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!
'. Conversely, if
To
L
'(R 2 ) then
Mo
>HI ~ CIEI.
Then
is bounded on
is bounded on Lq(R 2 ), for p'< q < p [16].
To prove this assume first that
To
is bounded on LP(R 2). Then
the first step is to notice that this implies that
This is proven by observing that if we dilate the region Ro by a huge factor p to get
R~
and translate
R~
properly (by rk) then R:•rk
looks like a half plane with boundary line making an angle of ok with the positive x axis. Then if rk(t) are the Rademacher functions, and Tp,rkfA
0
and
=X
p,rk
Ro
f
then
87
MULTIPARAMETER FOURIER ANALYSIS
Taking the limit as p
oo,
4
\1 (~1Tkfk12)
we have
1~
IILP
~ell (~ifk12)
1~
IILP,
where Tk is the Hilbert transform in the direction Ok. The next step is to use the above inequality to prove a covering lemma for rectangles in
mo.
Let IRkl be a sequence of such rectangles.
""' 1, R ~ "" Select R given that R 2 ,· .. , Rk-l have been chlSen provided IR n [j~k
Rj ll < ~ JR J.
Then if R is unselected
Mo<Xu'R'/ ;: - ~
on R
so that
Let Ek
=
Rk-
u
~J.
j 1/100. Applying Tk =Hilbert transform in the direction perpendicular to ek to Tkfk we see that Tk(Tt/k) > 1/100 on all of R~. Repeating twice more we get
I
I x~) I(~
IILP ~ c I(~ lx~k 12 )
1/2
This shows that
Me
1/2
IILP ~ Cllxu~k\ILP ·
is of weak type (p/2)'.
Conversely, assume that
Me
is bounded on L(PI 2 >'. Define
Sk=l(o:((1'( 2 )\2k~( 1 01 whose values lie in Rn+l: F(x,t) = (u 0 (x,t),·· ·,un(x,t)) where the ui{x,t) satisfy the "Generalized Cauchy-Riemann equations," n
au.
I~ (x,t) i=O
=0
(t
=xo)
1
and
au.1
ax. J
auj
- "'L"" for all i,j. = ax1
These Stein-Weiss analytic functions are then said to be HP(R~+ 1 ) if and only if sup t>O
(J
1 IF(x,t)IPdx) /p = JIFII
[17].
HP(Rn)
Rn
Again, these functions have an interpretation in terms of singular integrals, since if a Stein-Weiss analytic function F(x,t) is sufficiently "nice" on R~+l, then the boundary values ui(x) satisfy ui{x) = Ri [u 0 ]{x) where Ri is the ith Riesz transform given by Ri(f Xx) = f
cnxi * --. lxln+l
1
In particular we may consider an H (Rn+
+1
) function {by
identifying functions in R~+l with their boundary values) as a function f with real values in L 1(R 0 ) each of whose Riesz transforms Rif also belong to L 1 (Rn). An interesting feature of HP spaces is that they are intimately connected to differentiation theory as well as singular integrals. To discuss this, let us make some well-known observations. For a harmonic function u(x,t) which is continuous on R~+l and bounded there, u is given as an average of its boundary values acc01ding to the Poisson integra 1:
91
MULTIPARAMETER FOURIER ANALYSIS
Let f(x) = l(y,t)\ \x-y\ < tl. Then since convolving with Pt at a point x can be dominated by an appropriate linear combination of averages of f over balls centered at x of different radii, it follows that if u*(x)
=
sup
\u(y ,t)\ , then u*(x)
'S cMf(x)
.
(y,t)lf{x)
Unfortunately, if u(x,t) is harmonic, for p :S 1, u
=
P[fJ, and
fan \u(x,t)\Pdx ~ C then the domination u* ~ CMf is not useful, since M is not bounded on LP , and it is not true in general that u*(x) for a,.e.
F
say
X l i
< oo
Rn. On the other hand, suppose F is Stein-Weiss analytic,
H 1 (R~+l).
Then a beautiful computation shows that if 1 >a > 0 is close enough to 1
{a~ n~l) then 6(\F\a)?: 0 so that \Fia is subharmonic.
H
s(x,t) is subharmonic and has boundary values h(x) then s is dominated by the averages of h, i.e., s(x,t) ~ P[h](x,t) . Applying this to G = \Fia (which has we see that G*
'S M(h)
Ll/a so that M(h)
l
for some h
t
JG 1 /a(x,t)dx
~ C for all t > 0)
Ll/a. Now M is bounded on
Ll/a and so G* (Llla. It follows that F*
l
L1 •
Just as for a random f ( L 1(Rn) we do not necessarily have Rif ( L 1 (R 0 ) (singular integrals do not preserve L 1 ) it is also not true that for an arbitrary L 1 function f that for u = P[f]. u* then u* (
L 1(Rn).
l
L 1 . But if f
Thus the nontangential maximal function F*(x) ~
sup
\F(y ,t)\
l
L 1(Rn)
(y,t)lf(x)
if and only if the analytic function f
l
H 1 (R~+l).
l
H1 (R!+ 1 )
92
ROBERT FEFFERMAN
We know so far that we can characterize HP functions in terms of singular integrals and maximal functions. There is another characterization which is of great importance. To discuss it, let us return to HP functions in R ~ as complex analytic functions, F
=
u + iv. It is an
interesting question as to whether the maximal function characterization of HP can be reformulated entirely in terms of u. That is, is it true that F* c LP if and only if u* c LP ? In fact, this is true, and the best way to see this is by introducing a special singular integral, the LusinLittlewood-Paley-Stein area integral,
s 2(u)(x) =
JJ
l'i7ui 2 (y.t)dtdy
l(x)
which we already considered in the first lecture. As we shall see later, for a harmonic function u(x,t), IIS{u)ll
LP
~ llu*!l
LP
for all p
> 0 [18].
The importance of S here is that the area integral is invariant under the Hilbert transform, i.e., S(u)
=S(v),
since l'i7vl = l'i7ul.
When we combine the last two results, we immediately see that
It is interesting to note that the first proof of !IS{u)ll
""'!iu*ll P, LP L 1 ~ p > 0 was obtained by Burkholder, Gundy, and Silverstein [19] by using probabilistic arguments involving Brownian motion. Nowadays direct real variable proofs of this exist as we shall see later on. To summarize, we can view functions f in HP spaces by looking at their harmonic extensions u to R~+l and requiring that u* or S{u) belong to LP(Rn).
93
MULTIPARAMETER FOURIER ANALYSIS
It tums out that there is another important idea which is very useful
concerning HP spaces and their real variable theory. So far, we have spoken of HP functions only in connection with certain differential equations. Thus, if we wanted to know whether or not f take u
= P[f] which of course satisfies
~u =
l
HP we could
0.
This is not necessary. If f is a function and f/J < c;(Rn) with
fR
¢n =1, then we may form f* (x) =
sup
If *f/Jt(y)l, ¢t(x) =
(t,y)E r(x)
en rp(x/t) and if 1/J < c;(Rn) and J.P = 0 we may form s3.(f Xx) ==
is suitably non-trivial (say radial, non-zero)
rr
J,
'I'
if * .Pt(y)l 2 dy dt tn+l
[18] .
rex)
Then C. Fefferman and E. M. Stein have shown that 11£11 P n "" llf*ll P n "" IIS,,.(f)il P n for 0 H (R ) L (R ) 'I' L (R )
< p < oo
•
Thus, it is possible to think of HP spaces without any reference to particular approximate identities like Pt{x) which relate to differential equations. In addition to understanding the various characterizations of HP spaces, another important aspect is that of duality of H 1 with BMO, which we shall now discuss. A function rp(x), locally integrable on Rn is said to belong to the class BMO of functions of bounded mean oscillation provided
~~I
J
lr/J{x)- r/Jo I dx
'S
M
for all cubes Q in Rn ,
Q
f. f/J. IQI o
where r/Jo = _!_
The BMO functions are really functions defined
modulo constants and
II
llaMo is defined to be sup
l~l J0
lr/J-r/Jol·
ROBERT FEFFERMAN
94
According to a celebrated theorem of C. Fefferman and Stein, BMO is the dual of H 1 (18]. This result's original proof involves knowing that singular integrals map L"" to BMO and also a characterization of BMO functions in terms of their Poisson integrals which we now describe. Suppose p. :2: 0 is a measure in R~+l and Q S.: Rn is a cube. Let S(Q)
=
{y,t)ly £Q, 0 < t <side length (Q)I. Then we say that p. is a
Carleson measure-on R~+l iff p.(S(Q)) ~ CIQI. The basic property that characterizes Carleson measures is
for all functions u on R~+l. In connection with this type d measure there is the characterization of functions in BMO(Rn) in terms of their Poisson integrals. A function cp(x) on Rn with Poisson integral u(x,t) is in BMO if and only if the associated measure
is a Carleson measure. C. Fefferman proved this and used it to prove that every function in BMO acts continuously on H 1 :
These are the basic facts of HP spaces that will concern us here and which we shall later generalize to product spaces. Let us now prove that for a harmonic function u(x ,t)
95
MULTIPARAMETER FOURIER ANALYSIS
We begin with the estimate llu*llp
:S Cpi!S(u)llp, and to do this we
shall show that
liu*>Cal[
~ C~2 ~
J
S2(u)(x)dx +
liS(u)>al~.
J
S(u}all we have
~ J~ ..-
Hu. n
1 Au .(a)da
p -
f
0
j .•- f.~ j t 1
0
.S(u)(JJ)d {3 + •• ,.
j
J 00
aP- 3dad{3 +
aP-l>..S(u)(a)da .
0
00
Assuming p < 2 as we clearly may, this is
~
I
00
f3P-l>..S(uif3)d{3
-v
1\S(u)l\~.
0
To prove the estimate on llu* >Call, we set the notation that
E: = ~M(xE) > l.
f
H,
and then claim
S 2 (u)(x)dx:::: c JJI'Vu(y.t)\ 2 t dtdy where
ln fact, if (y ,t) dR then \B(y,t) n!S(u)>all
al
R
S(u):Sa
!R
t
\B(y,t)l .
.
96
ROBERT FEFFERMAN
Then
I
I ( I iVul 2(y,t)t 1 -ndyd~
S2 (u)(x)dx "
xd S(u~a
S(u}Sa
(4.1) -=
Jf
I
dx
T(x)
\'Vu\ 2 (y,t)t 1 -nJixj(y,t)d'{.i),x/lS(u)>allldydt.
Rn+l
+
But fa (y,t)
f
R,
\lx\(y,t)£r{x),xi!S(u)>all\-= \B(y,t)nciS(u)>al\ ~ ~ \B(y,t)\ and (4.1) is
as claimed.
II. The next step is to write \'Vu \2 to 9t :
ffA(u 2)(y,t) t dydt = JJ' R
Now
t
=
A(u 2 ), and apply Green's theorem
J~an
t- u 2
aR
~do. an
~ c > 0 for some c so the above gives
Since, for purposes of all estimates we may assume that u is rather nice, we may assume u('Vu)t vanishes at t = 0, so
97
MULTIPARAMETER FOURIER ANALYSIS
J
u(Vu)t do=
dR
J
u(Vu) t do
dR
where aR is the part of dR above IS(u) > al. It is not hard to see that !Vult~ a on dR so that
Putting all of our estimates together, we see that
I
u 2du
~J
I
t(u~a
dR
S2 (u)(x)dx +a'IS(u)
III. Next we wish to define a function
f
by
f(
x)
=
>
.,,l
)
u(x, r(x)) where
(x, r(x)) t: dR defines the function r. We claim that in the region R
lui~ Plr] +Ca.
(4.2)
This is done by harmonic majorization. It is enough to show this on dR, and this in turn is just saying that for any point p t: c1R, jU(p)l is dominated by the average over a relative ball on dR of u +Ca. This follows from the estimate
IVult~ Ca
on dR. Anyway, from (4.2) we
have, for xi IS(u) > al, u*(x) ~ CP[f]*(x) + Ca, so that finally llu* > C'all
~ IM((x) >all~ ~ llfll~ a
< _g_ a2
I S(u)~
This completes the proof.
S 2 (u)(x)dx + c llS(u)
>all .
98
ROBERT FEFFERMAN
The proof that \\u*\\p ~ Cp\IS(u)\\p which we ;ust gave has been lifted from Charles Fefferman and E. M. Stein's Acta paper [18]. To prove the reverse inequality we want to go via a different route, and we shall follow Merryfield here [20]. We prove the following lemma. In the next lecture we show how this lemma proves \lu*l\p ~ Cp\IS(u)l\p. LEMMA.
Let f(x) and g(x)
f
L 2 (Rn),
and suppose
rp
f
and u ~ P[£]. Then
for some
rp
f
c;(Rn) with
f rp
== 0 (
rp
real-valuecl).
Proof.
-2
Jf Rn+l
+
u(x,t)
~ u(x,t)(g*cJ>t) 2(x,t)dtdx
c;(R") radial
MULTIPARAMETER FOURIER ANALYSIS
where
I
=
JJ
u(tV(g*cf>t)) t -l 12 · Vu(g*cf>t) t 1 12dt dx
Rn+l
+
and
We see that
but
99
100
ROBERT FEFFERMAN
But
so
Putting this together gives
5.
More on HP spaces At this point we wish to discuss the theory of multi-parameter HP
spaces and BMO. We saw, in the last lecture, that HP(Rn) could be defined either by maximal functions or by Littlewood-Paley-Stein theory. All of these spaces, HP and BMO were invariant under the usual dilations on R 0
,
x-+ 8x, and this is hardly a surprise, since they can be
defined by the maximal functions and singular integrals which are
MULTIPARAMETER FOURIER ANALYSIS
101
invariant under these dilations. Here we shall define HP and BMO spaces which are invariant under the dilations (in R 2 ) (x 1 ,x 2 ) -. (8 1x 1 ,8 2 x 2 ), 8 1 ,8 2 > 0. For convenience we shall work in R 2 but all of this could just as we 11 be carried out in Rn x Rm, n, m > 1 . We shall call our HP and BMO spaces "product HP and BMO" and denote them by HP(R~ x R~) and BMO(R! x R~) so that we reserve HP(R 2 ) for the one-parameter space ci functions on R 2 • Let (x 1 ,x 2)
f
R 2 and denote by r(x) the set
Let u(x,t) be a function in R~ x R~, x
f
R2 , t
f
R+ x R+' which is
biharmonic, i.e., harmonic in each half plane separately. Then the nontangential maximal function and area integral ci u are defined by u*(xl'x 2) =
iu(y 1 ,t 1 ,y 2 ,t 2)1
sup (y,t)tf'<x 1 ,x 2 )
and
S2 (u)(x) =
Jf
1";;\ \72u(y.t)l 2dyldy2dtldt2.
f'(x)
More generally, if ¢
f
C~(R 2 ) and if
then for a function f on R 2 f*(x) =
sup (y, t) f f'<x)
If* ¢t t (x)l 1' 2
102
ROBERT FEFFERMAN
then
Given f(x 1 , x 2 ), we define its hi-Poisson integral by u(x 1 ,t 1 ,x 2 ,t 2 ) = P[f](x,t) = f * Pt t (xl'x 2), where the hi-Poisson (or just Poisson for 1' 2
short) kernel is defined by Pt t (x 1 ,x 2 ) 1, 2
=
P(x~\
t1 1t; 1 P(x 1)
t2)
tl
and
where P is the !-dimensional Poisson kernel. Then, of course, P[f] is bi-harmonic in R! x R!. In analogy with the 1-parameter case, we define f c HP(R!xR!) if and only if u* c LP(R 2) where u = P[fl. It is not hard to see, just as in the single parameter case that for any ¢ c c;(R 2), \lf*ll
L
P "'
llu* II
LP
f¢
=
1,
for p > 0 ,
so that we may use any approximate identity which is sufficiently nice to define product HP spaces. In terms rf area integrals, we also have, for 1/J(x 1 ,x 2) suitably nontrivial, say ¢
even in x 1' x 2 and not = 0,
To complete the chain of equivalences, we would like to know that
IIS(u)\1
L
P "'
\lu*\1
L
P.
103
MUL TIPARAMETER FOURIER ANALYSIS
In fact, this is true, but is not obvious, and so we in tend to present the proof here. The proofs are by iteration, but they are not of the same totally straightforward nature as the iteration in the jessen-MarcinkiewiczZygmund theorem. Often, this is the case in the analysis of product domains, namely, the proof is by iteration, but this requires a different way o{
looking at the one-parameter case than one is used to.
Proof that of S{u) ( LP, then, u*
l
LP (Gundy-Stein) [21 ]. We can
assume that u -= P[f]. Then since the 2-parameter area integra 1 is invariant under taking Hilbert transforms in each variable separately, we see that we may write
where ~2
f++
is supported in
g1, ( 2 > 0
and
f+-
is supported in ( > 0,
< 0, etc., and we have S(f±±) ( LP. By reflection we may assume f
is analytic (i.e., P[f]
is hi-analytic) and then show that f*
l
LP. But
for u = [f] a hi-analytic function, we know that for a > 0, ju(x 1 ,t 1 ,x 2 ,t 2 ) Ia is subharmonic in each half plane (xi,ti) l separately; this implies
Ri
that
lfa 0. In order to simplify things a little, we shall take a modified definition of u* in what follows, namely
u*(x) =
sup (y,t)fr
10
lu(y,t)l
10(x)
where
This is an irrelevant change, since a trivial computation shows that ·llu*llp is, for a larger aperture, no more than a constant times llu*llp for a smaller aperture, the constant depending only on the apertures involved.
In (5.3), take ¢(x) Let us estimate
=
1 for all lxl < 1/3, and g(x) =X
u
*< x )a)all < 2 ~ 0 jR(y,t)ll and where R(y;t) is
the rectangle in R 2 with sides parallel to the axes and with side lengths 2t1'2t 2 centered at y. Notice that if !R(y;t)nlu*>all < 1 ~ 0 !R(y;t)l then g * t(Y)
=
=
Ptg(y) > c for some c > 0. It follows that
i + ii + iii + iv. Consider i: If "' Qt(gXy)
*I 0
then u*(x)
~a
for some x
But then ju(y ,t) I ~ a so i is less than or equal to
f
R(y; t).
MULTIPARAMETER FOURIER ANALYSIS
ii is less than or equal to
Again
(iii) is similar to (ii). Finally, (iv) is less than or equal to
So we have
I lu*;Sal
S 2(u)(x)dx
~c
a 2 !1u*>all +
J u*(x);Sa
u* 2 (x)dx
109
110
ROBERT FEFFERMAN
and we have seen before that this implies that !IS(u)il
L
P
:S. Cpllu*ll
LP
, 2 > p > 0.
The next topic that we shall consider is that of duality of H 1 and BMO in the product setting. In the classical case there were four results which expressed this duality. 1) The characterization of Carleson measures 11 for which the Poisson transform f ... P[f] is bounded from LP(dx) to LP(dl1) , p > 1 . 2) The characterization of functions in BMO(R 1) by a condition on their Poisson integrals in terms of Carleson measures. 3) The characterization of functions in the dual of H 1 by the BMO condition. 4) The atomic decomposition of H 1 . Let us try to guess what the analogous theory should look like in product spaces. For simplicity we consider the dual of H 1 (R~xR~). Then what should an element of BMO(R!xR~) look like? We might look at tensor products of functions in BMO(R 1 ) to get a feel for the answer. So, for example if ¢ 1 and ¢ 2 are in BMO(R 1 ) then ¢ 1 (x 1 )¢ 2 (x 2) might be our model. Of course, this function ¢(xl'x 2) satisfies
(5.4)
~~I
I
l¢(x 1 ,x 2 )- c 1 (x 1 )-c 2 (x 2 )1 2 dx 2dx 2 :S. C
R
for the appropriate choice of functions c 1 and c 2 of the x 1 ,x 2 variable. A Carleson measure in R~ x R~ would be a non-negative measure 11 for which (5.5)
where P is the hi-Poisson integral. The obvious guess is that 11
111
MULTIPARAMETER FOURIER ANALYSIS
satisfies (5.5) if and only if IL(S(R)) ~ CIRI for all rectangles R
C
R2
with sides parallel to the axes, where the Carleson region S(R) is defined by S(I xj) == S(I) xSQ) for R =I xj. In terms of these Carleson measures, it is not hard to show that ¢ satisfies (5 .4) if and only if its hi-Poisson integral u satisfies
And finally, all of this in some sense is equivalent to asserting that every f
f
H 1 (R~ xR~) can be written as I A.kak where I IA.kl ~ Cilfll
H
1
and
ak(x 1 ,x 2 ) are "atoms," i.e., ak is supported in a rectangle Rk == Ikxjk such that
J
ak(x 1 ,x 2 )dx 1
0
for all x 2
ak(x 1 ,x 2 )dx 2 = 0
for all x 1
=
Ik
I Jk
and
In 1974 [22], L. Carleson showed that 11(S(R)) ~ CIRI was not sufficient to guarantee the inequality
J
ifiPdx.
R2
From here it is not difficult to produce examples of functions ¢(xl'x 2 ) which satisfy
112
ROBERT FEFFERMAN
R
where C 1 ,C 2 dependon R, yet ¢/LP(R 2 ) forany p>2. Therefore, this condition is not strong enough to force ¢ to belong to the dual of H1 . In other words the simple picture of the structure of H 1 (R~xR~) and BMO(R~ xR~) suggested above as the obvious guess is completely wrong.
Rather one considers the role of rectangles to be played instead by arbitrary open sets. Although this may seem a bit frightening at first glance, it turns out, and this is of course the final test of the theory, that nearly all the classica 1 theory of HP and BMO can easily be carried out using the approach suggested here. By way of introduction, we shall prove that for any function ¢
f
H 1 (R~xR;)*, if u = P[¢] we have a Carleson condition with respect
to open sets satisfied by the appropriate measure. To describe this result, we !!Jake the following definition ([23], [24], and [25 ]). Let n ~ R 2 be an arbitrary open set, and let R(y; t) be the rectangle in R 2 centered at (yl'y 2 ) = y and with side lengths 2t 1 and 2t 2 • Then S(fl) the Carleson region above n is defined as
U S(R) = l(y,t) ( (R;) 2 R(y; t) ~ n!.
