Acad. Tokyo 20:706-714.
(1945). Markov processes and the Dirichlet problem. Proc. Imp. Acad. Tokyo 21:227-233. Kato, T. (1973). Schrodinger operators with singular potentials. Israel J. Math. 13:135148.
Katznelson, Y. (1968). An introduction to harmonic analysis. John Wiley, New York.
Kellog, O. D. (1929). Foundations of potential theory. Springer-Verlag. Reprinted by Dover Books, New York, 1954.
Khas'minskii, R. Z. (1959). On positive solutions of the equation Au + Vu = 0. TPA 4:309-318. Khintchine, A. (1933). Asymptotische Geseteze der Wahrscheinlichkeitsrechnung. Ergebn. Math., Berlin.
Refer nc.N
321
Knight, F. B. (1971). A reduction of continuous square-integrable martingaleN to Brownian motion. In Martingales, ed. H. Dinges, pp. 19-31. Springer LNM 190. (1981). Essentials of Brownian motion and diffusion. AMS, Providence, RI. Kolmogorov, A. (1925). Sur les fonctions harmoniques conjuguees et les series de Fouricr.
Fund. Math. 7:24-29. Koosis, P. (1980). Lectures on HP spaces. LMS Lecture Notes. Cambridge Univ. Press, Cambridge. Kunita, H., and S. Watanabe (1967). On square integrable martingales. Nagoya Math. J. 30:209-245. Ladyzenskaya, O. A., V. A. Solonnikov, and N. N. Ural'ceva (1968). Linear and quasilinear equations of parabolic type. AMS Translations of Mathematical Monographs 23. Latter, R. (1977). A decomposition of HP(R") in terms of atoms. Studia Math. 62:92-101. (1979). The atomic decomposition of Hardy spaces. In Harmonic analysis in Euclidean space, ed. Weiss and Wainger (1979).
Lebesgue, H. (1924). Conditions de regularite, conditions d'irregularite, conditions d'impossibilite dans le probleme de Dirichlet. C. R. Acad. Sci. Paris, pp. 349-354. Lenglart, E. (1977). Sur la convergence presque sure des martingales locales. C. R. Acad. Sci. Paris 284A: 1085-1088. Lepingle, D. (1978). Sur le comportement asymptotique des martingales locales. In Sem. XII, pp. 148-161. Springer LNM 649. Levy, P. (1939). Sur certain processes stochastiques homogenes. Compositio Math. 7:283339.
(1940). Le mouvement Brownien plan. Amer. J. Math. 62:487-550. (1948). Processus stochastiques et mouvement Brownien. Gauthier-Villars, Paris. (1951). Wiener's random function. In Proceedings of the Second Berkeley Symposium, pp. 171-186. Lieb, E. (1976). Bounds on the eigenvalues of the Laplace and Schrodinger operators.
BAMS 82:751-753. Lieb, E., and B. Simon (1982). Pointwise bounds on eigenfunctions and wave packets in n-body quantum systems VI : Asymptotics in the two-cluster region. Adv. Appl. Math. 1:324-343. Littlewood, J. E. (1926). On a theorem of Kolmogorov. J. LMS 1:229-231. (1929). On a theorem of Zygmund. J. LMS 4: 305-307. Loomis, L. H. (1946). A note on the Hilbert transform. BAMS 52:1082-1086. Lyons, T. J., and H. P. McKean (in press). Winding of plane Brownian motion. Preprint. Oxford Univ. Press, New York. McKean, H. P. (1969). Stochastic integrals. Academic Press, New York. (1977). -A plus a bad potential. J. Math. Phys. 18:1277-1279. Marcinkiewicz, J., and A. Zygmund (1938). A theorem of Lusin. Duke Math. J. 4:473-485. Messulam, P., and M. Yor (1982). On D. Williams' punching method and some applications. J. LMS. Meyer, P. A. (1966). Probability and potentials. Blaisdel, Waltham, MA. (1976). Un cours sur les integrabes stochastiques. In Sem. X, pp. 245-400. Springer LNM 511. (1977). Le dual de H'(R') demonstrations probabilistes. In Sem. XI, pp. 132-195. Springer LNM 581. Millar, P. W. (1968). Martingale integrals. TAMS 133:145-166. Moser, J. (1961). On Harnack's theorem of elliptic differential equations. CPAM 14:377 591.