S(Q) =
1
RCfl
Then we say that 11 ~ 0 in (R;) 2 is a Carleson measure if and only if ll(S(fl))o:;Cifll for every open set n~R 2 ·f (H 1 (R~xR~)* if and only if for u = P[f],
In fact, this follows immediately from the inequality (5.3). To see this, notice that lv't V2 ul is invariant under the Hilbert transform Hx.(i = 1,2) 1
so that if we prove this when f when f is of the form
f
L 00(R 2 ), we will have proven it also
113
MULTIPARAMETER FOURIER ANALYSIS
A function a(x) on R 2 will be in H 1 (R~xR!) if and only if a Hx a, Hx a , and Hx Hx a 1
2
1
2
l
l
L 1 (R2),
L1 .
In fact, if a iH 1 (R~xR;) then S(a)iL 1(R 2) hencesoare S(H
x1
a),
S(Hx a) and S(Hx Hx a); therefore Hx.a,Hx Hx a ( L 1 • Conversely, 2
1
2
1
1
2
if a,Hxia, and Hx 1Hx 2a iL 1 (R 2 ) thenwecanform F++,F+_,F_+, and F __ iL 1 (R 2) suchthat a =~F±± and reflections ofthe F±± are boundary values of hi-analytic functions. A bianalytic function F with (distinguished) boundary values in L 1(R 2) has F* l L 1 by a subharmonicity argument applied to IF Ia, a< 1. So a*
i
H 1 (R~xR;>*.
Define a map from
i
L 1 and a
H 1 (R;xR~) ..P__, ED
i
H 1 . Let
L 1 (R 2)i by
i=1
Then ll~fll
1 "'-' 11£11 1 . ~ is obviously one to one, so t?- 1 = ~ exists eL H and is bounded on lm(~). The map o ~ extends, by Hahn-Banach to
an element of the dual e L 1 = e L 00 • Then
Thus every element of (H 1 )* is of the form
114
ROBERT FEFFERMAN
So it suffices to show that if f
But in (5.3), take g
L 00(R 2) with u
t
"'x 0 (x 1 ,x 2),
[-1 ,1] then ~t g{x) = 1 if (x, t)
l
=
P[f] then
and notice that if
f ¢>
"=
1, supp if>(x)
~
S(O) . This is because for such {x, t),
R(x; t) ~ 0 and g * cf>t(x) = fR2 if>t(x -u)du
=
1. It follows from (5.3) that
Jf117,17,u! t,t,dxdt s cnfll~fflii,.l'l dtt dx 2
S(0)
6. Duality of H 1 and BMO and the atomic decomposition
In this lecture we shall consider in greater detail the spaces HP(R~xR~) and BMO(R~xR!), which we discussed briefly in section 5. There we saw that in product spaces, the most obvious guesses at characterizations of HP atoms of BMO failed. In order to circumvent these difficulties we must take a slightly different approach than we are used to in the classical !-parameter case. In what follows we shall be working with functions in HP(R! xR~) or
BMO(R~xR!> only. The theory for R!xR!
X"'X
R! or for R~tl x R~+l
is quite similar and only requires minor changes. Now let !R be the family of all rectangles with sides parallel to the axes and Sid be the subfamily of 9l whose sides are dyadic intervals. If f( x l'x 2) is a sufficiently nice function on R 2 , and «/!
l
C 00{R 1),
ifJ is even, ifJ i 0 real valued and supp{«/1) ~ [-1 ,1], and «/! has a large
115
MULTIPARAMETER FOURIER ANALYSIS
number of moments vanishing, then for
I ~~(e-),2d~/e00
=
1
0
we have
In fact, taking Fourier transforms of both sides, for the right-hand side we have
We can use this representation to decompose the function f as follows: R
t
!Rd. Set ~l(R) = l(y,t)
f
R;xR;\ y
f
R, f 1 < t 1 ~ 2fi where
fi , i = 1,2 is the side length of R in the xi direction I. Since
R~ xR.!
=
U
W(R), if we define
Rf!Jld
fR(xl,x2)
=
ryf{y,t)I/Jt t (xl -yl'x2-y2)dy
J"
1 2
/!
1 2
~(R)
where f(y,t)
=
hY,t(y), then f =
I
fR, and each fR is supported in
Rt!Rd
"' the double of R and has the property that R
ROBERT FEFFERMAN
116
"' I
JfR(x 11 x 2)dx 2 = 0
for all x 1
"'
J
where ~ = Tx]. It will be convenient to define a norm I IR on functions supported on a rectangle R 1 as follows.
where N is a large integer. With these preliminaries we can pass to a theorem characterizing (H 1)* in a number of useful ways. THEOREM
[25]. For a function on R 2 the following are equivalent:
(1) ¢ t H 1 (R~xR~)*.
(2) ¢ = g 1 + Hx (g 2) + Hx (g 3) + Hx Hx (g 4) for some g 11 g 2,g 3 and 1
2
1
2
g 4 in L ""(R 2).
(3) If u = P[¢] in R! x R~ 1 then
If
lvl \72ul 2(ylt)tlt2dt
~ Clilll
for all open sets
n s; R 2 .
S({l)
(4) If ¢(y 1 t) = ¢
If S(il)
* 1/Jt(y),
l¢(y,t)l 2dy dtt
then
~ CIUII
for all open sets
n c R2.
MULTIPARAMETER FOURIER ANALYSIS
117
l: cRbR where bR(xl'x 2) are
(5) ¢J can be written in the form
RfRd
supported in
~, \bR\R ~ 1
and
. R2 . In
I c~ ~ C\01
R~U
for all n
Proof. To begin with, we proved in section 5 that (1) -=9 (2).
~ (1), since if f
trivial that (2)
f
and since f
f
open
It is also
H 1(R!xR!),
f
f(x)Hx Hx (g)(x)dx =fHx Hx (fXx)g(x)dx 1
H1
,
2
1
2
Hx Hx (f) ( L 1 • 1
2
Next, we recall that (2)
=--~
(3) was also proven in the preceding
section. Now we claim that (3) or (4') implies (1). We show that (4) implies (1), \
the other proof being similar.
We,~o
this via the atomic decomposition of
H 1 which we shall describe here only enough to derive our implication. We shall present the decomposition of H1 in greater detail later. Let f ( H 1 (R~xR~). Then S.p(f) ( L 1 (R 2 ) (this follows by vector iteration, just as in the argument that S(f) ( L 1 implies f ( L 1 ). Consider the sets nk = IS.p(f) ak(x 1 ,x 2) =
> 2k1, k ( Z. Set ~
fR(x).
RdRd
\RnUkl > 1/2\R\ \Rnnk+ 11 < 112\R\ Then, as we shall elaborate later on ~k(x 1 ,x 2 ) H 1(R +2 xR +2 ) atom where
is an
118
Then f
ROBERT FEFFERMAN
=
~ ,\kak where .\k
=
2klflkl, and by the strong maximal theorem
lflkl ~ Cl!lkl so that
~ ,\k ~ CIISr/,(OIILl ~ C'II£11Hl . Now consider ¢(x 1 ,x 2) satisfying (4). Then it will be enough to show that
f
ak(x1,x 2)¢(x1,x2)dx < c
R2
and then simply sum over k. But
(6.1)
where the sum is taken over
Then (6.1) becomes
¢(x)
*
dx. J..ryf(y,t)lfrt(x -y) ~yt 1 2 2
lukl
2{(R)
=J
oo, for each N
> 0.
so
Joo ~~(~)1 2 ~~
(x 2) where qt(i) is odd, C 00 and decreasing at oo like lxii-N (depending on how many moments of if! vanish). Now suppose we choose 7J(x) on R 1 sothat supp(7])~(1/4,4), 71£C 00{R 1), 71 evenand 00
I
k=-oo
11( \)
11(;1{).
2
= 1 . Let 71 0 (x) = I
k 0, let 71k(x) =
Set I/Jk,j(x 1 ,x 2 ) = qt{l)(x 1)71k(x 1)· '1'< 2 >(x 2)71k(x 2). Then
122
ROBERT FEFFERMAN
(a) supp('l'k,j)
s; 4R(O; 2k,2j)
(b) 'l'k,j is odd in each variable separately (c) 'l'kj is c;(R 2 ) and
By Minkowski's inequality, we have
Now, to estimate
we use the same argument as that given·above, except that now supp('l'k,j) ~ 4R(O; 2k,2j) and not the unit square. If (y,t) R(y; t)
Thus
s; {}
£
S(O), then
and the support of ('l'k,j)t( · -y) will be contained in
123
MULTIPARAMETER FOURIER ANALYSIS
By the strong maximal theorem ll)kj I :S C(k+j)2k+j_; and it is easy to see that
J0oo f0
00
1Wkj1 2
d~
decreases like a large power of T(k+j) as
k,j ... ""· So our desired estimate follows from (6.2). So far we have proven the equivalence of (1), (2), (3) and (4). We shall not go into the details of the equivalence of (5) except to say the proof is given in the Annals paper of Chang-Fefferman [25]. Rather, let us point out a beautiful application of the equivalence of (5) with the other definitions of BMO which occurs already in the one-parameter setting. This is the theorem of A. Uchiyama [26], which tells us which families of multipliers homogeneous on Rn of degree 0 determine H1 (Rn). He showed that for multiplier operators I, Kl' K2 .-··, Km with multipliers
1, Oi(O that f, Kif E L 1(Rn) implies f
l
H1(Rn) if and only if the Oi
separate antipodal points of sn- 1 , i.e., if and only if for every ~ l sn- 1 , there exists i such that Oi(~) i Oi(-4). The way Uchiyama proves this is to show that the dual statement is true, namely, every ¢ EBMO(Rn) can be written as
This depends on a simple lemma. LEMMA.
If
()i
are as above, then given f
there exist functions g 0,···, gm
l
f
L 2 , and a vector
II l
cm+ 1 ,
L 2 so that
To prove the formula (6.4), Uchiyama decomposes ¢=I C 1 ¢ 1 as in our (5), and applies the lemma to get functions g 1(x) such that Kg 1(x) C1 ¢ 1(x) for which g(x) is perpendicular to the correct when modified ooly slightly to
g1 ,
11,
has the property that I
=
and the result,
g1 l
L"". For
the details see Uchiyama's recent paper in Acta [26]. Now, finally we wish to discuss the atomic decomposition of H1(R,!xR.;>
in greater detail. There are interesting applications of this
124
ROBERT FEFFERMAN
decomposition besides duality with BMO(R~xR~) which was presented above. We shall be content with one more application here which sheds a good deal of light on the nature of these atoms. Namely, we intend to give a second proof, directly by real variables, that on R~ x R~ if S.p(f) c L 1(R 2) then f* c L 1(R 2) [27). Suppose S.p(f) c L 1(R 2). Then in our discussion of duality we defined atoms
To simplify this notation we define w = Ok and A(x) = 2klfik\ak(x). Let
~ 1 and ~ 2 c C;'(R 1) with ~i(x) ~ 0,
f ~i
= 1 , and supp(~i) C [-1, +1).
Set
Define W = M< 2 ><xw) > - 1- • We need to estimate A * ~t t (x) for 1010 1' 2 x I (;). To do this let us make the following definitions. If R is a rectangle then R 1 , R 2 will be its sides, so that R = R 1 xR 2 . Let
~
A:(x)=
A~(x)
fR(x),
=
J
Rc9lw \R 1 \=2j
Then to estimate A
* ~t
t (x), since supp(~t( · -x)) s; R(x; t) = S, in
1' 2
the definition of A we need only consider those f R for which For any such rectangle R dRk, since R C w,
where (;) denotes again
M<x
w
)>-
1- • Then 1010
Rn S I
0.
MULTIPARAMETER FOURIER ANALYSIS
A* cf>t t (x) 12
=
I,
. 2 3;js1 lt t (x)
I,
t-
12
. 2 1 ;js 2 lt t (x)
fR * cf>t t (x)
m
REv
R'ns .f. 0.
125
IR21
IS 2 I < P
and
18, and the reason this subtracted
term occurs is that we have double counted these fR whose R sides are both very small. In order to estimate A~* cf>t t (x) we use the following trivial lemma. 1
1 2
LEMMA. On R 1 suppose that c/>(x)
f
C 00 and is supported in an interval
{}, Suppose a{x) is supported on disjoint subintervals of {}, Ik whose lengths are all ~ yj{}j. Assume that a(x) has N vanishing moments over each Ik. Then
f
a{x)c/>(x)dx
~ Cllc/>(N+l)lloo(yj{}j)N+ 1 Jla{x)jdx.
{}
We estimate A~ * cf>t t (x) using the fact that for each fixed x 2 , 1 1 2 A~(· ,x 2 ) has N vanishing moments over disjoint x 1 intervals over 1
.
length 2 · 21. (Actually, we sould have to break up A~ into 3 pieces to 1
insure this, but we spare the reader this trivial complication.) It follows from the lemma that
Convolving in the x 2 variable, we have
lA~
* cf>t t (x)l;:; C 1 1' 2
21. I N +1 51
1
~ lSI
f 'S
IA~Idx'. 1
126
ROBERT FEFFERMAN
For this we get
By symmetry
and also
Thus
To sum up our findings, we have seen that if x I 6> then
127
MULTIPARAMETER FOURIER ANALYSIS
(6.5)
lA
* cPt
t (x)l ~ C 1 2
I ~ Sncui)N/4~1 IS I IS I S' ~
""
J
J(y)dy
where
To finish the proof, we need another lemma: LEMMA.
Let g(x)
I fR(x) where B is a collection of dyadic
=
RdB rectan~les.
Then
Proof. Let llhll
J
g(x)h(x)dx
2
2 =
L (R )
=f I,
1 . Then
LJf(y,t)l/lt(x-y)dy
R£~
1 2 ~(R)
=
I
RdB
:~
ryf(y,t)h(y,t)dy
J.,
l"CR)
td~ 1 2
· h(x)dx
128
ROBERT FEFFERMAN
Now, notice that, by the lemma,
IIJII~ ~
If
\f(y,t)\ 2 dy
t~tt 2 ~
Rf~R)
J S~(f)(x)dx ~
C ·2 2 k·lwl.
fik;flk+l
The same estima.te holds for \lA II~ . Then
Also away from
W
so
~ Clwi 112 iwl 112 2k = C2klwl. It follows that IIA *11 1
~ C2klwl and also llakll 1 ~C.
ROBERT FEFFERMAN DEPARTMENT OF MATHEMATICS UNIVERSITY OF CHICAGO CHICAGO, ILLINOIS 60637
BIBLIOGRAPHY [1]
B. Jessen, J. Marcinkiewicz and A. Zygmund, Notes on the Differentiability of Multiple Integrals, Fund. Math. 24, 1935.
[2]
E. M. Stein and S. Wainger, Problems in Harmonic Analysis Related to Curvature, Bull. AMS. 84,1978.
MULTIPARAMETER FOURIER ANALYSIS
129
[3]
A. Cordoba and R. Fefferman, A Geometric Proof of the Strong Maximal Theorem, Annals of Math., 102,1975.
[4]
J. 0. Stromberg, Weak Estimates on Maximal Functions with Rectangles in Certain Directions, Arkiv fur Math., 15, 1977.
(5]
A. Cordoba and R. Fefferman, On Differentiation of Integrals, Proc. Nat. Acad. of Sci., 74, 1977.
[6]
A. Nagel, E. M. Stein, and S. Wainger, Differentiation in Lacunary Directions, Proc. Nat. Acad. Sci., 75, 1978.
[7]
A. Cordoba, Maximal Functions, Covering Lemmas, and Fourier Multipliers, Proc. Symp. in Pure Math., 35, Part I, 1979.
[8]
F. Soria, Examples and Counterexamples to a Conjecture in the Theory of Differentiation of Integrals, to appear in Annals of Math.
[9]
B. Muckenhoupt, Weighted Norm Inequalities for the Hardy Maximal Function, Trans. of the AMS, 165, 1972.
[10] R. Hunt, B. Muckenhoupt, and R. Wheeden, Weighted Norm Inequalities for the Conjugate Function and Hilbert Transform, Trans. AMS, 176, 1973. [11] R. R. Coifman and C. Fefferman, Weighted Norm Inequalities for Maximal Functions and Singular Integrals, Studia Math., 51, 1974.
[12] M. Christ and R. Fefferman, A Note on Weighted Norm Inequalities for the Hardy-Littlewood Maximal Operator, Proceedings of the AMS, 84, 1983. [13] R. Fefferman, Strong Differentiation with Respect to Measures, Amer. Jour. of Math., 103, 1981.
[14]
, Some Weighted Norm Inequalities for Cordoba's Maximal Function, to appear in Amer. Jour. of Math.
[15] C. Fefferman, The Multiplier Problem for the Ball, Annals of Math., 94, 1971. [16] A. Cordoba and R. Fefferman, On the Equivalence between the Boundedness of Certain Classes of Maximal and Multiplier Operators in Fourier Analysis, Proc. Nat. Acad. Sci., 74, No. 2,1977. [17] E. M. Stein and G. Weiss, On the Theory of HP Spaces, Acta. Math., 103, 1960. [18] C. Fefferman and E. M. Stein, HP Spaces of Several Variables, Acta Math., 129, 1972. [19] D. Burkholder, R. Gundy, and M. Silverstein, A Maximal Function Characterization of the Class HP, Trans. AMS, 157, 1971.
130
ROBERT FEFFERMAN
[20] K. Merryfield, Ph.D. Thesis: HP Spaces in Poly-Half Spaces, University of Chicago, 1980. [21] R. Gundy and E. M. Stein, HP Theory for the Polydisk, Proc. Nat. Acad. Sci., 76, 1979. [22] L. Carleson, A Counterexample for Measures Bounded on HP for the Bi-Disc, Mittag-Leffler Report No.7, 1974. [23] S. Y. Chang, Carles on Measure on the Bi-Disc, Annals of Math., 109, 1979. [24] R. Fefferman, Functions on Bounded Mean Oscillation on the Bi-Disc, Annals of Math., 10, 1979. [25] S. Y. Chang and R. Fefferman, A Continuous Version of the Duality of H 1 and BMO on the Bi-Disc, Annals of Math., 1980. [26] A. Uchiyama, A Constructive Proof of the Fefferman-Stein Decomposition of BMO(Rn), Acta. Math., 148, 1982.
ELLIPTIC BOUNDARY VALUE PROBLEMS ON LIPSCHITZ DOMAINS Carlos E. Kenig*
Dedicated to the memory of jack P. Burke PREFACE This paper is an outgrowth of a series of lectures I presented at the Summer Symposium of Analysis in China (SSAC), held at Peking University in September, 1984. The material in the introduction and parts (a) and (b) of Section 1 comes from the expository article 'Boundary value problems on Lipschitz domains' ([19]), which I wrote jointly with D. S. Jerison in 1980. The rest of the paper can be considered as a sequel to that article. Some of the material in part (b) of Section 2, and all d. Section 3 comes from the recent expository article "Recent progress on boundary value problems on Lipschitz domains'' ([23] ). The results explained in Section 2, (b) and Section 3 are unpublished. Full details will appear elsewhere in several joint papers. Acknowled~ements.
I would like to thank Peking University, and the
organizing committee of the SSAC, Professors M. T. Cheng, S. L. Wang, S. Kung, D. G. Deng and R. Long for their invitation to participate in the SSAC, and for their warm hospitality during my visit to China. I would also like to thank Professor E. M. Stein for his many efforts to make the SSAC a success. Thanks are also due to Mr. You Zhong and Mr. Wang Wengshen for taking careful notes of my lectures. Finally, I would like to thank B. Dahlberg, E. Fabes, D. Jerison and G. Verchota for the many discussions and fruitful collaborations that we
*Supported in part by
the NSF.
131
132
CARLOS E. KENIG
have had throughout the years, which resulted in the work explained in this paper.
Introduction A harmonic function u is a twice continuously differentiable function on an open subset of Rn, n ~ 2, satisfying the Laplace equation n
L\u
=
l:
2
au
j=l ax~
= 0. Harmonic function arise in many problems in mathe-
J
matical physics. For example, the function measuring gravitational or electrical potential in free space is harmonic. A steady state temperature distribution in a homogeneous medium also satisfies the Laplace equation. Moreover, the Laplace equation is the simplest, and thus the prototype, of the elliptic equations, or systems of equations. These in turn also have many applications to mathematical physics and geometry. A first step in the understanding of this more general situation is the study of the Laplacian. This will be illustrated very clearly later on. Initially we will be concerned with the two basic boundary value problems for the Laplace equation, the Dirichlet and Neumann problems. Let D be a bounded, smooth domain in Rn and let f be a smooth (i.e. C"") function on aD, the boundary of D. The classical Dirichlet problem is to find and describe a function u that is harmonic in D,
continuous in
D, and equals f on aD. This corresponds to the problem
of finding the temperature inside a body D when one knows the temperature f on iJD. The classical Neumann problem is to find and describe a function u that is harmonic in D' belongs to C 1 (D), and satisfies
~
=
f on
ao'
where
~ represents the normal derivative of u on
ao.
This corresponds to the problem of finding the temperature inside D when one knows the heat flow f through the boundary surface
ao.
Our main purpose here is to describe results on the boundary behavior of u in the case of smooth domains, and to study in detail the extension
ELLIPTIC BOUNDARY VALUE PROBLEMS
133
of these results to the case of minimally smooth domains, where we allow corners and edges, i.e. Lipschitz domains. This class of domains is important from the point of view of applications, and also from the mathematical point of view. Their importance resides in the fact that this is a dilation invariant class of domains with some smoothness. They have the borderline amount of regularity necessary for the validity of the results we are going to expound on. In a smooth domain, the method of layer potentials, (which we are going to develop soon) yields the existence of a solution u to the Dirichlet problem with boundary data f ( ck,a(aD), and the bound
k
=
0,1,2,···
O(y')l
s Mjx'-y'j
and DnB = IX=(x',xn):xn>c/>(x')lnB.
H for each Q the function cf> can be chosen in C 1 (Rn- 1), then D is ca lied a C 1 domain. If in addition,
llc/>
satisfies a Holder condition
of order a, lllc/>(x')-llcf>(y')i :S·Cix'-y'la, we call D a C 1•a domain. Notice that the cone re ~ I (x ',xn): xn < -M lx'll satisfies re n B
c co.
Similarly, ri ~ {(x ',xn): xn >Mix'il satisfies ri n BCD. Thus, Lipschitz domains satisfy the interior and exterior cone condition. The function ¢ satisfying the Lipschitz condition i¢(x ') -- cf>(y') I ~ Mjx'-y'l is differentiable almost everywhere and 17¢
l
L ""(Rn- 1 ), llllc/>lloo :SM.
136
CARLOS E. KENIG
Surface measure o is defined for each Borel subset E C aD n B by
o(E) =
J
(1 + \'Vcp(x')\ 2 ) 1 12 dx',
E*
where E* =lx': (x',cp(x'))lEI. The unit outer normal to !D given in the coordinate system by ('i]cp(x '), -1 )/(1 + \'Vcp(x ')\ 2 / 12 exists for almost every x'. The unit normal at Q will be denoted by NQ. It exists for almost every Q
(a>,
with respect to do. In order to motivate the use of the method of layer potentials, we need to recall some formulas from advanced calculus, and some definitions. We will start with the derivation corresponding to the Dirichlet problem. We first recall the fundamental solution F(X) to Laplace's equation in Rn: ~F = 5, where
1
n >2
(n-2) wn \X \n-2 F(X) =
lrr log \X\
n=2
where wn is the surface area of the unit sphere in Rn. F(X) is the electrical potential in free space induced by a unit charge at the origin. It provides a formula for a solution w to the equation tlw = 1/J , with
0
1/J c C (Rn),
w(X) = F *1/J(X) =
f
F(X-Y)I{I(Y)dY.
Rn It will be convenient to put F(X,Y) ~ F(X-Y). Notice that tl.yF(X,Y) =
8(X-Y). The fundamental solution in a bounded domain is known as the Green function G(X,Y). It is the function on
DxD
continuous for X ~ Y
and satisfying tlyG(X,Y) = 8(X-Y), X c D; G(X,Y) = 0, X £ D, Y £!D.
137
ELLIPTIC BOUNDARY VALUE PROBLEMS
G(X,Y) as a function of Y is the potential induced by a unit charge at X that is grounded to zero potential on aD. The Green function can be obtained if one knows how to solve the Dirichlet problem. In fact, let ux(Y) be the harmonic function with boundary values ux(Y)\an = F(X,Y)\ao. Then, G(X,Y) = F(X,Y)-ux(Y).
(1)
On the other hand, if we know G(X,Y), we can formally write down the solution to the Dirichlet problem. In fact, u(X) =
=
J
J
D
D
u(Y)o(X,Y)dY =
f [u(Y)~yG(X,Y)
u(Y)AyG(X,Y)dY =
-Au(Y)· G(X,Y)]dY
=
D
J[u(Q)
~Q (X,Q)- ~Q (Q)G(X,Q~ du(Q)
=
ao
=
I
u(Q)
a;Q G(X,Q)du(Q)'
ao
where the fourth equality follows from Green's formula. Thus, we have derived the formula
(2)
u(X) =
J
f(Q)
:;Q
(X,Q)du(Q)
an
for the harmonic function u with boundary values f. The problem with
138
CARLOS E. KENIG
formula (2) is of course that we don't know G(X,Q). Because of formulas (1) and (2), C. Neumann proposed the formula
w(X) =
J :Q J f(Q)
(X,Q)da(Q) =
1
f(Q)
= (un
<X-Q,NQ> IX--Q I" da(Q)
av as a first approximation to the solution of the Dirichlet problem, L\u in D,
=
0
ulao =f.
w(X) is known as the double layer potential of f. First of all w is harmonic in D. Also, one can show that as X .... Q t CD, w(X) ....
t
f(Q) +
Kf(Q), where K is the operator on CD given by
Kf(Q)
=
J-n
J ao
I 'I"P f(P)da(P) . P-Q
If Kf were zero, we would be done, and in some sense it is true that Kf
is small compared to
~ f, when the domain D is smooth. In fact, CD
has dimension n-1 , while it is easy to see that if CD is C 00 ,
Thus, the operator K: C(CD) .... C(I , and
~ Cpllfll
What do we do when
.
LP(da)
ao
is merely C 1 , or even merely Lipschitz?
As I mentioned before, the LP boundedness of K is even in doubt. In 1977, A. P. Calderon ([1]) showed that for any C 1 ·domain, K: LP(OD) .... LP(OD), 1 < p < oo is a bounded operator. Shortly afterwards, Fabes, Jodeit and Riviere ([11]) showed that K is in fact compact in this case. They were thus able to extend the theorems above to the case of C 1 domains. Before going on to the main subject matter of this paper, i.e. the method of layer potentials on Lipschitz domains, I want to discuss another important method for the Dirichlet problem for Laplace's equation. (b) The method of harmonic measure Another way of studying the Dirichlet problem for Laplace's equation is in terms of the notion of harmonic measure. Let D be a bounded Lipschitz domain in Rn. As we mentioned before, then D satisfies the exterior cone condition, and so, by a classical result ci. Zaremba and Lebesgue, we can solve the classical Dirichlet problem for !l in D, for any f tC(aD). Given X mapping f
1-->
t
D, the maximum principle shows that the
u(X) defines a positive continuous linear functional on
C(OD). Therefore, by the Riesz representation theorem, there is a unique positive Borel probability measure wx on
u(X) =
J
f(Q)dwx(Q)
an
an
such that
142
CARLOS E. KENIG
wx is called the harmonic measure for D, evaluated at X. For example, harmonic measure for the unit ball B , evaluated at the origin is a constant multiple of surface measure: w 0 = u/o(aB). (This follows from the mean value property of harmonic functions.) For different X, the measures wx are mutually absolutely continuous (a simple consequence of Harnack's principle). We fix X 0
X
t
D, and denote w = w 0 • The importance of
harmonic measure to the boundary behavior of harmonic functions on Lipschitz domains can be illustrated by the following theorem of Hunt and Wheeden (1967): If u is a positive harmonic function in a Lipschitz domain D, then u has finite non-tangential limits almcst everywhere with respect to w. Conversely, given any set E C aD, with w(E) = 0, there is a positive harmonic function u in D with lim u(X) X
-->
Q, for every Q
t
= + oo as
E. Despite its advantages, harmonic measure has
some inherent difficulties. First, it is hard to calculate it explicitly. Second, it is tied up to the maximum principle, positivity, and the Harnack principle, and so it is not useful for the Neumann problem, or for the Dirichlet problem for systems of equations. In general, harmonic measure may be very different from surface measure. H D is a C 1 •a domain, then harmonic measure and surface measure are essentially identical in that each is a bounded multiple of the other. This can be proved by the classical method of layer potentials. Along the same lines, as we saw before, one can use layer potentials to solve the Dirichlet and Neumann problems with boundary data in LP. On C 1 domains, it is no longer true that harmonic measure is a bounded multiple of surface measure, or vice versa. Moreover, as was explained before, the applicability of the method of layer potentials is not obvious. The situation for general Lipschitz domains is even less obvious. In 1977, B. E. J. Dahlberg ((4]) proved that on a C 1 or even a Lipschitz domain, harmonic measure and s•uface measure are mutually absolutely continuous. Using a quantitative version of mutual absolute continuity, and the theory of weighted norm inequalities, he proved ([5])
ELLIPTIC BOUNDARY VALUE PROBLEMS
143
that in a Lipschitz domain D one can solve the Dirichlet problem as in the theorem above with f
£
L 2 (ao, da). In fact, he showed that given a
Lipschitz domain D, there exists e = e(D) such that this can be done for f
£
LP(ao, da), 2-e 'S p 'S
oo.
Also, simple examples to be presented
later show that given p < 2, we can find a Lipschitz domain D for which this cannot be done in LP. By establishing further properties for harmonic measure on C 1 domains, he was able to show the results above in the range 1
< p < oo for C 1 domains. (The best possible regularity
result for harmonic measure on C 1 domains is due to D. Jerison and C. Kenig (1981): if k =
~~, then log k £ VMO(dD) .)
A shortcoming of Dahlberg's method of proof, as was explained before, is that, by studying harmonic measure, it relied on positivity and the Harnack principle. This made the method inapplicable to the Neumann problem, or to systems of equations. Also the method does not provide useful representation formulas for the solution. (c) The method of layer potentials revisited In 1979, D. Jerison and C. Kenig [16], [17] were able to give a simplified proof of Dahlberg's results, using an integral identity that goes back to Rellich ([30]). However, the method still relied on positivity. Shortly afterwards, D. Jerison and C. Kenig ([18]) were also able to treat the Neumann problem on Lipschitz domains, with L 2 (aD,da) data and optimal estimates. To do so, they combined the Rellich type formulas with Dahlberg's results on the Dirichlet problem. Thus, it still relied on positivity, and dealt only with the L 2 case, leaving the corresponding LP theory open. In 1981, R. Coifman, A. Mcintosh andY. Meyer [2] established the boundedness of the Cauchy integral on any Lipschitz curve, opening the door to the applicability of the method of layer potentials to Lipschitz domains. This method is very flexible, does not relie on positivity, and does not in principle differentiate between a single equation or a system of equations. The difficulty then becomes the solvability of the integral
144
CARLOS E. KENIG
equations, since unlike in the C 1 case, the Fredholm theory is not applicable, because on a Lipschitz domain operators like the operator K in part (a) are not compact, as simple examples show. For the case of the Laplace equation, with L 2 (aD, do) data, this difficulty was overcome by G. C. Verchota ([33]) in 1982, in his doctoral dissertation. He made the key observation that the Rellich identities mentioned before are the appropriate substitutes to compactness, in the case of Lipschitz domains. Thus, Verchota was able to recover the L 2 results of Dahlberg [S] and of Jerison and Kenig [18], for Laplace's equation on a Lipschitz domain, but using the method of layer potentials. These results of Verchota's will be explained in the first part of Section 2. In 1984, B. Dahlberg and C. Kenig ([16]) were able to show that given a Lipschitz domain DC Rn, there exists e = e(D) > 0 such that one can solve the Neumann problem for Laplace's equation with data in LP(OD, do), 1 < p::; 2 +e. Easy examples (to be presented later) show that this range of p's is optimal. Moreover, they showed that the solution can be obtained by the method of layer potentials, and that Dahlberg's solution of the LP Dirichlet problem can also be obtained by the method of layer potentials. They also obtained end point estimates for the Hardy
space H 1(aD,do), which generalize the results for n [21 ], and for
C1
=
2 in [20] and
domains in [12]. The key idea in this work is that one
can estimate the regularity of the so-(x)){x)) or simply by x. Nx or NQ will denote the unit normal to ao =A at Q = (x, c/>{x)). If u is a function defined on Rn\A and Q (
ao,
lim
u ±(Q) will denote
u{X) or
X-->Q
xtri lim
u(X), respectively. If u is a function defined on D, N(u)(Q) =
X->Q
xtre(Q) sup \u(X)\. xtri We wish to solve the problems (D)
~Au = 0 { u\ao
=
Au= 0 in D
in D f
t
(N) {
L 2 (aD, du)
~I
ao
= f
t:
L 2(aD,du)
The results here are THEOREM 2.1.1. There exists a unique u such that N(u)
£
L 2(aD, da),
solving (D), where the boundary values are taken non-tangentially a. e .. Moreover, the solution u has the form
u(X) =;
n
for some g c L 2(aD,da).
J ao
<X-Q,NQ> I In g{Q)da(Q) , Q-X
147
ELLIPTIC BOUNDARY VALUE PROBLEMS
THEOREM
that N(V'u)
2.1.2. There exists a unique f
u
tendin~
to 0 at
oo,
such
L 2 (aD,da), solving (N) in the sense that No·Vu(X) ... f(Q)
as X ... Q non-tangentially a.e. . Moreover, the solution u has the form
u(X)
=
f
2)
/
"'n n-
1 IX-Q ln-2
g(Q)da(Q) ,
dD for some g
E
L 2 (ao, da).
In order to prove the above theorems, we introduce the double and single layer potentials
J
<X-Q,N >
Kg(X) = Jn
IX-Qjl?
g(Q)da(Q)
dD and
Sg(X) = -
l
cun n
2)
f
1
IX-Qin-2
g(Q)da(Q) .
dD If Q = (x, ¢(x)), X = (z ,y), then
Kg(z,y) =
J n
Sg(z,y)
=-
J Rn-1
1
cu {n-2)
n
THEOREM 2.1.3. a) If g
l
y-¢(x)-(z-x). V¢(x) g(x)dx llx-zl2 + [¢(x)-¢(z)]2]n/2
I
Jl + IV¢(x)l2
Rn- 1
llx-zl 2 + l¢(x)-y]2] 2
LP(di), da), 1 < p < oo,
n-2 -
g(x)dx.
then N('i7Sg), N(Kg)
also belong to LP(ao, da) and their norms are bounded by
Cll&ll
LP(dD,da)
.
148
{b)
CARLOS E. KENIG
lim
...!..
e-+0
e
a. e. and \\Kg\\
1 . 11m (;) e _.. 0 n
cf>(z) -~(x)-{z-x) · V~(x) g(x)dx = Kg(z) exists [\x-z\2 +[~{x)-¢{z)]2]n-2
LP(dD,do)
J -
~ C \\gl\
Lp(dD,do)
1
< p e
[\z-x 12 + [~(z) _ ~{x)]2]n/2
LP(e COROLLARY
2.1.4. {Nz\]Sg)±{z)
L 2 (il>, do) adjoint of
=
+} g{z)- K*g{z),
where K* is the
K.
The proof of Theorem 2.1.3 a) follows by well-known techniques from the deep theorem of Coifman, Mclntos h and Meyer ([2 ]). THEOREM
([2]). Let (}: R _.. R be even, and CDC>. Let A,B: Rn- 1 ... R
be Lipschitz. Let K(z,x) = A(z)-A(x) \z-xln operator
M*g(z) =sup E-+0
I
ofB(z)-B(x)l. [ lz-x\ J
I
K(z,x)g{x)dx\
\t-x\>e
is bounded on LP(Rn- 1 ), 1 < p < oo, with
Then, the maximal
•
149
ELLIPTIC BOUNDARY VALUE PROBLEMS
\I 'VAll""~ M, ll\7811"" ~ M.
where C = C(M,O,n,p), and
The proof of (b), (c) follow from the theorem above, together with the following simple lemmas. LEMMA.
If f c C':)'(Rn-l), then
lim _!_ e .. own
I
¢(z)-¢(x)-(z-x). \7¢(x) f(x)dx = [jx-zl2+[¢(x)-¢(z)]2]n/2
lz-x\>e
=-~_!_I ~ wn 1
zk-xk A
Iz-x 1n-1
[¢(z)-¢(x)~ lx-zl
L(x)dx' axk
where A(O) = 0, A'(t) = (1 +t 2)-n/ 2 , and where the equality holds at every z at which ¢ is differentiable, i.e. for a.e.z. LEMMA.
If a c Rn- 1 , a
.J ¢(a), f r C';;'(Rn- 1) and A is as in the
previous lellUlla, then
_!_f a-¢(a)- {a-x). 'V¢(x) wn
=
f(x)dx =
[Ja-xl2+[¢(x)-a]2]n/2
_21 sign (a-¢(a))f(a)- w1 n
f~
xk-ak A
~ lx-ajn-1
{¢~x) -j\ 1!
\
x--a
J uxk
(x)dx .
Moreover, the integral on the right-hand side of the equality is a continuous function of {a,a) c Rn. It is easy to see that (at least the existence part) of Theorems 2.1.1 and 2.1.2 will follow immediately if we can show that {~I+
K)
and
150
t
CARLOS E. KENIG
are invertible on L 2 (aD,do). This is the result of
I+K*
G. Verchota ([331). THEOREM 2.1.5.
(±~I+ K)
,
(± ~ I+ K*)
are invertible on L 2(aD, do).
In order to prove this theorem, it suffices to show that (±
~
I+ K*)
are invertible. In order to do so, we show that if f f L 2(aD, do), I (2!. I+ K*\ fll 2 ~ 11(-12 I- K*) fll 2 , where the constants } L (liD,do) L (ao,do) of equivalence depend only on the Lipschitz constant M. Let us take this for granted, and show, for example, that ~ I+K* is invertible. To do this, note first that if T
=
~ I+ K* , IITfll L 2 ~ C llfll L 2 , where C
depends only on the Lipschitz constant M. For 0::; t operator Tt =
~ I+
K:, where K:
~
1, consider the
is the operator corresponding to the
aT
domain defined by t¢. Then, T 0 = ~ I, T1 = T, and (/ : LP(Rn-l) .... LP(Rn- 1), 1 < p < oo with bound independent of t, by the theorem of Coifman-Mclntosh-Meyer. Moreover, for each t, independent of t. The invertibility of T now follows from the continuity method: LEMMA 2.1.6. Suppose that Tt: L 2 (Rn- 1) .... L 2 (Rn- 1) satisfy (a) liT tfll
L
2
> C 1llfll 2 L
(b) IITtf-Tsfll 2 ~C 2 \t-s\\fl\ 2 , O::;t,sSl. L L 1 2 2 (c) T0 : L (Rn-l) .... L (Rn- ) is invertible. Then, T1 is invertible. The proof of 2.1.6 is very simple. We are 1 I+ K*) fll 2 thus reduced to proving (2 .1. 7) II (-2
L (ao,da)
~ II (-21 I- K*) fll L 2 (ao,d [)
In order to prove (2.1.7), we will use the following formula, which goes back to Rellich [30] (also see [28], £29], [27}). LEMMA 2.1.8. Assume that u fLip (D), L\u
=
0 zn D , and u and its
151
ELLIPTIC BOUNDARY VALUE PROBLEMS
derivatives are suitably small at
Then, if en is the unit vector in
oo.
the direction of the y-axis,
fl\7Ul 2 du=2
ao Proof. Observe that div (en1Vul 2) = div
-t; V'u t
V'u · V'u +
=
i·
t
div V'u
=
J~·~do.
ao
!Vui 2 =
2fy \lU · V'u,
~ V'u · Vu.
while
Stokes' theorem now
gives the lemma. We will now deduce a few consequences of the Rellich identity. Recall that Nx
=
(-V'cp(x),l)/Vl + IV'cp(x)! 2 , so that
COROLLARY
1 (l+M2)
11 ::; :S
2
1.
2.1.9. Let u be as in 2.1.8, and let T 1(x), T2(x),
Tn_ 1 (x) be an orthogonal basis lor the tangent plane to aD at (X, cp(X)). n-1
Let IY'tu(x)l 2 = j~1 I 1 2 • Then,
J (~Y ao
da:S
c
J ao
1vtul 2 du.
152
CARLOS E. KENIG
COROLLARY 2.1.10. Let u be as in 2.1.8. Then,
Proof.
fao
IV'ufda::;
2(fao IV'ul 2 da) 112 Van 1~1 2 daf 12 ,
by 2.18,
and the corollary follows.
lnordertoprove2.1.7,let u~Sg. Becauseof2.1.3c, V'tu iscon tinuous across the boundary, whileby2.1.4,
{~}± ={+~
1-K*} g.
We
D, to obtain 1.1.7. This finishes the proof
now apply 2.1.11 in D and of 2.1.1 and 2.1.2.
We now turn our attention to L 2 regularity in the Dirichlet problem. DEFINITION 2 .1.12. f
f
L~(A), 1
< p < oo, if f(x, ifJ(x)) has a distribu-
tional gradient in LP(Rn- 1). It is easy to check that if F is any extension to Rn of f, then V'xF(x,ifJ(x)) is well defined, and belongs to LP(A). We call this V'tf. The norm in Lf(A) will be IIV'tfiiLP(A). THEOREM 2.1.13. The single layer potential S maps L 2 (A) into Lf(A) boundedly, and has a bounded inverse. Proof. The boundedness follows from 2.1.3a). Because of the L 2 -Neumann theory, and 2.1.11, IIV'tS(£)11 2 L
> cji~. C!ifll 2 A . The un L ( ) L ( )
-1 . Fix now a p > 2, and choose {3 so close to 2TT that p TT/{3-p < -1. Then, N('i7w) iLP(aO{:J). If {} I- K*) were invertible in LP(aO{:J), then, since would have that w(z)
=s((} 1-K*)-l(~))(z)
~
€
L 00 (d~), we
has a non-tangential
maximal function in LP(dQ{:J). By the L 2 -uniqueness in the Neumann problem, w-w is constant in 0{3, but this is a contradiction. This shows that given p > 2, we can find a Lipschitz domain so that {} I-- K*) is not invertible in LP. The example can also be used to
154
CARLOS E. KENIG
show that } I+ K is not always invertible in L q , when q < 2. In fact, fix q < 2, and let p satisfy
p~-p 0 such that, tiven
THEOREM
f t LP(Cl), da), 2- e ~ p
with N(u)
t
< oo, there exists a unique u harmonic in D,
LP(aD, da) such that u converges non-tangentially almost
everywhere to f. Moreover, the solution u has the form
Jn
u(X) =
J
<X-Q,NQ> IX-Qin g(Q)da(Q) '
ao for
some g t:
LP(aD,da).
THEOREM 2.2.2.
f
l
There exists e = e(M) > 0, such that, given
LP(aD, da), 2-e ~ p < oo, there exists a unique u harmonic in D,
tending to 0 at
oo,
with N(V'u)
t:
LP(aD, da) , such that NQ V'u(X) con-
verges non-tangentially a.e. to f(Q). Moreover, u has the form
u(X) =
for some g
t
(1 2) wn n-
f ao
g(Q)da(Q) , 1 IX-Q ln-2
LP(aD,da).
THEOREM 2.2.3.
There exists e = e(M) > 0 such that given f
t:
1 < p < 2 + e, there exists a harmonic function u, with IIN(V'u)l! -
CI!V'tfll
LP(
A, and such that V'tu )
=
for some g
£
(1 2) cun n-
LP(aD,da).
J ao
L
P
~)
S
V'tf (a.e.) non-tangentially on A·u
is unique (modulo constants). Moreover, u has the form
u(X) =-
Li(A),
1 2 g(Q)da(Q), IX-Qin-
156
CARLOS E. KENIG The case p - 2 of the above theorems was discussed in part (a). The
first part of 2.2.1 (i.e. without the representation formula), is due to
B. Dahlberg (1977) ([5]). Theorem 2.2.3 was first proved by G. Verchota (1982) ([331). The representation formula in 2.2.1, Theorem 2.2.2, and the proof that we are going to present of 2 .2 .3 are due to B. Dahlberg and
C. Kenig (1984) ([6]). Just like in part (a), 2.2.1, 2.2.2, and 2.2.3 follow from. THEOREM 2.2.4. There exists
e = e(M) > 0 such that (±}I- K*} is
invertible in LP(OO, do), 1 < p ::; 2 + e, ( ±
LP(aD, da), 2-e < p < -
oo,
~ I+ K)
is invertible in
and S: LP(CD, da) ... LP(aD, da) is invertible 1
l(x)+d, and pis large. The right-hand
f
sideequals lim e-->0
harmonicity of u, and
OD~
•
2=
~=lim f e{t/J-1]~,
e
aop
E-->0
faoe 0~ ~.