322
Reference.
Nevanlinna, F. and R. (1922). Uber die Eigenschaften analytischer Funktionen in der Umgebung einer singularen Stelle oder Linie. Acta Soc. Sci. Fenn. 50, no. 5. Neveu, J. (1975). Discrete parameter martingales. North Holland, Amsterdam. Olenik, O. A. (1966). Alcuni risultati sulle equazioni lineari e quasilineari ellitico-paraboliche a derivate parziali del second ordine. Rend. Classe Sci. Nat. A cad. Naz. Lincei, Series 8, 40:775-784.
Paley, R., N. Wiener, and A. Zygmund (1933). Note on random functions. Math. Z. 37:647-668. Paley, R. E. A. C., and A. Zygmund (1932). A note on analytic functions in the unit circle. P. Camb. Phil. Soc. 28:266-272. Petersen, K. E. (1977). Brownian motion, Hardy spaces, and bounded mean oscillation. LMS Lecture Notes. Cambridge Univ. Press, Cambridge. Pichorides, S. K. (1972). On the best values of the constants in the theorems of M. Riesz, Zygmund, and Kolmogorov. Studia Math. 46:165-179. Pitman, J. (1979). A note on LZ maximal inequalities. In Sem. XV, pp. 251-258. Springer LNM 850.
Plessner, A. (1928). Uber das Verhalten analytischen Funktionen auf dem Rande des Definitions-bereiches. J. Reine Angew. Math. 158:219-227. Port, S., and C. Stone (1978). Brownian motion and classical potential theory. Academic Press, New York. Privalov, I. I. (1916). Sur Ies fonctions conjuguees. Bull. Soc. Math. France (1916), pp. 100103.
(1924). Generalization of a theorem of Fatou. Mat. Sbornik 31:232-235. Protter, M. H., and H. F. Weinberger (1966). On the spectrum of general second-order equations. BAMS 72:251-255. (1967). Maximum principles in differential equations. Prentice-Hall, New York. Rao, K. M. (1969). On decomposition theorems of Meyer. Math. Scand. 24:66-78. Ray, D. (1954). On the spectra of second-order differential operators. TAMS 77:299-321. Riesz, F. (1923). Uber die Randwerte einer analytischen Funktion. Math. Z. 18:87-95. Riesz, M. (1927). Sur les fonctions conjuguees. Math. Z. 27:218-244. Rosenblatt, M. (1951). On a class of Markov processes. TAMS 71:120-135 Royden, H. (1968). Real analysis. 2d ed. Macmillan, New York. Sharpe, M. J. (1980). Local times and singularities of continuous local martingales. In Sem. XIV, pp. 76-101. Springer LNM 784. Simon, B. (1974-1975). Pointwise bounds on eigenfunctions and wave packets in n-body
quantum systems. (I) PAMS 42:395-401; (II) PAMS 45:454-456; (III) TAMS 208:317-329. (1979). Functional integration and quantum physics. Academic Press, New York. (1980). Brownian motion, L° properties of Schrodinger operators and the localization of binding. JFA 35:215-229.
(1981). Large time behavior of the L° norm of Schrodinger semigroups. JFA 40:66-83. (1982). Schrodinger semigroups. BAMS (new series) 7:447-526. (1983). Instantons, double wells, and large deviations. BAMS (new series) 8: 323326.
Sion, M. (1958). On general minmax theorems. PJM 8:171-176. Skorohod, A. V. (1956). Limit theorems for stochastic processes. TPA 1:261-290. (1961). Existence and uniqueness of solutions to stochastic differential equations. Sibirsk. Mat. Zur. 2:129-137.