OD~\ao;
p
1• '
=
aop
since,bythe
0. Let ODpe 1 = l(x,y) c iJDPE : y > c/>(x) + e I, •
Then, lim
f
e [t/J-1]
aop = fao t/Ja- fan
E--> 0
lim Jaoe [t/J-11 ~ = fao [t/J-1]a e ... o p, 2 t/1=0 on suppa. Moreover, J0 _B\7u\7t/I=J0
~
= lim
a=
E-+
0
f
ao~,l
fiKJ t/Ja
{t/J-1] ~ +
= 0, since
\lu·\lt/1*, where c/l*=c/Jo(/),
by our construction of B. The last term is also 0 by the same argument, and so a = 0. We now show that u (and hence
u ) is bounded.
We will
assume that n ~ 4 for simplicity. Since lla II 2 A < C, we know that L ( )
u(X)-= en
fa
£( Q)
D IX-Qin-2
l(x,y):y>¢(x)+1l,
da(Q), with 1\fll 2 A O. For R?:Ro= diamB*, set b(R)= f N('Vu) 2 , where, AR=I(x,¢(x)):R..\l!d..\+Ce
f."" ..\e- 1 (J. 0
f.
lm>..\1
m 2 d..\< -
\ m2 )d..\.
h>~
162
CARLOS E. KENIG
By a well-known inequality (see [14] for example), lEAl~ Callm >All. Thus, fm 2 +e ~ Ce f 0oo A1 +e llm>AIIdA+Ce f000 Ae-l(Jh>A m2)dA ~ Ce fm 2+E+ Cfm 2 he. Ifwenowchoose e0 sothat Ce 0 A,h~l-
that the same argument gives the estimate !IN(VU)llp ~ Cil'Vtullp, 2 < p < 2 + e , and the LP theory is thus completed. §3. Systems of equations on Lipschitz domains (a) The systems of elastostatics.
I"n this part we will sketch the extension of the L 2 results for the Laplace equation to the systems of linear elastostatics on Lipschitz domains. These results are joint work of B. Dahlberg, C. Kenig and G. Verchota, and will be discussed in detail in a forthcoming paper ([8]). Here we will describe some of the main ideas in that work. For simplicity here we restrict our attention to domains D above the graph of a
cp: R 2 ... R.
Lipschitz function Let ..\ > 0, p.
~
0 be constants (Lame moduli). We will seek to
... =
solve the following boundary value problems, where u
i
(ul'u 2 ,u 3)
~\{ + (..\ + p.) 'V div \: = 0 in D
(3.1.1)
...
ulao
f
.... =
f
2
£
L (il>,da)
~t; + (..\ + p.) 'V di v 1: =
0 in D
(3.1.2) { ..\(div \:)N + p.IV\:+(Vi:)tlNlao =
f
£
L 2 (il>,da).
164
CARLOS E. KENIG
(3.1.1) corresponds to lmowing the displacement vector
ii
on the
boundary of D, while (3.1.2) corresponds to knowing the surface stresses on the boundary of D. We seek to solve (3.1.1) and (3.1.2) by the method of layer potentials. In order to do so, we introduce the Kelvin matrix of fundamental solutions (see [24] for example), r(X) = (rij(X)), where A 0 ii
c xi xi TXi + 4rr·IXI3 , and
rii(X) = 4"
1 [1 A = 2 p:
1 1 1 [1 1 ] +_2il.+XJ ' c = ~ p: -z;;x
.
We will also introduce the stress operator T, where Tti' = ~ (div i:) N +
pi \l~ + vi: t I N . The double layer potential of a denc;ity g(Q) is then given by i:(X) = Kg(X) =
fao
IT(Q)r(X-Q)Itg(Q)da(Q), where the operator T is applied
to each column of the rna trix
r.
The single layer potential of a density g(Q) is
ti'(X) = Sg(X) =
J
r(X-Q) · g(Q)da(Q).
ao Our main results here parallel those of Section 2, part a). They are THEOREM
3.1.3. (a) There exists a unique solution of problem 3.1.1 in
L 2 (00, da). Moreover, the solution U' has the form
D, with N( ti')
f
ti'(X) = Kg(X),
g
f
L 2 (00, da).
(b) There exists a unique solution of (3.1.2) in D, which is 0 at infinity,
with N(\]u) Sg(X),
f
L 2 (aD, da).
Moreover the solution ~ has the form ~(X)=
gfL 2 (aD,da). 2
~
(c) If the data f in 3.1.1 belongs to L 1 (00, da), then we can solve
(3.1.1), with N(\7~)
f
L 2 (cD,da). Moreover, we can take ~
~
~
u(X) = Sg(x), g
f
2
L (00, da).
The proof of Theorem 3.1.3 starts out following the pattern we used to prove 2.1.1, 2.1.2 and 2.1.14. We first show, as in Theorem 2.1.3. that the following lemma holds:
165
ELLIPTIC BOUNDARY VALUE PROBLEMS
LEMMA
3.1.4. Let Kg, Sg be defined as above, so that they both solve
j.LAii + (A+p.)'V div ii (a)
(b)
=
0 in R 3\CID. Then:
IINll L P 'I+ p.I \]u.... + \]u->t I) 2 do
t II 2 do .
ao
ao
Lemma 3.1.7 is proved by first doing so in the case when the Lipschitz constant is small, and then passing to the general case by using the ideas of G. David ([9] ). Lemma 3.1.8 is proved by observing that if ~ is any row of the matrix A.(div i:)I+p.I\Ji:+\Jirtl, then ~ is a solutioo of the Stokes system
/'t.~ = \]p in D { (S) div ~ = 0 in D
~lao= 7( L 2 (aD,do) This is checked directly by using the system of equations
p.!'t.\I + (A+p.)\7 div II= 0. One then invo~s the following Theorem of E. Fabes, C. Kenig and G. Verchota, whose proof will be presented in the next section. THEOREM
3.1.9. Given
f ( L 2(0D, do),
there exists a unique solution
(;,p) to system (S) with p tending to 0 at oo, and N(~) ( L 2(aD,do).
Moreover 1\N(;)II 2
L ¢(x)l, and so that T¢ is a bi-Lipschitzian inapping. Also, it is clear that T¢ is smooth for (x,y) with y > 0, and T¢(x,O) = (x, ¢(x)). We will den~te by A¢ the point T¢(0,1). Lemma 3.1. 7 is an easy consequence of LEMMA 3.1.11. Given M > 0 and ¢ with l\IV' ¢111 ~ M, there exists a constant C = C(M) such that lor all functions t; in D¢, which are Lipschitz in 5¢, which satisfy ~t; +(A+ p.)\7 div t; vii(A,J.) = V'ii(A,J.)t' we have IIN,J.(\i'ii)\1 2 'fJ
~tl vii+ vtrt Ill
'fJ
2
L
'fJ
O, e>O, a€(0,0.1), ifProposition (M,e) holds, then Proposition ((1-a)M,l.le) holds.
We postpone the proof of Proposition 3.1.13, and show first how Proposition 3.1.12 and Proposition 3.1.13 yield Lemma 3.1.11.
Proof of Lemma 3.1.11. We will show that Proposition (M,e) holds for any M,e. Fix M,e, and choose R so large that if e(10M) is as in Proposition R
3.1.12, then (l.l)Re(lOM)? e. Pick now aJ· > 0 so that
.n
(1-aJ·)=l/10.
J=l
Then, since Proposition (10M, e(lOM)) holds, by Proposition 3.1.12, applying Proposition 3.1.13 R times we see that Proposition (M,e) holds. We will not sketch the proof of Proposition 3.1.13. We first note that it suffices to show that
where
Ncp
is the nontangential maximal operator with a wider opening of
the non-tangential region. This follows because of classical arguments relating non-tangential maximal functions with different openings (see [14]
174
CARLOS E. KENIG
for example). Pick now ¢ with 111\7¢-i Ill~ 1·le, llli Ill :S (1-a)M. We will choose N¢ as follows: Since ao¢\B¢ is smooth, it is easy to see that we can find a finite number of coordinate charts (i.e. rotates, translates and dilates of
D.p ),
which are entirely contained in D¢ ,
such that their bottoms B.p are contained in aD¢, such that T.p((x,O): lllxlll < 1/2) cover (/JJ¢• and such that the o/'s involved satisfy
IIIVI/1111
S.
(t- f) M, and there exist
fl.p
:S (x,O): 1\\x\\1 < ~ )
such that llli.plll
l.lle . The non-tangential region defining N¢, on T.p ( is defined as follows: let F C {<x,O): \llxlll < ~} sider the cone on R~. y = l<x,y)
f
R~: b \xI
be a closed set. Con-
< y I, where b is a small
constant. Consider now the domain DF on R~, given by DF =
U ((x,O)+y). Then DF is the domain above the graph of a xfF
Lipschitz function (), for which IIIV Olil s_ Cb, for some absolute constant C (independent of F ). It is also easy to see that we can take now b so small (depending only on M and e )
~hat
domain above the graph of a Lipschitz function
if!,
which satisfies
IIIV~III S.
(1-
to) M,
tangential region defining N¢, for Q
Tl/1
~x,O):
if!
with
IIIVI/1-a".plll f
T .p(f~,_F) is the ~
:S l.llle.
lllxlll
0, consider the open-set Et ~ lm > tl. We now produce a Whitney type decomposition of Et into a family of disjoint sets IUj I with the property that each Uj is contained in T.p (<x,O): 1\\xl\l < for a coordinate chart cube in
D.p,
t)
each Uj contains T.p(lj), where Ij is a
11\x \\1 < ~ , and is contained in T.p(lj), where 1j is a fixed
multiple of Ij. Finally, we can also assume that there exists a constant 'Tlo suchthatif diam(Uj)S.Tlo• thereexistsapoint Qj in ao¢, with
dist (Qj ,Uj) ~ diam Uj, such that m(Q j) s_ t. Let now f3 > 1 be given. We claim that there exists
o> 0
so small that if Ej
=
Uj
n lm > {3t, iii S. otl
175
ELLIPTIC BOUNDARY VALUE PROBLEMS
then a(Ej)
s (1-17M)a(Uj)
where 17M> 0. Assume the claim for the time
being. Then
J-
Joo ta(Ef!t)dt ~ f 2{3
ta(E,ldt - 2{3 2
0
J
2
J 00
ta(Ej)dt + 2{3 2
ta(m>8t)dt S
~ 2{3 2 (1-17M)
0
0
ta(Uj n E{3t)dt S
0
0
00
J 00
J 00
ta(Uj)dt +
0
J
Thus, if
we chocse {3 > 1 , but so that {3 2 ·(1-17M) < 1 , the desired result follows. It remains to establish the claim. We argue by contradiction. Suppose not, then a(Ej) >(1-17M) a(Uj). Let Ej
=
Tt(Ej). If 17M is chosen suffi-
ciently small, we can guarantee that IEj nijl? ·99~1- Let now Fj = Ej n Ij, and construct now the Lipschitz function as
i~ the definition of N¢.
lll'iJ«/J-a«/1111 S l.llh.
Thus,
~ :2: «/!,
=
lxdj
:f=~l.
We now apply Lemma 3.1.10"'to
then !FjnFjl ?C a(Uj), with
and such that there exists af, with lll'iJf-aclll S
corresponding to it,
lll'iJ;III S
at a time, to find a Lipschitz function f, with f Fj
«/!
t l.llle <E.
ill acill S
~
(~ -f0 ) M, «/!,
one variable
«/! on Ij, such that if III'Vf\11
s
(1
-f0)M,
(1 - laO) M so that
We can also arrange the truncation of our non-
tangential regions in such a way that on the appropriate rotate, translate and dilate of Df (which of course is contained in the corresponding coordinate chart associated to IA(div u)I t
11! 'iJu + 'iJut II S 8t.
D«/1,
which is contained in D¢ ),
To lighten the exposition, we will still
denote by Df the translate, rotate and dilate of Df. Note that Proposition (M,E) applies to it. We divide the sets Uj into two types. Type I
176
CARLOS E. KENIG
are those with diam Uj
~
TJo, and type II those for which diam Uj s_ TJo·
We first deal with the Uj of type I. In this case, Df has diameter of the order of 1. Because of the solvability of problem 3.1.2 for balls, and our normalization, we see that on a ball BCD¢, diam B A¢ (B, we have JB !Vlti 2
s_c
~
1,
JB \Adiv lti+,JVlt+vtttl\ 2 . Joining Af
to A¢ by a finite number of balls, and using interior regularity results for the system ~\i + (A+p.) 'V div lt = 0, we see that IV~(Af)l
S. C8t,
for
some absolute constant C. Then
J
s_ C a(Uj)8 2t 2 + c
Nl(A(div
lt)I+p.IV~+Vltt1) 2 da,
aof by (M,e). Thelastquantityisalsoboundedby Ca(Uj)8 2t 2 , whichis a contradiction for small 8. Now, assume that Uj is of type II. Note that in this case
th~e
exists Qj caD¢, with dist(Qj,Uj) ~ diam Uj, and such that !Vu(X)i
S.
t
for all X in the nontangential region associated to Qj. Because of this, it is easy to see, using the arguments we used to bound ivii'(Af)l in case I, that for all X in a neighborhood of Af and also on the top part of Df, we have that !Vu(X)\ m(Q) Nt
1 , if 8 is small enough, we see that we must have
Q)? m(Q).
Hence, N1 (vU -[IIU(Af); vU' -2 vir tlJ
faof Nf \'Vu
2
~
.... [Vir(A ) - virt(Aff])2 Nf Vu f 2 da ~
da ~ Ca(Uj)8 2t 2 , a contradiction if 8
is small. This finishes the proof of Proposition 3.1.13, and hence of Lemma 3.1.11. (b) The Stokes system of linear hydrostatics
In this part I will sketch the proof of the L 2 results for the Stokes system of hydrostatics. These results are joint work of E. Fabes, C. Kenig and G. Verchota ([13]). We will keep using the notation introduced in part (a).
II =
We seek a vector valued function
(u 1 , u 2 ,u 3 ) and a scalar valued
function p satisfying
(3.2.1)
~
flit = Vp in D div it= 0 in D .... .... 2 u lao ~- f ( L (00, da)
in the non-tangential sense.
THEOREM 3.2.2 (also Theorem 3.1.9). Given f ( L 2(aJ, da), there
exists a unique solution ( J,p) to (3.2 .1 ), with p tending to 0 at
oo,
and N(J)£L 2 (CD,da). Moreover, ~(X)=Kg{X), with g£L 2(aJ,da). In order to sketch the proof of 3.2.2 we introduce the matrix f'(X) of fundamental solutions (see the book of Ladyzhenskaya [25] ), f'(X) =
a..
x.x.
" lXI .
" \X\3 x.
(f 1.,.(X)), where f 1.J.(X) =-81 ~ + -81 -
.
vector q(X) = q 1 (X)), where q 1(X)
1
_J , and its corresponding pressure
1 -. = --
477\X\ 3
Our solution of (3 .2 .2) will
..
be given in the form of a double layer potential, u(X) =
fan
K.g(X) . =
IH'(Q)r(X-Q)Ig(Q)da(Q), where (H'(Q)f'(X-Q))if = oijqe(X-Q)nj{Q) +
178
CARLOS E. KENIG
ar.f act" (X-Q)nj(Q).
We will also use the single layer potential ;(X)=
J
In the same way as one establishes 3.1.4, one has: LEMMA
3.2.3. Let Kg, Sg be defined as above, with
Then, they both solve Ali= 'Vp in D, and
o-.
Also
(a)
!INII L 2 ( an,du) ~ Cll gI L 2 ( an,du) .
(b)
(Kg)±(P) = ±} g(P) - p. v.
(c)
IIN(\.lSg)\1 L 2
•
3.2.8. Let u,p be as m 3.2.5. Then,
Jl~t da ~
an
J ac
~
lvtt: 12 da +
l
J
/ns
an
~r da,
where the constants of equivalence depend only on M. Proof. 3.2.5 clearly implies, by Schwartz's inequality, that C
fan
~~~ 2da.
fan
iVI:i 2 daS:
Moreover, arguing as in the second part of the Remark 2
after 3.1.5, we see that 3.2.5 shows that
Ian
IVII 12 da S: C
Ian
IVtli 12 da
Ilac
+I
p ns hf
tx:
da
By Corollary 3.2.7, the right-hand side is bounded by
3.2.8 follows now, using 3.2.7 once more. ~
To prove 3.2.4, let u continuous across
aD.
~
=
~
S(g). By d) in 3.2.3, \ltu and ns
~s
ax.
are
J Using this fact, 3.2.3 e) and Corollary 3.2.8,
3.2.4 follows. In closing we would like to point out another boundary value problem for the Stokes system, which is of physical significance, the so-called slip boundary condition
181
ELLIPTIC BOUNDARY VALUE PROBLEMS
(3.2. 9)
This problem is very similar to (3.1.2). Using the techniques introduced in part (a), together with the observation that if ~ir = V'p, div
ir = 0
in
D, the same is true fa each row v of the matrix [\7~+ vtrt - p·l], we have obtained THEOREM 3.2.10. Given f c L 2 (dD,du) there exists a unique solution
( ~.p) to (3.2.9), which tends to 0 at ->
Moreover, u(X)
=
-+
•
-+
S(g )(X), wzth g
f
oo,
and with N(Vir) c L 2 (dD, du).
2
L (dO, du).
DEPARTMENT OF MATHEMATICS UNIVERSITY OF CHICAGO CHICAGO, ILLINOIS 60637 REFERENCES [1]
A. P. Calderon, Cauchy integrals on Lipschitz curves and related operators, Proc. Nat. Acad. Sc. U.S.A. 74 (1977), 1324-1327.
[2]
R. R. Coifman, A. Mcintosh and Y. Meyer, L'integrale de Cauchy definit un operateur borne sur L 2 pour les courbes lipschitziennes, Annals of Math. 116(1982), 361-387.
[3]
R. R. Coifman and G. Weiss, Extensions of Hardy spaces and their use in analysis, Bull. AMS 83 (1977), 569-645.
[4]
B.E.J. Dahlberg, On estimates of harmonic measure, Arch. Rational Mech. and Anal. 65 (1977), 272-288. - - - - · On the Poisson integral for Lipschitz and C 1 domains, Studia Math. 66 (1979), 13-24.
[5] [6]
B.E.J. Dahlberg and C. E. Kenig, Hardy spaces and the LP Neumann problem for Laplace's equation in a Lipschitz domain, to appear, Annals of Math.
[7]
----,Area integral estimates for higher order boundary value problems on Lipschitz domains, to appear.
[8]
D.E.J. Dahlberg, C.E. Kenig and G. C. Verchota, Boundary value problems for the systems of elastostatics on a Lipschitz domain, in preparation.
182 [9] [10]
CARLOS E. KENIG
G. David, Operateurs integraux singuliers sur certaines courbes du plan complex, Ann. Sci. del'Ecole Norm. Sup. 17 (1984), 157-189. , personal communication, 1983.
[11] E. Fabes, M. Jodeit, Jr., and Nf Riviere, Potential techniques for boundary value problems on C domains, Acta Math. 141, (1978), 165-186. [12] E. Fabes and C. E. Kenig, On the Hardy space H1 of a C 1 domain, Ark. Mat. 19 (1981), 1-22.
[13] E. Fabes, C. E. Kenig and G. C. Verchota, The Stokes system on a Lipschitz domain, in preparation. [14] C. Fefferman and E. Stein, HP spaces of several variables, Acta Math. 129 (1972), 137-193. [15] A. Gutierrez, Boundary value problems for linear elastostatics on c 1 domains, University of Minnesota preprint, 1980. [16] D. S. Jerison and C. E. Kenig, An identity with applications to harmonic measure, Bull. AMS Vol. 2 (1980), 447-451. [17]
, The Dirichlet problem in non-smooth domains, Annals of Math. 113(1981), 367-382.
[18]
, The Neumann problem on Lipschitz domains, Bull. AMS Vol. 4 (1981), 203-207
[19]
, Boundary value problems on Lipschitz domains, MAA Studies in Mathematics, Vol. 23, Studies in Partial Differential Equations, W. Littmann, editor (1982), 1-68.
[20] C. E. Kenig, Weighted HP spaces on Lipschitz domains, Amer. J. of Math. 102(1980), 129-163.
[21]
, Weighted Hardy spaces on Lipschitz domains, Proceedings of Symposia in Pure Mathematics, Vol. 35, Part 1, (1979), 263-274.
[22]
, Boundary value problems of linear elastostatics and hydrostatics on Lipschitz domains, Seminaire Goulaouic-MeyerSchwartz, 1983-84, Expose no. XXI, Ecole Polytechnique, Palaiseau, France.
[23]
, Recent progress on boundary value problems on Lipschitz domains, to appear, Proceedings of Symposia in Pure Mathematics, Proceedings of the Notre Dame Conference on Pseudodifferential Operators, Volume 43 (1985), 175-205.
[24] V. D. Kupradze, Three dimensional problems of the mathematical theory of elastocity and thermoelasticity, North Holland, New York, 1979. [25] 0. A. Ladyzhenskaya, The mathematical theory of viscous incompressible flow, Gordon and Breach, New York, 1963.
ELLIPTIC BOUNDARY VALUE PROBLEMS
183
[26]
J. Moser, On Harnack's theorem for elliptic differential equations, Comm. Pure Appl. Math. Vol. XIV (1961), 577-591.
[27]
J.
Necas, Les methodes directes en theorie des equations elliptiques, Academia, Prague,1967.
[28] L. Payne and H. Weinberger, New bounds in harmonic and biharmonic problems, J. Math. Phys. 33 (1954), 291-307. [29]
, New bounds for solutions of second order elliptic partial differential equations, Pacific J. of Math. 8 (1958), 551-573.
[30] F. Rellich, Darstellung der Eigenwerte von ~u + Au durch ein Randintegral, Math Z. 46 (1940), 635-646. [31] J. Serrin and H. Weinberger, Isolated singularities of solutions of linear elliptic equations, Amer. J. of Math. 88(1966), 258-272. [32] E. Stein and G. Weiss, On the theory of harmonic functions of several variables, I, Acta Math. 103 (1960), 25-62. [33] G. C. Verchota, Layer potentials and boundary value problems for Laplace's equation in Lipschitz domains, Thesis, University of Minnesota, (1982), also, J. of Functional Analysis, 59(1984), 572-611.
INTEGRAL FORMULAS IN COMPLEX ANALYSIS Steven G. Krantz* §1. Three basic methods for obtaining integral formulas We begin by discussing three ways to think about integral formulas on domains in C 1 , with a view to finding techniques which might generalize to en . These are (I)
Exploit symmetry of the domain;
(II)
Use differential forms and Stokes's theorem;
(III) Use functional analysis. Discussion of I. Let 11
=
lz E"C: lz I < 11. For n E" Z 1 define cpn(rei())
=
rlnleinO. Then direct calculation shows that
ciJn(O)
=
2~
J cpn(ei~d() 277
1
all n.
(1.1)
0
Now if f is harmonic on a neighborhood of 11
1
then f has an L 2
convergent Fourier expansion
(1.2) n=-oo
By linearity 1 (1.1) and (1.2) yield
*Work supported in part by the National Science Foundation. The splendid lecture notes prepared by Li Hui Ping, Li Xin Min and Ye KeYing greatly simplified the task of writing this paper. 185
186
STEVEN G. KRANTZ
J f(ei~dO. 21T
f(O) =
irr
(1.3)
0
Of course formula (1.3) holds in particular for holomorphic f; then we rewrite (1.3) as •
f(O) "' 1 .
2iTi
J~ 27T
elO -0
(iei 0d0) = 1 .
0
where y(O)
=
t
2m j
f((;) d' ,_0 y '
(1.4)
eiO, 0 ~ 0 < 277.
Now (1.4) is the Cauchy integral formula on the disc for the point z
=
0. In order to obtain the general Cauchy formula, we exploit more
symmetry. Recall that if z £A is fixed then the function
called a Mobius transformation, has the following properties: (a)
cp: A .... A
is biholomorphic (i.e. holomorphic, one-to-one, and
onto, with a holomorphic inverse). (b) cp(O) = z
(c)
cp,
¢- 1 are smooth on a neighborhood of ~.
If f is holomorphic on a neighborhood of
X,
define g(~) = f orf>(~).
Then (1.4) applied to g yields
f(z)
= g(O) = 2;i
f g(~) d~ ~ f f(rp~)) d~. =
y
Change variables by ~ =
y
cp - 1((;) = ('-z)/(1-z (;).
Then
187
INTEGRAL FORMULAS IN COMPLEX ANALYSIS
and logarithmic differentiation of ¢- 1 gives
f(z) =
2 ~i ~ f('') [:~z + 1 ~ ,:] d'
(1.5)
y
Now the numerator of the second integrand in (1.6) is a holomorphic function of '
which vanishes at 0. By (1.4), the second integral
vanishes. Formula (1.6) now becomes
which is the Cauchy Integral Formula for the disc. REMARK
1. If f is only assumed to be harmonic, then we cannot argue
that the second integral in (1.6) vanishes. Instead, a little algebra applied to (1.5) gives
This is the Poisson Integral Formula. REMARK
2. Among bounded domains in C, only the disc (and domains
biholomorphic to it) has a transitive group of biholomorphic self maps. (This follows from the Uniformization Theorem; see also [32].) Thus approach I has serious limitations in C1 . In en the limitations are even more severe. Indeed in Section 7, after we develop a lot of machinery, we shall return to the concept of symmetry in en and gain some new insights.
188
STEVEN G. KRANTZ
Discussion of II. We need some notation. Recall that in real differential
analysis on R2 we use the basis Jx ,
t
for the tangent space (i.e. all
linear first order differential operators are linear combinations of these) and dx , dy for the cotangent space. We have the pairings
< axa• dx >
=
a • dy > < ay
=
1 •
< axa. dy >
=
a• dx > < ay
=
0.
In complex analysis, it is convenient to define differential operators
The motivation for this notation is twofold. First,
Secondly, if f( z) = u(z) + iv(z) is a C 1 function, with u and v real valued, then
(au _av
ar =0 ~ ax = dY az
and
auay =_- av) ax •
(1.7)
which is the Cauchy-Riemann equations. Thus af = 0 means thett f is
az
holomorphic. We also define
dz = dx + idy , dz = dx - idy . It is immediate that
=
= 1 ,
< aa. dz > = < ~. dz > z
az
=
0.
189
INTEGRAL FORMULAS IN COMPLEX ANALYSIS
An arbitrary 1-form is written
(1.8)
u(z) = a(z)dz + b(z)dz and we define the exterior differentials
au = ~ dz az Clearly du = au
+au.
STOKES' THEOREM.
A
dz 1
au = az aa dz
A
(1.9)
dz ,
Recall If {l CC Rn is a bounded domain with smooth
boundary and u is a smooth form on fi then
In our new notation, if
n ~ C 1 ~ R2
and u is a 1-form as in (1.8), (1.9),
then Stokes' Theorem becomes
(1.10)
I(~
-;)
dz
A
dZ .
n Now we can prove THEOREM.
If {l ~ C is smoothly bounded and f is holomorphic on a
nei~hborhood of
fi then
f(z)
=~ 21Tl
If(() d(. all z ( ( -z
an
n.
190
STEVEN G. KRANTZ
Proof. Fix z
and
!(')
.,-z
t:
0. Let e 0' am> E I. If E is sufficiently small, say
Discussion of Ill. If define
{}E =
lz
l {}:
dist(z,
0 < E < Eo I then ne will also be smoothly bounded. Define
H 2(0)={f holomorphicon 0:
Jlf(()l 2 ds(() (}{}
is normal projection then (flan)
a
=1, uz j 1 az j 1
and all other pairings are 0.
j=1, .. ·,n,
194
STEVEN G. KRANTZ
If a= (a 1,···,ak),
fJ
= (f3 1 ,···,fJe) are tuples of non-negative
integers (multi-indices) then we write dza =dz
a1
A··· Adz
ak'
cfZ fJ
=azfJ 1 A···
Aazt:;!o ·
,...L
A differential form is written
u
=I aaf3 dza Acrzf3
(2.1)
a,fJ
with smooth coefficients a a{3" (If 0 ranges over
lal
=
p,
1{31
=
:S p ,q c Z and the sum in (2 .1)
q only, then u is called a form of type
(p,q) .) We then define
au =
~ aaafJ
B ~ -r:- dzj A dza A Oz' , ~ . U.t: ) a, f3 ,J
By a calculation (or functoriality), du = called holomorphic if d"u
=0.
=
~ iJaafJ
.J3 .
~ - - Qz j A dza A dz f3 oz.J a, ,j
au + au.
A C 1 function
(Note: this means that df ifi:.
is
=0,
J
j
=1, ·· · ,n,
so f is holomorphic in the one variable sense in each
variable separately.) Finally, we introduce two special forms: if w = (w 1,···,wn) is an n-tuple of smooth functions then we define the Leray form to be ~
. 1
71(w) = ~ (-1)1+ wj A dw 1 A ... A dwj_ 1 A dwj+l A··· A dwn. j=l
Likewise cu(w)
= dw 1 A · · · A d wn .
INTEGRAL FORMULAS IN COMPLEX ANALYSIS
195
We define a constant
J
W(n) =
lU(I) " lU(() .
B(O,l)
Here B(z ,r) = I( len: ,, -z I < rl. Now we may formulate a generalization of approach II in Section 1. THEOREM (The Cauchy-Fantappie formula). Let 0 ~en be a smoothly
bounded domain. Assume that w = (wl····,wn)
l
C 00(0
X
n\~)'
wj =Wj(z,(), and n
~ wj(z,() • ((j-zj) ~ 1 on
0 X 0\ ~.
(2.2)
j=l
Iff fc 1
di>
is holomorphic on n, then lor any z f0 we have
f(z)
=
nW~n)
If(() Tf(w) "6J(() .
an Before proving this result, we make some detailed remarks. REMARK 1. In case n = 1 , then w = w1 = - 1-
(-z
The Cauchy-Fantappie formula becomes
f(z) = 1 . Jf(() d( ,
2m
(-z
an which is just Cauchy's formula.
(of necessity). So
(2.3)
196
STEVEN G. KRANTZ
REMARK
2. As soon as n
~
2, the condition (2.2) no longer uniquely
determines w. However an interesting example is given by
Let us calculate what the theorem says for this w in case n = 2. Now
which by direct calculation
(2 .4)
Thus, by the theorem, we get a form of the Bochner-Martinelli formula:
for f (
c 1( n).
holomorphic on
n.
Now we turn to the proof of the Cauchy-Fantappie formula. For simplicity, we restrict attention to n = 2. Let
INTEGRAL FORMULAS IN COMPLEX ANALYSIS
If a 1 , a 2
f
1
197
then we define
B(a 1 , a 2 ) = det (a 1 , 1 the system is then "over-determined" and a compatibility condition is necessary. For n = 1 the system is not over-determined. The three basic considerations about a POE are existence, Wliqueness and regularity. It is easy to check that
a is elliptic on functions in the
interior of a given domain; hence, if u exists, it will be smooth whenever f is (we will see this in a more elementary fashion later). So interior regularity is not a problem. Also, since the kernel of
a consists of all
holomorphic functions, uniqueness is out of the question. So, for us, existence is the main issue. The following example shows that the compatibility condition does not by itself guarantee existence of u. EXAMPLE.
Let {l C C 2 be given by
n = (B(0,4)\B(0,2)) U B ((2,0), ~)
at ~ 0
202
STEVEN G. KRANTZ
v
Let U=B((l,O).o and V=B((1,0),H asshown. Let 1jfC;(u) satisfy 11
=1
on V. Finally, let
Then f is smooth and
aclosed on
holomorphic on supp (Ji,) then the function h
n n.
= u - _1__1 zl-
n\(s ((1,0), }) n lz 1 =11).
0 since
1 is well-defined and z 1 -1
If there existed a u satisfying ~ = f would be holomorphic
(dh ~ 0)
on
But u would necessarily be smooth near
(1 ,0) (since f is) hence h has a singularity at, for instance, (1 ,0). Thus we have created a function holomorphic on 8(0,4)\8(0,2) which does not continue analytically to (1 ,0). This contradicts the Haitogs extension phenomenon (an independent proof of this phenomenon will be given momentarily). o Now that we know that ~
=
f is not always solvable, let us turn to
an example where it is useful to be able to solve the
a equation.
203
INTEGRAL FORMULAS IN COMPLEX ANALYSIS
EXAMPLE. Consider the following question for an open domain 0 CC en:
If w = n n lzn = Ol -10 and if g is holomorphic { on w (in an obvious sense), can we find G
(3.1)
holomorphic on 0 such that Giw = g ?
If 0 is the unit ball, then the trivial extension G(zpz 2 ,···,zn)"' g(z 1 ,···,zn_ 1,0) will suffice. However if {} = B(0,2)\B(0,1) s; e 2 then g(z 1,0) = 1/z 1 is holomorphic on o> but could not have an extension G (else the Hartogs extension phenomenon would be contradicted). o THEOREM. Suppose that w C Cn is a connected open set such that
whenever f is a smooth a-closed (0.1) form on
ali
=
n
then the equation
f has a smooth solution. Then the answer to (3.1) is "yes."
Proof. Let 77: en-> en be given by 77(Zl····,zn) = (zl····,Zn-1'0). Let B = lz ( n: 17Z I wl. Then B, w are disjoint relatively closed subsets of
n,
tive neighborhood of w and ¢ Define
n
so there is a C 00 function ¢ on
'"" F
=0
such that ¢
=1
on a rela-
on a relative neighborhood of B.
on 0 by
'"" F(z)
{¢(z) · f(11(z))
if
0
else.
z
£
supp ¢
=
Then F gives a C 00 (but certainly not holomorphic) extension of f to 0.
204
STEVEN G. KRANTZ ~
~
We now seek a v such that F + v is holomorphic and F + vi(L) =f. With this in mind, we take v of the form zn · u and we want -~
a(F+v)=O or
Now f
is holomorphic on supp ¢ and z n is holomorphic so all that
o TT
remains is
acp . (f 017)
t-
zn .
au = 0
or
(- #).
(f 017)
(3.2)
zn The critical fact is that, by construction,
0
n {zn =01
acp
=
0 in a neighborhood of
so the right-hand side of (3.2) is smooth on {}. Also it is
easily checked to be
a closed.
Thus our hypothesis is satisfied and a ~
u satisfying (3.2) exists. Therefore F
=F +v
has all the desired
properties. o Our two examples show that solving the
a equation is (i) subtle and
(ii) useful. Thus we have ample motivation to prove our next result. LEMMA.
Let ¢
l
C~(C), k ~ 1 , and define f
u(z)
satisfies u (
c k(C)
Proof. We have
and
1 =- 2m -.
au = f.
=
cp(z)dZ. Then
lJ¢1
and let 0, flnna(P,e) is unbounded. It is useful to be able to construct singular functions. Often we ·can
nearly do this in the sense that we can find a neighborhood U cl P and a holomorphic function on
un0
which is singular at p (this is "called
a local singular function). Then the problem reduces to extending local singular functions to global ones. LEMMA.
Let 0 ~ en be a domain on which the
uniform estimates. If p (
ao
a operator satisfies
and there exists a local
si~ular
at P which is bounded off any B(P,e) then there exists a
function
~lobal
one.
211
INTEGRAL FORMULAS IN COMPLEX ANALYSIS
Outline of proof. Let g: U n 0 ... C be a local singular function at P. Let V be an open neighborhood of P such that satisfy ¢
=1
near P and ¢
a problem to find a bounded
=0
off V. Set f
V £ U. =
Let ¢ £ C 00(U)
¢ · g + u and solve a
u. Then f is a global singular function at
P. Fix a strongly pseudoconvex domain 0. We sha 11 prove later that {i) uniform estimates for the
a operator hold on
0 and {ii) local singular
functions satisfying the hypotheses of the lemma exist for each p £
ao.
By taking a suitable root of the local singular function and applying the lemma, we may construct for each P £ aG a singular function Fp at P which is in L 2{0).
Now we will prove that there is an L 2 holomorphic
function F on 0 that cannot be holomorphically continued past any boundary point. This shows that 0 is a domain of holomorphy and essentially solves the Levi problem (se17 [31 ]). For the construction of F, let !Pi lj: 1 be a countable dense set in 00. Let Hij be the L 2 holomorphic functions on 0 U B (Pi,T) , j = 1,2, ....
Let A 2 (0) be the L 2 holomorphic functions on 0. Con-
sider the restriction map Yij : Hij -> A 2 (0). Define Xij
=
image Yij
£ A2 (0).
Because FP. exists for each i, Xij /. A2 {0) for all i,j. We claim that 1
.u. Xij /.A2 (0). Assume the claim for now. Take F £A2 (0)\ .u. Xij' l,J
l,J
This is the F we seek. The claim now follows from: PROPOSITION.
Let X and Y be Banach spaces, and T: X-> Y a
continuous linear map. Then the following are equivalent: (1) T(X) is not of first category in Y .
(2) T is an open mapping. (3) T is onto.
Proof. This a variant of the Open Mapping Theorem for Banach spaces. (I am grateful toR. Huff for this proposition.) o
212
STEVEN G. KRANTZ
§5. Convexity and pseudoconvexity Let n ~ RN be an open set. Then n is called geometrically convex ifwhenever P,Q£n and
O~t~1
then (1-t)P+tQfn. Incalculus,
however, a C 2 function y = f(x) is called convex if fq? 0. How are these ideas related? If n has smooth boundary, then we may think of n as given by
n = !x
£
RN: p(x) < Ol
for a smooth function p with \lp i 0 on
an
(Exercise: use the
implicit function theorem). The function p is called a defining function
(an,
for n. If p
Tp(an)
let =
j 0 so small that
n=lz (en: dist (z, ll)
(z,()\
provided ( is near
TTZ ,
~~
1TZ,
1TZ ,
we conclude that
(\Re w(z,()\ + \Im(z,O\)
say \( -1TZ \ < r 0 • Then
which
224
STEVEN G. KRANTZ
The second integral is trivially bounded since when lz-'1 ~ r 0 then A 1 is bounded. The first is majorized by
c
J f
=C
+C
f
Now
c'r 0
0.
Therefore we may set
By a calculation (see [31 ]), the matrix (gij(z)) gives a non-{]egenerate Kahler metric on {} (called the
biholomorphic
mappin~s.
Ber~man
metric) which is invariant under
In particular it holds that if «: 0 1 .... f! 2 is
biholomorphic then
As a result, metric geodesics and curvature are preserved. The Bergman metric and kernel are potentially powerful tools in function theory, provided we can calculate them. To do so, we exploit the idea of Kerzman and Stein [231 to compare K with the Henkin kernel. However a complication arises: the Henkin integral (7.1) is a boundary
232
STEVEN G. KRANTZ
integral while the Bergman integra 1 is a solid integral. How can we compare functions with different domains? What we would like to do is apply Stokes' theorem to the Henkin integral and turn it into an integral over 0. However, for z
l
0 fixed, Henkin's kerne 1 has a singularity at (
=
z . So
Stokes' theorem does not apply. The remedy to this situation is to use an idea developed in [19], [30],
[33]: for each fixed z ( n' let
Now construct a smooth extension 'liz of 1/Jz to 0. The CauchyFantappie formula is still valid with 'Pz replacing takes place on the boundary where 'II z =
.Pz
(since the integral
ifJ z ). Thus Stokes' theorem can
be applied to the new Henkin formula containing 'II z. The resulting solid integral operator on L 2(0) can be compared with the Bergman integral via the program of Kerzman and Stein (details are in [33]). The result is that
K(z,()
=
'l'z(() +(terms which are less singular).
As a result, curvature, geodesics, etc. of the Bergman metric may be calculated. Also the dependence of these invariants on deformations of
an
can be determined (see [12], [13]; it should be noted that the methods
of [1] or [6] may be used for the deformation study instead of the KerzmanStein technique). The following are the three principal consequences of these calculations for a smoothly bounded strongly pseudoconvex (a) As Q
J
z
->
an,
n:
the Bergman metric curvature tensor at z converges
to the constant Bergman metric curvature tensor of the unit ball. The convergence is uniform over
an.
(/3) The kernel and the curvature vary smoothly with smooth perturbations of
an.
(y) 0, equipped with the Bergman metric, is a complete Riemannian
manifold.
233
INTEGRAL FORMULAS IN COMPLEX ANALYSIS Now we conclude this paper by coming full eire le and discussing
once again the topic of symmetry of domains. The reader should consider that, up to now, all of our effort has been directed at obtaining (a), (fJ), (y). Now we use those to derive concrete information about symmetries. If {} s;_ en is a domain, let Aut fl denote the group of biholomorphic self-
mappings. If two domains
nl
and
n2
are biholomorphic we will write
nl ~ n2. THEOREM (Bun Wong [41 ]). If {} CC en is smoothly bounded and
strongly pseudoconvex and if Aut fl acts transitively on !1, then
n ~ball. Proof (Klembeck). Let Pol n be any fixed point. Let IPjl Pj .... afl. By hypothesis, choose ¢j
(constant curvature tensor of the ball).
(*)
Thus the Bergman metric curvature tensor is constant on !1. We now use THEOREM (Lu Qi-Keng [34]). If M is a complete connected Kahler
manifold with the constant holomorphic sectional curvature of the ball then M ~ball. This theorem, together with (*), completes the proof. o THEOREM (Greene- Krantz [13] ). If fl
an
..\II 'S llx t: Rnl Na(u-un) >..\/211
~ [2CaAp.\- 1 llu-unllhP]P. Since p < oo, llu-u
II
n hP
... 0
as n ... "", and since ..\ > 0 is arbitrary, it
follows that
By taking a countable sequence ci a's which increase to infinity, we obtain a proof d Fatou's theorem. It is clear that the family of Euclidean balls plays an important role in this theorem, not only in the definition of nontangential approach regions, but also crucially in the definition of the Hardy-Littlewood maximal operator and the proof of its boundedness. We now recall how these
248
ALEXANDER NAGEL
balls are also involved in studying the fundamental solution for the Laplace operator. An important fundamental solution for /'t.. is given by the Newtonian potential:
N(x) = if n = 2
lrr log lxl where wn = 2rrn 12 if
cp
f
;r(T) . Then
C ~(Rn) cp(x) =
I
~N
=
8 as distributions. In particular,
N(x-y)/'t..cp(y)dy
(2)
R" and if
1/;(x) =
f
N(x-y)cp(y)dy
(3)
R" then /'t..f/;
=
cp .
Proofs of these facts can be found in Folland [5], Chapter 2. A great deal is known about the operator
f-+ N
* f(x) =
f
N(x-y)f(y)dy .
R" Basically, the fundamental idea is that, when measured with appropriate norms, N
*f
has two more orders of smoothness than f itself. For
example, if f satisfies a HCilder continuity condition of order a, 0 0 define It is easy to check that
We associate to the family of dilations a pseudometric p((x,t),(y,s)) =
~
lzjl 2 = lzl 2
Recall that n is the image of the unit ball B =
!
l(w 1 ,···,wn+l)l~tlwl+d(w,'lOJ.
What do the conesponding balls B(z, 8) look like? We see that
\V t B(Z,8) is essentially equivalent to the pair of inequalities: lz-wl < 8
Fix
z = (z,zn+l) tan.