Rsfv.ncss
323
(1965). Studies in the theory of random processes. Addison-Wesley, Reading, MA. Spanne, S. (1966). Sur l'interpolation entre les espaces Ann. Scusla Norm Sup. Pisa 20:625-648. Spencer, D. C. (1943). A function theoretic identity. Amer. J. Math. 65:147-160. Spitzer, F. (1958). Some theorems concerning two-dimensional Brownian motion. TAMS 87:187-197. Spivak, M. (1979). A comprehensive introduction to differential geometry. Publish or Perish, Inc., Berkeley, CA.
Stein, E. M. (1961). On the theory of harmonic functions of several variables II. Acta Math. 106:137-174. (1962). Conjugate harmonic functions in several variables. In Proceedings of the International Congress of Mathematics, Stockholm, pp. 414-420. (1967). Singular integrals, harmonic functions and differentiability properties of functions of several variables. Proc. Symp. Pure Math. 10: 316 335, (1969). A note on LlogL. Studia Math. 31:305 310. (1970a). Singular integrals and differentiability properties of functions. Princeton Univ. Press, Princeton. (1970b). Topics in harmonic analysis related to the Littlewood-Paler theory. Annals of Mathematical Studies, no. 63. Princeton Univ. Press, Princeton. Stein, E. M., and G. Weiss (1959). An extension of a theorem of Marcinkiewicz and its applications. J. Math. Mech. 8:263-284. (1960). On the theory of harmonic functions of several variables 1. Acta Math. 103:25-62. (1968). Generalization of the Cauchy-Riemann equations and representations of the rotation group. Amer. J. Math. 90:163-196. (1971). Introduction to Fourier analysis on Euclidean spaces. Princeton Univ. Press, Princeton. Stein, P. (1933). On a theorem of M. Riesz. J. LMS 8:242-247. Stroock, D. W. (1973). Application of Fefferman-Stein type interpolation to probability theory and analysis. CPAM 26:477-496. (1981). The Malliavin calculus and its applications. In Stochastic integrals, ed. D. Williams (1981), pp. 394-432.
(1983). The Malliavin calculus and its application to second-order parabolic equations. Math. Systems Theory 14:25-65.
Stroock, D. W., and S. R. S. Varadhan (1969). Diffusion processes with continuous coefficients. CPAM 22: (I) 345-400, (II) 479-530.
(1970). On the support of diffusion processes with applications to the strong maximum principle. In Proceedings of the Sixth Berkeley Symposium, vol. III, pp. 333-360. (1972a). On degenerate elliptic parabolic operators of second order and their associated diffusions. CPAM 25:651-713. (1972b). Diffusion processes. In Proceedings of the Sixth Berkeley Symposium, vol. III, pp. 361-368. (1974). A probabilistic approach to H"(Re). TAMS 192:245-260. (1979). Multidimensional diffusion processes. Springer-Verlag, New York. Taibleson, M. H. (1975). Fourier analysis on local fields. Mathematical Notes. Princeton Univ. Press, Princeton. (1979). An introduction to Hardy spaces on local fields. In Harmonic analysis in Euclidean space, ed. Weiss and Wainger (1979).
324
References
Taibleson, M. H. and G. Weiss (1979). The molecular characterization of Hardy spaces. In Harmonic analysis in Euclidean space, ed. Weiss and Wainger (1979). Tanaka, H. (1964). Existence of diffusions with continuous coefficients. Mem. Fac. Sci. Kyushu, Series A 18:89-103. Tsirel'son, B. S. (1976). An example of a stochastic differential equation having no strong solution. TPA 20:416-418. Uchiyama, A. (1982). A constructive proof of the Fefferman-Stein decomposition of BMO(R"). Acta Math. 148:215-241. (1983). A constructive proof of the Fefferman-Stein decomposition of BMO on simple martingales. In Conference on harmonic analysis in honor of Antoni Zygmund,
vol. II, ed. Beckner, Calderon, Fefferman, and Jones, pp. 495-505. Wadsworth, Belmont, CA. Varopoulos, N. (1977a). BMO functions and the a equation. PJM 71:221-273. (1977b). A remark on BMO and bounded harmonic functions. PJM 73:257-259. (1979). A probabilistic proof of the Garnett-Jones theorem on BMO. PJM 90: 201221.