The complex tangent space to
an
at z is
given by the equation (t)n+l- zn+l- 2i<w,z> If w = (w ,wn+l)
(an'
=
0
0
the distance from \V to this complex tangent plane
is essentially
=
\Re (wn+l-zn+l)+i(\w\ 2 + lz 12 -2Cw,z >)I
=
IRe (wn+l-zn+l-2i)+i\w-zl 2 1
~ lw-z[ 2 +(Re(wn+Czn+l-2i)l
< 82
""'
0
261
VECTOR FIELDS AND NONISOTROPIC METRICS
Thus the balls B(z, 8) are essentially "ellipsoids" of size 8 in the directions of the complex part of the tangent space to
an
at
z,
and of
size 8 2 in the orthogonal real direction, and hence in particular
Thus the doubling property of the balls is verified, and d{l (or Hn ) equipped with the pseudometric d is indeed a space of homogeneous type. We now want to discuss the analogue of Fatou's theorem for boundary behavior of holomorphic functions in n. This problem was first studied by Koranyi [9] for domains like {}, and was later generalized by Stein [17] to general smoothly bounded domains in en. Here we want to emphasize the role of the nonisotropic balls on the boundary, in analogy with the role of Euclidean balls on Rn in Fatou's theorem. We begin by defining appropriate nonisotropic approach regions in G. Let rr : {} ... aG be the projection
For a> 0 and
VI= (w,s+ilwl 2 ) l an let
Aa( VI) = l(z ,zn+t) l n:rr(z ,zn+l) l B( VI, ap(z,zn+ 1 ) 1 12 )1 where of course B(w,8) is the nonisotropic ball defined by the pseudometric d. It is clear that this definition is analogous to our earlier definition of nontangential approach regions ra<xo) in n~+l. Now (z,t+ilzl 2 +iy)
f
Aa-II~ A 1>--111fll,
THEOREM
oa
it
P = 1.
8. Suppose u is continuous on {} and pleurisubharmonic on
0. For a > 0 there is a constant Ca (independent of u ) so that for all
'(an
263
VECTOR FIELDS AND NONISOTROPIC METRICS
Theorem 7 of course follows from Theorem 5 and the Marcinkiewicz interpolation theorem (see Stein [16], Chapter 1), since we already know
an
is a space of homogeneous type. Before proving Theorem 8, we point
out some of the consequences of these results. For 1 ~ p ~ "", we let Hp(O) denote the space of holomorphic functions on
n
which satisfy
sup J\F(z,t+i\zl 2+iy)[Pdu(z,t): y>O
1\Fil~
AI=
U lNaFE>AI
eSf (see Koninyi and Vagi [10]). Finally, we consider fundamental solutions. On Hn let
a a X·=...,-+2y-"""!;", J ox, J 01:
a
J
J
where we write z · J
=X·
J
a
a T="""t:
Y-=...,--2x·~·
oy.
J
J
ot
01:
+ iy-. These vector fields form a basis for the J
left invariant vector fields on Hn. Put - = -1 (X -+iY .) - (X ·-iY.), Z. Z · = 21 J
and consider for a
t
C.
J
J
J2
J
J
267
VECTOR FIELDS AND NONISOTROPIC METRICS
This second order operator arises in the following way: if we identify Hn with
an,
the vector fields
zj
annihilate the boundary values of holo-
morphic functions. Thus, in analogy with the operator
a,
we consider
n
abf
=I zid(i,
on functions,
j=l
and we extend this in the usual way to (O,q) forms on
an.
In L 2 (Hn)
we can define a formal adjoint (~)*, and the Kohn Laplacian is then
On q forms, o{,q> acts diagonally, and is given by the operator ~a where a = n - 2q. The operator c::1tJ is not elliptic but Kohn's fundamental work [8a] showed that one can obtain subelliptic estimates for ~a. Folland and Stein [6] discovered a fundamental solution for ~a· Define:
THEOREM 10 (Folland and Stein). ~acf>a
tions, where ca is a constant, and ca
=
cao in the sense of distribu-
.J 0 if a t
±n, ±(n+2), ±(n-t-4),
etc. Thus except for the exceptional values of a, solution for ~a, and it is easy to verify that
where
o = d((O,O), (z,t)).
1 cf> is a fundamental ca a
One also obtains corresponding estimates for
derivatives of cf>a, so that again there is a complete analogy with the estimates (4) for the Newtonian potential.
268
ALEXANDER NAGEL
In the case of the Heisenberg group, the basic vector fields are xl,···,Xn, y
I•"' ,Y n
I
and the vector field T which is given "weight"
two. We shall later see how the general construction applied to these vector fields gives the nonisotropic pseudometric d. Part II. Metrics defined by vector fields Our object in this part of the paper is to outline the construction of metrics from certain families of vector fields. Many details of the arguments will be omitted, and complete proofs can be found in [13]. In [7J, Hormander studied differentiability along noncommuting vector fields, and used the techniques of exponential mappings and the Campbell-Hausdorff formula. The case of vector fields of type 2 was studied in [11]. Balls reflecting commutation properties c:i vector fields have also been studied by Fefferman and Phong [4a], by Folland and Hung [Sa], and by SanchezCalle [lSa]. Let
nc
RN be a connected open set, and l:_t y l•'",Yq be C 00
real vector fields defined on a neighborhood of 0. We associate to each vector field Yj an integer d j = d(Y ;)
? 1 which we call the formal
degree, and we make two fundamental assumptions about this collection of vector fields. -
N
(1) For each x dl, the vectors IY 1 (x),· .. ,Yq(x)l span R . (2) For all j, k, we can write [Y;,Y k] =
e
I
c~k(x)Ye
where
de:Sd/dk 1
-
cjk c C 00(0). Here [X,Y] =- XY- YX is the commutator of the two vector fields. There are several basic examples to keep in mind.
=i;
and let dj =1 for 1 ~ j 0, there are
(3) Given x and y, there
curves ¢,1/1: [0,1] .... 1} with ¢(0)=x, ¢(1)=y, t/J(O)=y, t/J(1)=z, q
q
cp'(t)=Ia/t)Y/¢(t)), t/J'(t)=Ibj(t)Yj(lj!(t)), with 1
la;(t)\
1
S (d(x,y)+e)
d(Y .) 1 ,
lb;(t)\
S d(y,z).+e)
d(Y .) 1
•
Define 8: [0,1] .... fi by ¢(at) O(t) = {
.;, (at-1) ! (3y) 2
JJ J 2rr
-y/2 0
We now need the following result, which is where the hypothesis
lirb
~
0
is used: There arc constants C 1 , C 2 which depend only on rn so that for 8 > 0,
LEMMA.
(i)
(ii)
o < c 1 ~ A(O,cS)/t(cS) ~ c 2 )
We defer the proof for a moment, and return to the estimate for u(O,iy). In our last integral, when r is between between
~
J.
and 3
o~
and
o 3J , t(r)
is
It follows from the lemma that in this range,
A(r) ~ y/5(y) 2 , and it follows from part (iii) of the lemma that the range of s integration in the integral is contained in 1\s \ < Cy I for some constant C. Thus: 277 C5(y)
iu(O,iy)I(z) is a real polynomial of degree at most m with 0
Sj
~ m, and if !'!c/J(z) 2' 0 for lzl ~
ai~ (0) az'
=
0,
8 0 then for 8 < 8 0 •
In fact, if we can prove this for 8 = 8 0 = 1, then given c/J, we apply the result to 1/J(z) = c/J(8z), and the result for general 8 follows, so it
.
suffices to study 8 = 8 0 = 1. But now we let
,I= ~c/J(z)
real polynomials of degree < m such that
iAb
--:- (0) = 0, 0 < j < m; 6.cf>(z) > 0 if lz\
az1
and
- -
I
-
I
(!'- +13c/J jazaaz13 co>
=
l
1\ .
:S 1 ;
296
ALEXANDER NAGEL
It is easy to see that }": is a compact set of polynomials and that the
maps
II 27T
c/J-+}
0
1
c/J(reiO)rdrdO
0
are continuous on I. Moreover, these functions are strictly positive
!l.c/J ? 0 and !l.c/J i 0, on jz I 'S 1 , and so the functions are bounded below by a constant Am > 0. The general result now follows by dividing since
a general polynomial
c/J by I
I
fP+f3c/J (0)!. ifzaazf3
Finally, in order to show
when for
c/J is a polynomial of degree :S m.
o = 1.
it again suffices to check this
But
is a norm on this space of polynomials. Hence lllc/JIII
:S CA(0,1) :S Ct(1).
As an easy corollary of the theorem we obtain an estimate for the Szego kernel S(z,() for the domain 0 on the diagonal z = (. Recall that the orthogonal projection S of L 2 (an) onto H2 (U) is called the Szego projection, and formally, this projection is given by integrating against a kernel: Sf(z) =
J
f(()S(z,()da((),
an
z(n.
297
VECTOR FIELDS AND NONISOTROPIC METRICS
In fact S(z,() = ~ cPj(z)c/>j(() where
lc!>jl is a complete orthonormal
basis for H2 (Q), and this series converges uniformly on compact subsets of {} x {}. (See Krantz [lOa}, Chapter l, for further details.) Now it is easy to check that for z
f {}
S(z,z) 1 12 =sup IF(z)l
..
where the supremum is taken over all F by our theorem, if F
H 2 (D) with IIFII 2 ~ 1 . But H
H 2 (D),
f
IF(z)l
f
:S CIBI- 1
J
IF(()! da(()
B
< CIBI- 1 12
I
1/2
IF(()I 2 da(()
an
where B = B(rr(z),o) is the ball centered at the projection rr(z) of z, and A(rr(z),o) = -p(z). Thus we obtain: COROLLARY.
If z
f {}
S(z,z)
:S CIB(rr(z),n)l- 1
§11. Estimates for the Szega kernel on aD As a final application, we show how one can make estimates of the Szego kernel S(z,() on the boundary, at least for certain very special domains {}. Thus, let
298
ALEXANDER NAGEL
where ¢ is a subharmonic, non-harmonic polynomial of degree m. We also make the very restrictive assumption that !:i¢(z) = !:i¢(x+iy)
is actually independent of y. Our approach is the following: if
then L is a global tangential antiholomorphic vector field, and
so we identify H2 with the kernel of the differential operator. When we identify
an
with
cX R
in the usual way' and write z
operator becomes:
We now make a change of variables on C x R ~(x,y,t) =
(x,y,t-A(x,y))
where
J lC
A(x,y) = -
~ (t,y)dt .
0
Then if we put
b'(x) =
f 0
lC
!:i¢(t)dt
~ X + iy
' this
299
VECTOR FIELDS AND NONISOTROPIC METRICS
and
"'L dXa .[dYa b ,(x) dta] +
=
+
1
it is easy to check that L(f ou(x.y.t)dydt
R2 so that
u(x,y,t) =
JJ
e 211 i(Yl7+tr)u(x,n,r)dndr.
R2
cr-1:-cr
LU=J
LJU
where
Let t/l(x, 11 ,r) = e-211(71x+b(x)T)
300
ALEXANDER NAGEL
and let h\pg(x,77,r)
=
r/I(X,'fl,r) g(x,77,r) .
Then
Now
and
are isometries. Thus L is similar to the operator
9x,
acting on func-
tions which satisfy
The kernel of this operator thus consists of functions g(77, r) so that
Let
Thensince b"'(x)?O, l=l(77,r)\r1
TJ,Tg-
I J ""'
g(r) e411(7Jr+Tb (r)) dr
0()
e4TT(71r+rb(r))dr
-0()
Thus
J
0()
P17 ,rg(x) =
-0()
g(y) K17 ,r(y)dy
302
ALEXANDER NAGEL
where
J
if (71, r)
E
l:
""
e47T(l1r-+f'b(r))dr
-00
if (71,r)/I.
0
Ff(x,y ,t)
=Iff
f(r,s,u)S((x,y,t); {r,s,u))drds dy
where
S((x,y,t); {r,s,u)) =
=
Jfe277i[(y-s)l1+(t-u)r]e2~[l1(x-r)+r(b(x}-b(r))]
J"" e -277r[(b(x)+b(r))+i(t-u)] 0
I I 00
e217l1{(x+r) + i(y-s))
00 ~------------d71
-oo
e47T(l1r-rb(r))dr
-00
This is the kernel we have to estimate. We begin by estimating the inner integral. For r > 0, set
dr.
303
VECTOR FIELDS AND NONISOTROPIC METRICS
Then replacing r by r +}, and 'T/ by 'T/ + rb'(~} it follows that F(.\+it,r) =
J
00
t :+.
-~
.
e21Ti'Tiid'T/
~•'(})- 0, (1-iA)-Y:r-V2
= ..\-'f,...-V 2 (1/.\-ir%-f; 2 ,
where we have fixed the
principal branch of z-'h---f/2 in the plane slit along the negative halfa'X!is. The pewer series expansion of (w-i)-V:r--f/2 (which holds for \w\ < 1 ), then gives the desired asymptotic expansion (1.4). Step 2. Observe next that if TJ ( C~ and f is a non-negative integer, then
I
(1.5)
""
-00
To prove this let a be a C"" function with the property that a(x) for \xi ~ 1, and a(x)
=
0 when
The first integral is dominated by
lxl
=1
~ 2, and write
cl+l.
The second integral can be
written as
with
~f
=
~ d~ ~x . A simple computation then shows that this term
is majorized by eN -
..\N
J
1X 1e-2N dX= ctN/\, -N cf-2N-l
lx\~s
if f - 2N < - 1 . Altogether then the integral in (1.5) is bounded by
314
E. M. STEIN
CNief+t +A-Nl- 2 N+lJ and we need only take
(with N >
e2t)
,
E
= .\- 1 12 ,
to get the conclusion (1.5).
A similar (but simpler) argument of integration by parts also shows that
J
ei.\x 2e(x)dx = 0(.\ -N), every N :::- 0
(1.6)
whenever
e(s' and e vanishes near the origin.
Step 3. We prove the proposition first in the case cf>(x)
=
x 2 • To do this
write
where
'J;
is a C~ function which is 1 on the support of 1/J. Now for
each N, write the taylor expansion
Substituting in the above gives three terms 00
I
(a)
-oo
-oo
:\ 2
e~x
2 .
e-x xl dx
OSCILLATORY INTEGRALS IN FOURIER ANALYSIS
315
For (a) we use (1.4); for (b) we use (1.5); and for (c) we use (1.6). It is then easy to see that their combination gives the desired asymptotic expansion for
J e j,\x 2rp(x)dx.
Let us now consider the general case when k = 2. We can then write rf>(x)
=
c(x-x 0 ) 2 + O(x-x 0) 3 with c
.J 0
and set rf>(x) = c(x-x 0) 2 [1+e(x)],
where e is a smooth function which is O(x-x 0 ), and hence Ie(x)l < 1 when x is sufficiently close to x 0 . Moreover, rf>'(x) J 0, when x J x 0 , but x lies sufficiently close to x 0 • Let us now fix such a neighborhood of x 0 , and let y = (x-x 0 ) (1 + e(x)) 1 12 • Then the mapping x .... y is a diffeomorphism of that neighborhood of x 0 to a neighborhood of y = 0, and of course cy 2 = r/>(x). Thus
with
'J;
f
C~ if the support of r/1 lies in our fixed neighborhood of x 0 .
The expansion (1.3) (for k=2 ), is then proved as a consequence of the special case treated before. REMARKS:
(1) The proof for higher k is similar and is based on the fact that
I~ eMx\-x\1 dx ~ ck,f(l - ;.\ )-(f+l) /k 0
(2) Each constant aj that appears in the asymptotic expansion (1.3) depends on only finitely many derivatives of rf> and r/1 at x 0 . Note e.g. that when k = 2, we have a 0 = Vn"(-irf>'(x 0 ))-l 12 rp(x 0). Similarly the bounds occurring in (1.3') depend on upper bounds of finitely many derivatives of rf> and r/1 in the support of r/1, the size of the support of rp, and a lower bound for rf>(k)(x 0 ).
316
E. M. STEIN
References: The reader may consult Erdelyi [8], Chapter II, where further citations of the classical literature may be found.
2. Oscillatory inteArals of the first kind, n
~
2
Only some of the above results have analogues when n ?: 2, but the extension of Proposition 1 is simple. Continuing a terminology used above we say that a phase _function
rp
defined in a neighborhood of a point x 0
in R" has x 0 as a critical point, if (\/rp)(x 0 ) ~ 0. PROPOSITION 4. Suppose "' ( C~(Rn), and ¢
is a smooth real-valued
function which has no critical points in the support of "'. Then
1(,\)
=
J
eiArp(x)"'(x)dx
= 0(,\-N),
as A-+oo
for every N
~ 0.
Rn Proof. For each x 0 in the support of "', there is a unit vector ~ and a
small ball B(x 0 ), centered at x 0 , so that (~, 'Vx)r/J(x)?: c > 0, for X (
B(xo). Decompose the integral
reiA¢(x)l{!(x)dx
as a finite sum
where each "'k is C"" and has compact support in one of these balls. It then suffices to prove the corresponding estimate for each of these integrals. Now choose a coordinate system xl'x 2 , ···,xn so that x 1 lies a long ~ . Then
But the inner integral is 0(,\ -N) by Proposition 1, and so our desired conclusion follows.
OSCILLATORY INTEGRALS IN FOURIER ANALYSIS
317
We can only state a weak analogue for the scaling principle, Proposition 2; it, however, will be useful in what follows. PROPOSITION
multi-index
5. Suppose 1/J c C~, ~ is real-valued, and for some
a, \al > 0,
throughout the support of 1/J. Then
(2.1)
with k =\a\, and the constant ck(~) is independent of A and r/1 and remains bounded as long as the ck+l norm of ~ remains bounded. Proof. Consider the real linear space of homogeneous polynomials of degree k in Rn. Let d(k,n) denote its dimension. Of course lxatla\=k is a basis for this space. However it is not difficult to see that there are d(k,n) unit vectors e- has a non-generate
PROPOSITION
critical point at
X O.
small neighborhood of
(2.2)
J
If t/J
x0 ,
f
C~ and the support of t/J is a sufficiently
then 00
eiAcf>(x)l/l(x)dx"-'.\-n;z
~aj.\-i,
as ,\ ... ..,
Rn where the asymptotics hold in the same sense as (1.3), (1.3' ).
Note. Again each of the constants aj appearing in the asymptotic expansion depends on only finite many values of derivatives of cf> and 1/J at x 0 . Thus e.g. a 0 = frrn/ 2 . \' llt, ll2' ... , lln
-~
J-1
(-illj)- 1 12) .
are the e1genvalues of the matnx
•
r/J(x 0), where
h azcf>(x o) l 12 axjaxk r
Similarly
each of the bounds occurring in the error terms depend only on upper bounds for finitely many derivatives of cf> and r/1 in the suppcrt of r/1, the size of the support of "' , and a lower bound for det
~ a;~ 1 J
v2n
D (f) j'
l af =~ ~... Ili\X j
OJ
1 , since 71 is in 3° and Vrf>(x) = O(x) as x .... 0; thus. IVxcl>(x, 71)j ~ c'> 0, if the support of ';;J is a a sufficiently small neighborhood of the origin. Hence for the 71 in region 3° we may use Proposition 4 to conclude that the left-side of (3.2) is actually O(A.-N) for every N. The proof of Theorem 1 is therefore concluded. REMARK. We have used only a special consequence of the asymptotic formula (2.2), namely the "remainder estimate" analogous to (1.3') when N = r = 0. Had we used the full formula we can get an asymptotic expansion for d~({); its main term is explicitly expressible in terms of the Gaussian curvature at those points x direction
t
or
t
S , for which the normal is in the
-t.
(2) We shall now consider the problem in a wider setting. Here S will be a smooth m-dimensional sub-manifold, with 1
~
m S_ n-1, and our
assumptions on the non-vanishing curvature will be replaced by the more
324
E. M. STEIN
general assumption that at each point S has at most a finite order contact with any hyperplane. We shall call such sub-manifolds of finite type. (These have some analogy with the finite-type domains in several complex variables, which are also discussed in Nagel's lectures [21 ].) The precise definitions required for our considerations are as follows. We shall assume that we are considering S in a sufficiently small neighborhood of a given point, and th"en write S as the image of mapping
cp : Rm
... Rn,
defined in a neighborhood U of the origin in Rn. (To get a smoothly embedded
s
we should also suppose that· the vectors
aacp , Xl
c:!2 ,... ,~
UAm
UA
are linearly independent for each x, but we shall not need that assumption.) Now fix any point x 0 ( U C Rm, and any unit vector 71 in Rn. We shall assume that the function (cp(x)- cp(x 0 )) • 71 does not vanish of infinite order as x .... x 0 • Put another way, for each x 0 unit vector 11, there is a multi-index a, with 1 'S
(Jxt
f
U and each
\al , so that
(cp(x)· 1/)lx=xo f. 0. Notice that if (x',.,') are sufficiently close to
(x 0 ,71), thenalso
c~r¢(x')·71'1x=x'JO.
eachunitvector TJ then 3a, be called the type of the type of
cp
cp
lal~k,
with
Thesmallest k sothatfor
~(cp(x)·11)l 0 =10 axa x=x
will
at x 0 • Also if U1 is a compact set in U,
in ul will be the least upper bound of the types for x 0
in ul. THEOREM
2. Suppose
finite type.
Let dlL
(3.3)
S is a smooth m-dimensional manifold in
= c/fdo, \d~(e:)\
and in fact we can take e =
with c/f ( C;;'(Rm).
'S A\e:l-e,
lor some
Rn of
Then
e> 0,
1/k, where k is the type of S inside the
support of c/f. Proof. By a suitable partition of unity we can reduce the problem to
showing that
OSCILLATORY INTEGRALS IN FOURIER ANALYSIS
I
ei¢(x)·e-~(x)dx
=
325
O(ie-1-1 /k)
Rm with ¢ as described above, and the support of Now we can write there is an
e- = ATJ,
a with 1
Ia! :S
with
!TJI
=
k 1 so that
'J;
sufficiently small.
1 , and >.. > 0. Then we know that
(Jx)a ¢(x) ·
Tf
I 0 whenever x is 1
in the support of ~ (once the size of the support has been chosen small enough). Thus the conclusion (3.3) follows from (2.1) of PropositionS. References.
Theorem 1 in its more precise form alluded to in the remark
goes back to Hlawka [14]. See also Herz [13], Littman [18]. Randol [25]. and Hormander [16]. When S is a real-analytic sub-manifold not contained in any affine hyper-plane, then it is of finite type as defined above. For such real-analytic S estimates of the type (3.3) were proved by Bjork [2].
4. Restriction theorems for the Fourier transform The Fourier transform of a function in LP(Rn), 1 < p :S 2 is most naturally thought of as an Lp' function (via the Hausdorff-Young Theorem) and so at first sight it is viewed as defined only almost-everywhere. This impression is further supported by the case p ""' 2 , when clearly the Fourier transform can be completely arbitrary on any given set of zero Lebesgue measure. It is therefore a noteworthy fact that whenever n 2: 2 and S is a sub-manifold of Rn (with some appropriate "curvature") then there exists a p 0 = p(S), p 0 > 1, so that every function in LP, 1 -::; p-::; p 0 has a Fourier transform restricting to S (i.e. with respect to
the induced measure on S ). Let us make this precise. Suppose that S is a given smooth sub-manifold in Rn, with da its induced· Lebesgue measure. We shall say that the LP restriction property holds for S, if there exists a q
=
q(p), so that the inequality
326
E. M. STEIN
(4.1)
holds for each f
S,
f
whenever S 0 is an open subset of S with compact
closure in S. THEOREM
3. Suppose S is a smooth hypersurface in Rn with non-zero
Gaussian curvature. Then the restriction property (4.1) holds for
1 < p < 2n + 2 , (with q - - n+3
= 2 ).
.
Proof. Suppose 1/1 ~ 0 and 1/1 ( C~. It will suffice to prove the
inequality (4.2)
for p 0 = 2 n + 32, and f n+
l
S;
the case 1 < p < p0 will then follow by -
interpolation.* By covering the support of
-
1/J
by sufficiently many small
open sets, it will be enough to prove (4.2) when (after a suitable rotation and translation of coordinates) the surface S can be represented (in the support of
1/J) as
a graph: ~n = 4>(~l'····~n- 1 ). Now with d~t = 1/fda we
have that
where (Tf )(x)
= (h K)(x),
with
*1n fact the interpolation argument shows that we can take q so that {4.1) holds with q = (n- 1) p', which is the optimal relation between p and q.
n+1
327
OSCILLATORY INTEGRALS IN FOURIER ANALYSIS
Thus (4.2) follows from Holder's inequality if we can show that (4.3) where
p~
is the dual exponent to p 0 .
To prove (4.3) we consider the function Ks (initially defined for Re(s) > 0) by
es2_ (4.4) K x _ _ s< ) - f'(s/2)
f Rn
Here we have abbreviated (~ 1 , .. ·,~n- 1 ) by ~'; we have set ';/;(~') =
rp(~')(1 + I'V¢(~')! 2 ) 1 12 ,
so that
VJ(~')d~' = d~;
also 71 is a
C~(R)
function which equals 1 near the origin. Now the change of variables ~n ... ~n + ¢(~') in the above integral shows that it equals
with
-00
Now it is well known that ~s has an analytic continuation in s which is an entire function; also ~0
= 1; and i(s<xn)l :S. clxni-Re(s),
where lxnl ? 1 , and the real part of s remains bounded. From these facts it follows that Ks has an analytic continuation to an entire function s (whose values are smooth functions of x 1' ... , xn of at most polynomial growth). One can conclude as well that
328
E. M. STEIN
(a) K 0 (x)
=
K(x) ,
(b) IK_n/2+it(x)l ~A, all x
l R0 , all real t , n (c) IKl+it(t")l ~A, all ~ l R , all real t . ~
In fact (c) is immediate from our initial definition (4.4), and (b) follows from Theorem 1. Now consider the analytic family Ts of operators defined by Ts(f) =
h Ks. From (b) one has (4.5)
I!T_n/ 2 +it(f)ll
L
00
~ Al!fll 1 , all real t , L
and from (c) and Plancherel's theorem one gets (4.6)
liT l+it 0, and the metric could be defined in terms of the usual distance. The second are the dilations (z,t) ... (pz,p 2t), and the appropriate quasi-distance (from the origin) is then (lz14 + t2) 1 14 . The latter dilations and metric are closely tied with the realization of the Heisenberg group as the boundary of the generalized upper half-space holomorphically equivalent with the unit ball in en+l. This point of view, as well as related generalizations, is elaborated in Nagel's lectures [21].
336
E. M. STEIN
In the present context the first type of dilations and corresponding metric would be appropriate if one considered expressions related to ordinary potential theory in Hm viewed as R2 m+ 1 • However the two conflicting types of dilations (and related metrics) occur in e.g. the solutions of du.
=
f. (One sees this for example in Krantz's lectures [17],
where in the formula of Henkin we have a kernel made of prQducts of functions each belo~ging to one of the two above homogeneities.)* Other expressions of this type occur in the explicit formulae for the solutions of the a-Neumann problem (see [1], Chapter 7). Let us now consider the simplest operator on the Heisenberg group displaying simultaneously these two homogeneities. The prime example is given by Tf =hK
(6.1)
where convolution is with respect to the Heisenberg group, and the kernel K is a distribution of the form (6.2)
K(z,t)
=
L(z)B(t).
L(z) is a standard Calderon-Zygmund kernel in em
=
R 2 m, i.e. L(pz) =
p- 2 m L(z), L is smooth away from the origin, and L has vanishing
mean-value on the unit sphere. Here B(t) is the Dirac delta function in the t-variable, and in an obvious sense is homogeneous B(pt) = p - 1 B(t). Thus K is homogeneous at degree -2m- 1 with respect to the standard dilations, and at the same time homogeneous of degree -2m - 2 with respect to the other dilations; in both instances the degrees are the critical ones. We turn next to the question of proving that the operator (6.1) is bounded on L 2 (Hm). The most efficient way is to proceed via the Fourier transform in the t-variable. This leads to the problem of showing that the family of operators T,\ defined by
*In particular the terms A 1 and A 2 that appear in §6 of [17].
337
OSCILLATORY INTEGRALS IN FOURIER ANALYSIS
(TA,)(F)(z) =
(6.3)
J
L(z-w)eiAF(w)dw ,
em (with < z ,w > the anti-symmetric form which occurs in the multiplication law for the Heisenberg group) is bounded on L 2 (Cm) to itself, uniformly in A, -oo 0. Con-
We fix a curve t .... y(t) = (t, y 2 (t)), 0 y(t)
£
c2 ,
sider the transformation T, which maps function on the interval [0,1] to functions on R 2 , given by
Je~·y(t)f(t)dt. 1
(7 .1)
(T£)(0 =
0
THEOREM 9. Under the assumption above T is bounded from LP[0,1] to L q(R 2 ), whenever 3/q + 1/p ~ 1 and 1 :S p
< 4.
(Note that when p .... 4, then q .... 4 in the above relation between p and q .)
345
OSCILLATORY INTEGRALS IN FOURIER ANALYSIS
Proof. Write
F(~) = ((Tf)(~)) 2 =
(7.2)
f Je~·(y(s)+y(t))f(s)f(t)ds 1
0
1
dt ,
0
and we shall try to apply Plancherel 's theorem (more precisely, the Hausdorff-Young inequality) to F. To do this break the above integral into two essentially equal parts according to t
>s
or t 'S_ s , which
divides [0,1} x [0,1] into the union of two regions R 1 and R2 • We then consider the mapping of R 1 .... R2 given by x = y(s) + y(t), i.e. x 1 = s + t, x 2 -= Y/s) + y 2(t). It is easy to verify on the basis of our assumptions that this rna pping is one-one, and its Jacobian J satisfies IJ I -= \y;(s)-y;(t)\ ~ c\s-t\. Therefore
(7.3)
I
e ~·(y(s)+y(t))f(s)f(t)dsdt =
f e~·xf(x
x )dx dx
1' 2
1
2
R2
R1
with f(x 1 ,x 2 ) = f(s)f(t)\J\- 1 • So if we denote by F1 (~) the quantity appearing in (7 .3) then whenever 1 'S_ r (7.4)
:S
2, and 1/r' + 1/r = 1 , we know that
IIFlll Lr' (R2) :S cll£11 Lr (R2) ·
However
=
Jlf(s)\r lf(t)ifiJ\ 1 -r dsdt
346
E. M. STEIN
To estimate the last integral we need to invoke the theorem of fractional integration in one dimension in the form
llf\1~ whe~ p
So we take g(t)"' \f(tW, then \\gllu
=
ur. Then if we fix
a so that -1 +a= 1-r, then 32r. =
~, and p = 3 ~r. The limitation
=
0 < a becomes r < 2, and with q == 2r' we obtain from (7.4) that
IIF1 11 r' 2 L
(R )
:S c'
(J1
)2/p
\f(t)\P dt
,
0
with a similar estimate for F2 (0 which is the analogue of (7.4), but taken over R 2 . Since F = F1 + F2 and F = (Tf ) 2 we obtain IIT(f >II q
2
:s A1\fll L p[0,1] .
,
so
L (R )
Note that ~ = -1. = 3r- 3 = 1 - !. q 2r' 2r p 1
J. ~ !.p = 1 ,
q
and the limitation
:S r < 2 is equivalent with 1 :S p = 32~ r < 4. Theorem 9 is therefore
proved. It is clear that inequalities for the Fourier transform play a key role
in the above argument. If we want to generalize Theorem 9 it is natural to look for a corresponding extension of the L 2 boundedness of the Fourier transform and the Hausdorff-Young theorem. One result along these lines is as follows. Suppose we consider the family of operators T, depending on the parameter .\ , .\ > 0, defined by
T>,(f)(()
=
J Rn
ei.\tf)(x,O!/I(x,()f(x)dx,
OSCILLATORY INTEGRALS IN FOURIER ANALYSIS
347
where 1/J is a fixed C~(Rn x Rn) cut-off function; ell is a real-valued C"" phase function which we assume satisfies the assumption that its Hessian is non-vanishing, i.e.
(7.6)
PROPOSITION:
(7.7) COROLLARY: (7.8) where 1 ~ p ~ 2, and 1/p + 1/p' = 1.
REMARK. The boundedness of TA for any fixed A is trivial, but what is of interest is the decrease in the norm as A _. "". This decrease is consistent with the special case when Cl>(x, () is bilinear (and nondegenerate); when we take A ->
oc
in that case we recover the usual
(LP,LP') inequalities for the Fourier transform. Notice also that the corollary follows from the proposition by the use of theM. Riesz convexity theorem. To prove (7. 7) we argue as in the proof of Theorem 7; as in the treatment of the operator T"" it suffices to show that the operator norm of T~TA is bounded by AA-n. Now this operator has as its kernel the
function KA ((,n) given by
(7.9)
KA((,71)
=
J Rn
eiA(CI>{x,71}-CI>(x.())¢(x,J7)1/1(x,()dx .
348
E. M. STEIN
Now since
we can find a= (a 1 , .. ·,an), so that the aj depend smoothly on x and
l~(x,g,fl) ~ cle-111 .oo the support of K~(g,fl)· Set Dx = i~ (a,Vx>· Then since (Dx)~ei.\(«(x,fl)-«(xl)) = eiA(«(x,fl)-«(xl)), we can integrate by parts N times in (7.9) and obtain (7.10) It follows from (7.10) with N = n + 1, that the operator T~T,\ which has kernel K,\ has a norm bounded by A,\-n and the proposition is proved. We shall now formulate some theorems for oscillatory integrals of the form (7.11)
(T,\f)(0=J
eiA«(tl)l/l(t,g)f(t)dt, g(Rn,
Rn-1
which will generalize the restriction theorems (Theorem 9 above, as well as Theorem 3 in §4) and also give results for Bochner-Riesz summability. Notice that (7.11) are mappings from functions on Rn- 1 to functions on Rn. The basic assumptions on the real phase function «!> are as follows: We consider for each fixed (t 0 ,g ~ the associated bilinear form B(u,v) definedby B(u,v)=(v,Vt)(u,Vg)(«)(t 0 l~. with U£Rn- 1 , v
f
Rn. Our first assumption is that B is of rank n - 1 .
Thus there exists (an essentially unique), ii
f
Rn,
lui
= 1, so that the
scalar function t .... (U,Vg«(t,g}) has a critical point at (t 0 ,g 0 ). We shall also assume that this critical point is non-degenerate, i.e. we suppose the non-vanishing of the (n-1)x(n-1) determinant:
OSCILLATORY INTEGRALS IN FOURIER ANALYSIS
349
These assumptions will be supposed to hold at all (t 0 ,g~ in the support of r/J(t, 0, where r/1 is a fixed function in c;(Rn- 1 x Rn). THEOREM
10. Under the assumptions above the operator (7.11) satisfies
(7 .13) with
q=(~:Dp', ~+f=1,
(a) when n = 2, if 1
(b) when n
~
) (0) = 0, then the conditions (7.12) are near the origin equivalent with the non-vanishing Gaussian curvature of the graph tn = ¢(t 1 , .. ·,tn_ 1 ). If we apply the result (7 .13), letting A ...
oo,
it is not difficult to recover
Theorem 9 from part (a), and Theorem 3 from part (b). (2) The proof of part (a) follows the same lines as the proof given for Theorem 9, once we use (7.8) as the substitute for the Hausdorff-Young theorem; further details as well as relations with Bochner-Riesz summability may be found in the papers of Carleson and Sjolin [3] and Ht>rmander [15]. Since part (b) has not appeared before, we will outline its proof. This will also serve as a good review of many c:i. the notions we have discussed here. Proof of part (b). It suffices to prove the case p = 2, since the case
p = 1 is trivial and the rest follows by interpolation. Now the case p = 2 is equivalent by duality to the statement
(7.14)
IIT~(F)II
2 Rn-1
L (
)
~ AA-n/r'IIFII r Rn L (
)
350
E. M. STEIN
with r
o:
2 0 (0 so that (u,9~)cl> 0(~) i 0 to increase the rank of 9x 9~~ to n. Now, as in the proof of Theorem 3 in §4, we form K~ defined by 2
Kf(~,q) = ~~:/2)
f eiA(a>(x,77}-a>(x,~)).p(t,q)rf;(t,01xnrl+SJJ(xn)dx, R"
with dx
~
dt dxn, and where v is a C'(;' function which equals 1 near
the origin. We easily verify (7.16) since a>(x, 0 = cl>(t, 0, when x = (t,O) . Next (7.17)
K~+it is the kPrnel L 2 (Rn)
of a bounded operator from
to itself with norm ~ M-n/2.
This follows by applying the estimate (7.7) of the proposition above and using the non-degeneracy of the Hessian of iP(x, ~). Finally we claim that (7 .18)
00
f
-oo
352
E. M. STEIN
Then since lil_n/ 2 + 1 / 2 +it(u)\
'5. cluln/2-l/2 ,
as u
"" we see that to
prove (7.18) it suffices to show that (7.19)
1n proving this estimate for the integral KA given by (7.15) we may suppose that the integrand is supported in a sufficiently small neighborhood of a given point t = t 0 , (for otherwise we can write it as the sum of finitely many such terms). When we write ll>(t,71)
=
ll>(t,O
= (\7~ll>)(t,71)
·
(71-~) + 0(71-~) 2 we see that these are two cases to consider as in the
proof of Theorem 1 in §3: 1° when the directions 71 - ~ or ~ - Tf are close to the critical direction u arising in condition (7.12b); or 2° in the opposite case. In the first case we use stationary phase (i.e. Proposition 6 in §2) to obtain (7.19). 1n the second case, we actually get O(A-177-~ 1)-N, for every N ~ 0 as an estimate, by Proposition 4. This
completes the proof of (7.18), and shows that K~n/2+l 12+it is the kernel of a bounded operator from L 1 (Rn) to L 00(Rn), with bounds uniform in A. The proof of the theorem is then concluded by applying the interpolation theorem, as in the proof of Theorem 3. 8. Appendix Here we shall prove Lemma 2 and Theorem 6 which were stated in §S. First let Pd denote the linear space of polynomials in Rn of degree 0, !log u! (8.3) implies (5.4) whenever mp
=
=
log+u + log+fr ~ u + }u-r. Therefore
1 , and so that result also holds in
general. E. M. STEIN DEPARTMENT OF MATHEMATICS PRINCETON UNIVERSITY PRINCETON, NEW JERSEY 08544
REFERENCES (1]
M. Beals, C. Fefferman, and R. Grossman, "Strictly pseudo-convex domains," Bull. A.M.S. 8(1983), 125-322.
(2]
J. E. Bjorck, "On Fourier transforms of smooth measures carried by real-analytic submanifolds of Rn," preprint 1973.
354
E. M. STEIN
[3]
L. Carleson and P. Sjolin, "Oscillatory integrals and a multiplier problem for the disc," Studia Math. 44(1972), 287-299.
[4]
M. Christ, "On the restriction of the Fourier transform to curves," Trans. Amer. Math. Soc., 287(1985), 223-238.
[5]
R. Coifman andY. Meyer, in these proceedings.
[6]
R. Coifman and C. Fefferman, "Weighted norm inequalities for maximal functions and singular integrals," Studia Math. 51 (1979), 241-250.
[7]
S. Drury, "Restrictions of Fourier transforms to curves," preprint.
[8]
A. Erdelyi, "Asymptotics Expansions," 1956, Dover Publication.
[9]
C. Fefferman, "Inequalities for strongly singular convolution operators," Acta Math. 124 (1970), 9-36.
[10] C. Fefferman and E. M. Stein, "HP spaces of several variables," Acta Math. 129 (1972), 137-193.
[11] D. Geller and E. M. Stein, "Estimates for sin gular convolution operators on the Heisenberg group," Math. Ann. 267 (1984), 1-15. [12] A. Greenleaf, "Principal curvature in harmonic analysis," Ind. Univer. Math. J. 30 (1981 ), 519-537. [13] C. S. Herz, "Fourier transforms related to convex sets," Ann. of Math. 75 (1962), 81-92. [14] E. Hlawka, "Uber Integrate auf konvexen Korper. 1," Monatsh. Math. 54 (1950), 1-36. [15] L. Hormander, "Oscillatory integrals and multipliers on FLP," Ark. Mat. 11 (1973), 1-11. [16]
, "The analysis of linear partial differential operators. I," 1983, Springer Verlag.
[17] S. Krantz, "Integral formulas in complex analysis," in these proceedings. [18] W. Littman, "Fourier transforms of surface-carried measures and differentiability of surface averages," Bull. A.M.S. 69(1963), 766-770. [19] G. Mauceri, M.A. Picardello, and F. Ricci, "Twisted convolutions, Hardy spaces, and Hormander multipliers," Supp. Rend. Cir. MatPalermo 1 (1981), 191-202. [20]
J. Milnor, "Morse Theory," Annals of Math. Study
tt51, 1963,
Princeton University Press. [21] A. Nagel, "Vector fields and nonisotropic metrics," in these proceedings.
OSCILLATORY INTI!:GRALS IN FOURIER ANALYSIS
355
[22] D. H. Phong and E. M. Stein, "Singular integrals related to the Radon transform and boundary value problems," Proc. Nat. Acad. Sci. USA 80(1983), 7697-7701. [23]
, "Hilbert integrals, singular integrals, and Radon transforms,'' preprint.
[24] E. Prestini, "Restriction theorems for the Fourier transform to some manifolds in Rn in Harmonic analysis in Euclidean spaces," Proc. Symp. in Pure Math. 35, part 1(1979), 101-109. [25] B. Randol, "On the asymptotic behaviour of the Fourier transform of the indicator function ci. a convex set," Trans. Amer. Math. Soc. 139 (1969), 279-285. [26] E. M. Stein, "Singular integrals and differentiability properties of functions," 1970, Princeton University Press. (27] E. M. Stein and S. Wainger, "The estimation of an integral arising in multiplier transformations," Studia Math. 35 (1970), 101-104. [28] E. M. Stein and G. Weiss, "Introduction to Fourier analysis on Euclidean spaces," 1971, Princeton University Press. [291 R. S. Strichartz, "Restrictions of Fourier transforms to quadratic surfaces and decay of solutions of wave equations/' Duke Math. ] .
44(1977), 705-713. [30] P. A. Tomas, A restriction theorem for the Fourier transform," Bull. A.M.S. 81 (1975), 477-478. [31]
, "Restriction theorems for the Fourier transform in Harmonic Analysis in Euclidean spaces," Proc. Symp. in Pure Math. 35, part 1 (1979), 111-114.
[32] S. Wainger, "Averages and singular integrals over lower dimensional sets," in these proceedings.
[331 A. Zygmund, "On Fourier coefficients and transforms of functions of two variables," Studia Math. 50(1974), 189-201.
AVERAGES AND SINGULAR INTEGRALS OVER LOWER DIMENSIONAL SETS Stephen Wainger< 1 ) I.
Introduction These lectures deal with work primarily due to Alex Nagel, Nestor
Riviere, Eli Stein, and myself dealing with certain averages of and singular integral operators on functions, f, of n variables, n
~
2. These
averages and singular integrals differ in character from the classical theory in that the integration is over a manifold of dimension less than n.
Let us begin with an example of the type of problem we have in mind. The classical differentiation theorem of Lebesgue asserts for any locally integrable function f
f(x) = lim - 1r-oO IQrl
J
f(x-y)dy
a.e.
Q,
(where Qr is the square, Qr = !xt:Rn!sup!xil ~rl, and !Qrl denotes the Lebesgue measure of Qr ), and
f(x) = lim - 1r-oO IBrl
J
f(x-y)dt
Br
(where Br is the ball, Br = lxllxl <Sri).
* Supported in part by a grant from the National Science Foundation.
357
a.e.,
358
STEPHEN WAINGER
Our first problems are the following: Problem lA: Does lim
1)
r ... o
J
f(x-y)dor(y) == f(x)
a.e.?
aQr Here aQr denotes the boundary of Qr
~nd
dor is n-1 dimensional
Lebesque measure on aQr normalized so that dor(aQr) == 1 . Problem IB: Does 2)
lim r-+0
J
f(x-y)dttr = f(x)
a.e.?
asr
Here dllr is the unit rotationally invariant mass on aBr. 1) and 2) trivially hold if f is continuous, and the questions only become interesting when we consider functions in a class like L"", L 2 or
,
L1 .
In questions lA and IB, we are considering certain averages
3)
MQ/(x)
=I
f(x-y)dor(Y)
aQr
and
4)
Ma/(x)
=I
f(x-y)dllr(Y).
aar We are asking if 5)
MQ f(x) ... f(x) r
a.e.