(1980a). Aspects of probabilistic Littlewood-Paley theory. JFA 38:25-60. (1980b). The Helson-Szego theorem and Ap functions for Brownian motion and several variables. JFA 39:85-121.
Walsh, J. B. (1977). A property of conformal martingales. In Sem. XI, pp. 490-501. Springer LNM 581. Walsh, J. L. (1929). The approximation of harmonic functions by harmonic polynomials and harmonic rational functions. BAMS 35:499-544. Weiss, G. (1970). Complex method in harmonic analysis. Amer. Math. Monthly 77:465474.
Weiss, G., and S. Wainger, eds. (1979). Harmonic analysis in Euclidean space. Proceedings of Symposia in Pure Mathematics, vol. 35. AMS, Providence, RI. Widder, D. V. (1975). The heat equation. Academic Press, New York. Wiener, N. (1923). Differential space. J. Math. Phys. 2:131-174. (1924a). Certain notions in potential theory. J. Math. Phys. 3:24-51. (1924b). The Dirichlet problem. J. Math. Phys. 3:127-146. (1924c). Un probleme de probabilitese enombrables. Bull. Soc. Math. de France 52:569-578. (1925). Note on a paper of O. Perron. J. Math. Phys. 4:21-32. (1930a). Generalized harmonic analysis. Acta Math. 55:117-258. (1930b). The homogeneous chaos. Amer. J. Math. 80:897-936.
Williams, D., ed. (1981). Stochastic integrals. Proceedings LMS Durham Symposium. Springer LNM 851. (ND). A simple geometric proof of Spitzer's winding number formula for 2dimensional Brownian motion. Unpublished paper. University College of Swansea, Wales.
Yamada, T., and S. Watanabe (1971). On the uniqueness of solutions to stochastic differential equations. J. Math. Kyoto 11: (1) 155-167, (II) 553-563. Zygmund, A. (1929). Sur les fonctions conjuguees. Fund. Math. 13:284-303; correction, pp. 18, 312. (1949). On the boundary values of several complex variables. Fund. Math. 36: 207235.
(1959). Trigonometric series. 2d ed. Cambridge Univ. Press, London.
Index of Notation
Rd
1
D
36
,Pd
3
ky
36
.
C
5
he
38
w
5
GD(x, y)
39
sofa
106
P.
5
A
44
sat
106
11
49
2*
107
<X>, <X, Y>
52
.
V*
107
54
a*
107
10
n
55
PB a.s.
107
12
b1I°
55
Yf
119
55
S,(0)
126
°
Pµ
P, (X, y) 05
+
5 8
8
K.
106
'
106
A,u
106
.FS'
17
d
17
n1
56
N,u
127
T.
17
112(X)
57
hp
144
.
19
II Hllx
57
dp(u)
144
TA
19
.,r f2
57
144
Os
20
113 (X)
58
.FS
20
D(x, 6)
23
fbH J
59
Hp it X*
80
. ffp
147
i.o.
28
PB
96
148
28
144
147
5
102
IIXIIp X,*
28,71
VB
105
."p
150
G(x, y)
31
-T,
105
.Y'-p
150
H
32
2'
105
IlXllx.
151
35
N,u
106
IIXIII
151
D; A
GH(x,y)
149
325
326
Index of Noteda
192
Q
272
BMO
193
X,"
282
163
II(pII*
193
Xs X
288
it
170
<(p>>
198
a.s.
294
(AIP)*
184
203
'A
295
203
i.o.
295
<M >l Ho (A*X),
152 155
R.,1f10h
91#0
186
M N
IIXII*
186
ao(X)
211
Q(X)
296
"OZ
188
Kd
228
E(X I.F)
300
<<X%
188
Kd'O0
229
192
L
271
Al
Subject Index adapted process, 44 almost surely, 294 associative law, 62 asymptotic a-field, 17 atomic decomposition
forp=1,184 forp<1,216 Blumenthal's 0-1 law, 14 Borel Cantelli lemmas, 295, 299
bounded mean oscillation, 186, 193, 199, 205 Brownian motion, 1 Brownian paths nondifferentiability, 5 quadratic variation, 6 Holder continuity, 7, 15 BurklLNder-Gundy mequalihes, 155 Burkholder's weak type inequality, 164
Cameron-Martin transformation, 234 Cauchy process, 2, 33 Chebyshev's inequality, 296 Ciesielski and Taylor's "paradox," 260 conditional expectation, 300
conditioned Brownian motions in H, 94, 97 in R° - {0}, 100 in D, 126 conjugate harmonic functions, 162, 170 covariance of two local martingales, 54
of two stochastic integrals, 61 of two semimartingales,
Doob's inequality, 147, 306 sharp constant in, 151 Doob-Meyer decomposition, 52 distributive law, 68 duality theorem
for.*', 184 for H', 199 Dynkin's n A theorem, 9
explosions
Janson's theorem, 167 Jensen's formula, 132 Jensett's inequality (for conditional expectations), 301 John-Nucnherg inequality, :1114, 209
from too much drift, 240 from loo much variance, 277
Fellcr's test, 242 exponential nuirtingale, 27, 70
Fefferman's inequality, 187, 190
Fefferman-Stein decomposition. 202 Feynman-Kac formula, 229 filtration, 50 fine topology, 108
Garnett-Jones theorem, 211, 214
Girsanov's formula, 82 good A-inequalities, 153 Green's function, 39 Gundy's L log L inequality, 148
Hardy space H°, 144 Harnack's inequality (for solutions of Schrt dinger's equation). 261 heat equation, 220 Helson-SzegO theorem, 215 h-transform, 92, 94
68
Dirichlet problem, 43, 246 Donsker and Varadhan, 266
for a local martingNlp, ri4 for several semimartingales, AN
Kelvin's Iunnsformations, No, 414
Khasmin'xMu's
231
Kolmoguruv's extension IIICI m, I continuity t,tw,1uru, 6 three series Ilies trni, 11411
(other) weak IvIs' inequality, I I i Kunita-Watanahe inequality, !9
law of the iterated logarithm for Brownian motion, I 1 for continuous local martingales, 77 Lebesgue's thorn, 248 Levy's theorem, 75 Levy's arc sin law, 261 local martingale, 50
Markov property of Brownian motion, 7 of diffusions, 288 martingale, 302 martingale convergence theorem, 305 martingale transforms, 162 maximum principle, 72, 115 monotone class theorem, 10
independence, 296 integration by parts, 68
U's formula
natural scale, 74, 93, 241 327
32
Subject Iodex
Nevanlinna class, 128 nontangential convergence,
predictable sequence, 48, 303
105
optional projection, 191 optional a-field, 44 optional stopping theorem,
quadratic variation of Brownian motion, 7 of continuous local martingales, 65
310
orthogonality of martingale increments, 303
Pac-Man, 249 Picard's theorems, 139, 143 Poincarb's cone condition, 249
Poisson's equation, 43, 251 Poisson integral representation of harmonic functions nonnegative functions, 98, 99 bounded functions, 119 functions in H°, 158 potential kernels, 31 predictable a-field, 49
Rayleigh-Ritz formula, 270 regular point, 247 Riesz's inequality, 170 sharp constant in, 172 Riesz transforms, 169
Schrodinger equation, 255, 263, 264 Schrodinger semigroups,
stable subspace, 87 stochastic differential equations existence of solutions, 274, 278 uniqueness, 278, 286 examples, 283 stopping time, 18, 306 stretching function, 51 strong Markov property for Brownian motion, 21 for conditioned Brownian motion, 102 submartingale, 302 subordination (of analytic functions), 176 supermartingale, 302
232
semimartingale, 58 shift, 10 shift invariant events, 102 simple random walk, 48, 91, 303 Spitzer's theorem (on the winding of Brownian motion), 134
uniform integrability, 308 usual domination argument, 62
variance process, 52 Varopoulos' staircase, 213 volume potential, 225