AVERAGES AND SINGULAR INTEGRALS
359
and M8 f(x)
,6)
->
f(x)
a.e.
r
The standard approach to this type of problem involves considemtion of appropriate maximal functions. We define the maximal functions 7)
and
m8 f(x) = sup M8
8)
r>O
mB
(Jfj){x) . r
is called the spherical maximal function. Since 1) and 2) hold for f which are continuous 1) would follow for
every f in LP, 1
'S p < oo, if we could show
for every f in Lp, and 2) would follow if we could show
The argument showing that 9) and 10) imply 1) and 2) is the same as the argument showing that Lebesgue's differentiation theorem follows from the weak type inequality for the Hardy-Littlewood maximal function given in chapter 1 of [S]. While it is not quite as well known, there are appropriate estimates on maximal functions that guarantee 1) and 2) hold for all L 00 functions. In our case this means the following: Let E be a measurable set and XE its characteristic function. Then if (11)
360
STEPHEN WAINGER
where C(A) may depend on A but not on E, then 1) holds for every f in L"". H
12) then 2) holds for every f in L ""·. A discussion of this can be found in [BF]. Let us try to see if 9) or 10) could be true in some simple cases. We consider for example the one-dimensional case. Here Br = Qr = lxl-r<xQf(x) =
:Jil 8 f(x) = oo for every x. We can
also see that 11) and 12) are false in one dimension. We just take Ee = lxl~x::;el. Then !Eel .... 0 but
for all x. We could still ask if 1) and 2) hold in some interesting class even though 9), 19), 11 ), and 12) fail. However an important idea of Stein shows that the failure [Sl] of 9), 1 0), 11 ), and 12) implies that 1) and 2) fail even in the class of locally bounded functions. The statement of the ·main theorem of [SI] requires that the underlying space be compact. But if 1) or 2) were true for an LP class on Rn, it would also hold for the
corresponding LP class on the torus. Furthermore, the theorem of Stein requires the hypothesis that 1
:S p :S 2. However due to the positive
nature of the averages under consideration, his ideas can be modified to show that 1) fails for at least some L 00 functions. See [SW]. Thus we obtain negative results in one dimension. Similar reasoning gives the same negative conclusion for question lA in any number of dimensions.
AVERAGES AND SINGULAR INTEGRALS
361
One need only consider F(x 1 ,. .. ,xn) = f(x 1 )h(x 2 ,··· ,xn) where f is as above and h is a nice function. So there are no interesting positive results in problem lA. However as we shall see later there are positive results for problem 18 in 3 or more dimensions. One might ask if there is a simple geometric reason why . there should be positive answers for the sphere and only negative answers \
for the boundaries of squares. It turns out that the underlying basic reason that we have positive results for the boundary of balls and negative results for the boundaiy of squares is that spheres are round and boundaries of squares are flat. In other words an important word for us will be CURVATURE. We will come back to the role d curvature in our problem in a little while, but first we shall discuss the other problems that we will consider. Problem II: Let y(t) be a curve passing through the origin in Rn. Is it true that lim
kJ
0h
f(x-y(t))dt = f(x) a.e., for f in L.,., or L2 or L 1 ?
Problem III: Let v(x) be a smooth vector field in Rn. Does
J h
lim _hl h->0
f(x-tv(x))dt = f(x)
a.e.
0
for f in L""' or L2 or L1 ? Corresponding to problems II and III there are interesting singular integrals. We let y(t) be a curve and v(x) be a smooth vector field as in problems II and III. We set
13)
Hi(x) =
I
a
f(x-y(t)) dtt '
-a
(where sometimes we wish to think of a as finite and sometimes as ""' ), and
STEPHEN WAINGER
362
J 1
Hvf(x) =
14)
f(x-tv(x)) d( .
-1
We call Hy the Hilbert transform along the curve y and Hv the Hilbert transform along the vector field v(x). We then have the following two problems: Problem II': Can we have an estimate 15)
for some p's ? Problem III': Can we have an estimate 16)
for some values of p ? The classical development of singular integrals and maximal functions suggests that problems II' and III' should be related to problems II and III. In fact the progress on problems I, II, III, II', and III' is all interrelated. We have presented our problems as variants of Lebesgue's Theorem on the differentiation of the integral. These particular variants arose from other considerations. Riviere was led to problem II' from the consideration of a problem of singular integrals, namely from trying to generalize the method of Rotations of Calderon and Zygmund. Calderon and Zygmund developed the method of rotations to reduce the study of operators
where K is a kernel having "standard homogeneity" that is 17)
A>O
AVERAGES AND SINGULAR INTEGRALS
363
(K(x) is a function on Rn ) to the one-dimensional Hilbert transform
Hf(x) =Jf(x-t) dtt
( f a function on R 1 ). We will explain how the method of rotations can lead to problem ll'. Let K(x,y) be a function of two variables x and y which is odd, K(-x,-y) = -K(x,y) and which has a "parabolic homogeneity," that is 18)
We wish to consider the LP boundedness of the transformation
JJ 00
Tf(u,v) =
19)
00
-DO
f(u-x,v-y) K(x,y)dx dy .
-00
We now introduce parabolic polar coordinates into 19) x = rcos ()
and find
JJ ""
20)
Tf(u,v)
=
0
277 f(u-rcos (), v-r 2sin ())
0
where r 2 N(O) is the jacobian factor in the change of variables. N(O) is smooth and N(0+77) = N(O). By 18 we see that
364
STEPHEN WAINGER
J
211
21) Tf(u,v) =
I
DO
N(O)K(cosO,sinO)dO
0
}f(u-rcosO,v-rsinO)dr
0
-I
=
I} ~
217
N(O)K(cos(0+TT),sin(O+rr))d0
0
f(u-rcos O,v-rsin(O))dr
0
since K is odd. Thus
J
217
Tf(u,v) =-
J ~
N(O)K(cosO,sinO)dO
0
}f(u+rcosO,v+rsinO)dr
0
since N(O+rr) = N(O) . Finally
J
211
21A) Tf(u,v) =
J} DO
N(O)K(cost,sin O)dO
0
f(u-rcos O;v-r 2 sin O)dr .
-oa
Now adding 21) and 21A) we find that 211
Tf(u,v)
=}I
J
DO
N(O)K(cosO,sinO)dO
0
}f(u-rcos0,v-r2 sin0)dr.
-~
If K(cos O,sin 0) is in L 1 of [0,2rr], we can apply Minkowski's inequality 1T
IITfiiLP
:scI 0
where
dO
IIHofiiLp
AVERAGES AND SINGULAR INTEGRALS
Hef =
I
365
00
f(x-yo(r)) d: ,
-00
with
Now we prove
by showing
This is a problem of the type II'. Stein was led to consider problem II by his study of Poisson integrals on symmetric spaces. We are not going to launch into a discussion of symmetric spaces, but instead we consider an example. Let
If K(x,y) were dominated by a decreasing, radial, L 1 function, the
classical theory would imply lim Me f(x,y) = f(x,y) e~o
see [SWE]. However
a.e.
366
STEPHEN WAINGER
so the smallest ~adial majorant of K is
1 which is not integrable. l+X2+y2
In effect K has too much of its mass along the coordinate axis. The extreme case of this phenomena would be to have a kernel with all of its mass on the coordinate axis. In other examples, kernels have too much of their mass along curves, and the extreme case of difficulties arising in problems of Poisson Integrals on symmetric spaces lead to Problem II. Appropriate positive results to problem III would have implications for the boundary behavior of functions holomorphic in pseudoconvex domains in en. The natural balls in these problems are long, thin and twisting. The idealized situation is that of a vector field. In the case of a strictly pseudo convex domain, the balls satisfy the standard properties that ensure that the usual covering arguments apply. See [SBC). For progress in the case of pseudoconvex domains see [NSW) and [NSWB). Now that we have seen some of the roots of our problems, let us consider why these problems don't fit into the framework of the standard theory of Maximal functions and singular integrals as presented for example in [S]. In the standard treatment of averages over Euclidean balls an important geometric property of the Euclidean balls is used. If two balls B 1 and B 2 of the same radius, r, intersect, then B;, the ball having the same center as B 2 but having radius 3r contains B 1 . To see how badly this property fails for our problems let us suppose we were considering averages 1
he
Jh Jt2 +E
0
f(x-r,y-s)drds
t2
over slightly thickened parabolas or balls
AVERAGES AND SINGULAR INTEGRALS
367
We now consider the intersection of two of these balls of the same size
Clearly one of these "balls" is not contained in a fixed multiple of the other (uniform in
E ).
We can also see the difficulty of using the Calderon-Zygmund theory to study Problem II'. Suppose y(t) is the parabola (t,t 2 ) in R 2 and
Tf(x,y) = Jf(x-t,y-t 2 )d: .
Then we may formally write
Tf(x,y)
=
=
If
Jf
f(x-t,y-st 2 ) 8(1;s) ds dt
f(x-t,y-v)
~ 8 (1- t~) dv dt .
Or 22) where 22A)
K(x,y) =
1.. 8 {1-x3
\
.!) , x2
368
STEPHEN WAINGER
The Calderon-Zygmund theory deals with convolution operators with kernels
K(x~y)
less than
but in their theory K(x+h 1y+k)- K(x 1 y) should be much
I
K(x~y)
if h and k are much smaller than x or y. However
for our K if (X Y) is a point on the curve y : x 2 and (x+h y+k) is not 1
1
on the curve, no cancellation in the difference K(r+h,y+k)- K{x,y) can occur no matter how small h and k are. I
The Calderon-Zygmund Theory is based on 4) tools a) The Fourier transform (The Fourier transform is even used in the L 1 theory) b) Interpolation c) Covering lemmas d) Calderon-Zygmund decomposition. Perhaps the natural attack on our problems would be to find appropriate covering lemmas and suitable variants of the Calderon-Zygmund decompcsition. Some progress in finding covering lemmas for related problems was made by Stromberg [Str] and [STRO] and Cordoba [CORl], [COR2], Cordoba and Fefferman [CFl], [CF2] 1 [CF3] 1 and Fefferman [FEf]. Our approach will however be different. We shall try to use the Fourier transform or other orthogonality methods and interpolation to reduce our problems on averages and singular integrals to the more standard averages and singular integrals. In retrospect we see that some of these ideas occurred in [SPL] 1 {CS] 1 and in [KS]. We have said earlier that curvature and Fourier Transform would be important for us. Actually they go together. If one has a nice measure on a curved surface, the Fourier transform of that measure decays at infinity even though the measure is singular. Let us consider some examples. Define, for a test function ¢,
IJ.(¢):
I 0
1
¢(tiO)dt.
AVERAGES AND SINGULAR INTEGRALS
ll
369
is supported on a straight line, namely the x-i:! xis, and
~(~,71) = ll(ei~xei71Y)"'
Jei~tdt 1
0
which is independent of 71 and hence cannot decay at infinity along the 71-axis. Now let us consider a measure supported on a parabola, 00
23)
v(c!J) =
I
-oo
Then
=
I
00
e-t2e~tei71t2 dt
.
-oo
This integral may be computed exactly by completing the square, and it is easy to see that
Thus V(~,71) tends to zero at infinity. Another example is afforded by rotationally invariant Lebesgue measure on the n-1 dimensional sphere lxl this measure by diL, we have for
gl
dll(g) = CnJn- 2 2
= 1 in Rn. If we denote
Rn
Cl~l) 1~1- {n~2)
•
370
STEPHEN WAINGER
See [SWE]. Thus
Of course we want to have a tool to estimate the Fourier transform of measures in general, not in just a few specific cases. This tool is
a lemma of Van Der Corput.
VANDER CORPUT'S LEMMA. Let h(t) be a real function. For some j, assume lh(j)(t)l ? .\. in an interval a ~ t ~ b. If j
=
1, assume also that
h'(t) is monotone, then
II I
b
exp (ih(t))dt
For the proof of Van Der Corput's lemma for j = 1 and 2 see [Z]. The proof for higher j is similar. Let us consider the measure
d~-t(c;b) =
24)
f
2
c;b(t.t 2 )dt.
1
Then
This integral cannot be evaluated explicitly, but we wish to see that Van Der Corput's lemma may be applied. We take h(t) = e-t-+ 71t 2 • First we use the fact that h .. (t) = 71· Thus by VanDer Corput's lemma with j = 2, we see
AVERAGES AND SINGULAR INTEGRALS
A ldtt((.T/)1 ~
25)
c
1
I I 112
(1 + T/ )
Now if 1111 < 1.{!, 8
So by Van Der Corput 's lemma with j = 1 , 26)
if
Putting 25) and 26) together we have 27)
for some 8 > 0. Stein pointed out in retrospect that we can already see from an estimate like 27) that dtt has interesting properties from the point of harmonic analysis -namely even though dtt is singular, Tf = dtt
*f
maps LP into L 2 continuously for some p < 2. For
371
372
STEPHEN WAINGER
The second integral is bounded if q' is sufficiently large which means for some q > 1. But then the first integral is bounded for f c LP where
!.p
+~
2q
=
1.
II. The Hilbert transform
alon~ curves
The first progress in our series of problems was made on the Hilbert transform along curves. The Hilbert transform along a curve can be thought d as a multiplier transformation 28) where
29)
my(e-) =
I
00
e~·y(t) dtt .
-oo
To see that 29) is true we may either substitute the formula
373
AVERAGES AND SINGULAR INTEGRALS
f(x)
=fe
-ifx f(~ )d~
into 13) or recognize the fact that Hf = D
*f
where D is a distribution
D¢ =f¢(y(t))dt .
So A
AA
Hf =Of,
and D may be computed by evaluating D on an exponential. Thus to prove that Hy is bounded on L2 one needs to show that my is bounded. The first result of this type was obtained by. Fabes [F]. Fabes showed Hy is bounded on L 2 in 2-dimensions for the curve y(t)
=
(t, ltlasgnt) ,
a> 0.
So Fabes' proof consisted in showing that the integral
J exp(it~+ilt\a(sgnt)Tt) d: 00
m(~,Tf) =
-oo
is uniformly bounded in ~ and Tf. To this end Fabes employed the method of steepest descents. The method of steepest descents is a method of obtaining very precise asymptotic information for large A about integrals of the form
J
exp(iAh(t))dt
374
STEPHEN WAINGER
by contour integration. However to employ the method one has to have very precise information on where the real part of h(z) is positive and negative in the complex plane. Thus already to employ the method of steepest descents for the curve (t,t 2,t 3 ), one would have to understand the zero set of
uniformly in ~ 1 , ~2 and ~ 3 • So it is hard to imagine using the method of steepest descents, and for the curve t, t 2 , t 3 , t 4 , t 5 it would seem close to impossible. Fabes' result was very important in that it gave the first clue that problems such as II and II' could have positive answers. However a better method would have to be found -a method that needed less precise information about h(t). The next step was to . a1 a2 an-1 show that 1f y(t),. (t,t ,t ,···,t ) , 1 < a 1 < a 2 < ··· < an_ 1 a. a. a. (here t J can mean either It! J or ltl J sgnt ) then Hy was bounded on L 2(Rn) [SWA]. Here we had to prove the boundedness of the integral
I
00
exp[i(e-1t+···+entan-1)] dtt.
-oo
The proof was by way of the Van Der Corput lemma but was unnecessarily complicated because at that time we only knew the lemma for j = 1,2. Let us see how Van Der Corputs "lemma works in the case y(t) = (t,t2, ... , tn). We then have to show that
30)
J exp(~1 t+···+i~ntn)~t ~
C(n).
e ~Jt!~R
where C(n) does not depend on E,Rl 1,···ln. We shall prove 30) by induction on n. By changing variables, replacing t by ___!__ 1 1 we l~nl n
375
AVERAGES AND SINGULAR INTEGRALS
may assume f n = ± 1 in 30). Then by using Van Der Corput's lemma with j = n , we find
J t
31)
exp(if1s+···+ifn_ 1sn-l±isn) :::c(n).
1
An integration by parts together with 31) shows that
J
32)
exp (if 1t + ··· + ifn_ 1tn- 1 ± itn)
~ < C(n) .
1:::1 t!:=:R
'I
Now
f
exp (if1 t + ... +if n-1tn-1 ± itn) dtt
e 0. Hence by Stein's interpolation theorem we would know that T 0 = T was bounded in all Lp, p S 2.
AVERAGES AND SINGULAR INTEGRALS
381
Then by a duality argument T would be bounded in all LP , 1 < p < oo. It turns out that one can show by a messy calculation that T z is of Calderon-Zygmund type if Re z
> 0.
Let us try to understand why the kernel for Tz, Re z > 0, might be a little better than the kernel for T0 . T is essentially Hy y = (t ,t 2 ), and so the kernel K 0 of T0 is essentially
from 22) and 22A). If we introduce "parabolic polar coordinates" y
=
r 2 sin 0 x
=
rcos 0 , we see
where sin 0 0 - - =1. cos 2 0 0 We might expect if Re z
=
e > 0 Kz to be e better than K 0 . So we
might expect
Now we would like to explain why a
1
IO-Ooll-e
singularity is better than
a 8(0-0 0 ) singularity. To see the situation more clearly, let us examine the analogous situation for the standard polar coordinates with 0 0 = 0. Suppose 53)
where x = rcos 0 and y = rsin 0 , and
382
STEPHEN WAINGER
54) (The factor _!. for ordinary polar coordinates plays the same role as 1 r2
r3
for parabolic polar coordinates.) We are trying to see whether
I
55)
)K(x,y)-K(x-h,y-k)l < C .
For either K = K0 or K = K£ . Let us take h = 0 and k = 1 and consider first K = K0 • Let us look at the contribution from y's which are very close to 0. We have
JJ
r>C nearo
If y is very close to 0 [0'1 "'
~ : y
r
0
1
~I ~y-1
So the left-hand side of 55) is at least f~
} dr
= "".
So 55) can't
hold. Let us put the matter a little differently. If we consider the 8's with ()
~
0 where the difference
offers no cancellation, we find there is only one bad 8, ()
=
0. But still
383
AVERAGES AND SINGULAR INTEGRALS
J•l•
jK 0 (,, 6)- K 0 ( 0 an operator of CalderonZygmund type. This proved that HY was bounded in LP 1 < p
0. Note that also the function mz(e,.,.,) defined in 52) also has this type of homogeneity, namely mz(Ae,A-2 .,.,) Now experience has shown that
=
mz(e,.,.,), for A.> 0.
h~mobeneity
is a powerful friend not
"-'
to be tossed away lightly. However Dp does not have this homogeneity. This situation can be remedied by defining
59) where 00
J
60)
-00
Note that for A> 0 61) Let us see how formula 61) can help us. We would like to show
62) if Re z > - ~. By formula 61) we may assume Then by Van Der Carput's lemma with j
=
Tf =
± 1 , let us say
2 , we see that
Tf =
1.
AVERAGES AND SINGULAR INTEGRALS
=I
t
ei(seis2 dsj :S C .
1
So an integration by parts shows
Joo (l+t4)-z/4ei(tei71t 2 d t 1
:S C(z)
1
(·1J=l). Now
-1
J 1
:S C(z)
t2
ftti 'S C(z) .
-1
But we already know that
I -1
1
e~tei11t2 ~t
0. Then
f J ""
00
dt
T
-00
-00
=
l
e-iqy (1 + 17 2t4)-E/4 ei7Jt 2
-00
J~
e i"(x 2-y )(I + •'x•) _,/4 d"
-00
where Pe/ 2 is a modified Poisson kernel. Pe/ 2 decays exponentially fast at oo and Pe; 2(u)"' C as u ... 0. See [SWE]. lull-E/2 Thus Ke<x ,y) has a singularity near the curve (t,t 2) of the form 1 which is just the improvement over the a-function that we llH:J 0 11-e/2 seek. A modification of these ideas worked for curves
AVERAGES AND SINGULAR INTEGRALS
y(t) = (t a1
387
at a2 an •t •...• t )
< a 2 ,- ··, < an . See [NRW]. However, there is a natural generalization of these curves. All of
these curves satisfy an equation of the form
63)
y'(t) =
~ y(t)
where A is a real nxn matrix such that the real parts of the eigenvalues of A are positive. For example if y(t)
=
(t 1 t 2 )
A=(~ ~). A curve satisfying 63) 1 where all the eigenvalues of A have positive real part is called a homogeneous curve. A will generate a group of transformations TA = exp(A log>.). Then
64) where
65) Moreover there is a distance pA (x) defined on R 0 such that
In the case of the cure (t 1 t 2 ) we may 1 as we said before take
388
STEPHEN WAINGER
Then
and
It turns out that in the case of a general homogeneous cure, we can
obtain a satisfactory analytic family of operators by defining
66) where
67)
p A
*
is the distance function corresponding to A*, the adjoint of A. For a detailed description of the argument see [SW]. Here we shall
just make a comment. If some of the eigenvalues of A have non-zero imaginary part, y(t) can be an infinite spiral. For example the curve y(t) = (tacos (/3logt), flsin ({3logt)) is an example. So one could believe it might be rather messy to prove integrals involving exp (ig·y(t)) to be bounded. It might be difficult to show that at each t some derivative of
~ · y(t)
However, if one makes a change of variables t
would be non-zero. =
eu, we would be led to
consideration of integrals involving ~ · 71(u) where 71(u) -= y(eu). If y'(t) = ~y(t), 71(u) satisfies
68)
71'(u) = A71(u) .
AVERAGES AND SINGULAR INTEGRALS
389
We shall show that if 77(u) is a curve in Rn satisfying 68) where the eigenvalues of A have positive real part, then either 17(u) lies in a proper subspace of Rn or for every
1
~ j
Sn
e-1 0
and u there is a j ,
such that dj -. e. 77(u) 1- 0 . du1
69)
From 68) we see that
By the Cayley-Hamilton theorem, we can find numbers aj , 0 O
IM 8
f(x)l ,. r
394
STEPHEN WAINGER
Stein used g-functions to prove THEOREM
3:
77) if p > ~ and n > 3.
n-1
Simple exampl,es of the form f(x)
_...;1/J~(_x)_ _ n-1 Jx\ -n loga _!_
= _
ix\
where 1/J = 1 near the origin and has compact support show that p > n~ 1 is necessary in order that 77) hold. The situation for n
=
2 , p > 2 is
unknown at this time. I would like to present here Stein's original argument which proved
77) for p = 2 and n = 4 . We define
78)
Assume that we could prove Jlg(f )1\ 2 ~ C(n) \If\\ 2 •
79)
L
L
and let us see how 77) would follow. Now r
rnMrf{x)
=I fs
snMsf(x)ds
0
J r
=
n
0
J r
sn- 1 Msf(x)ds
+
sn
0
d~ Msf{x)ds .
AVERAGES AND SINGULAR INTEGRALS
395
Thus
Mrf{x)
~~
J r
sn-t M8 f(x)ds +
:n
0 =
J r
sn
Js Msf(x)ds
0
I{r) + II(r) .
Now I(r) is dominated by the Hardy-Littlewood Maximal function and
ll{r)<S
r~J
r
sn-l/ 2 s 1 12 M8 f(x)ds
0
~1/2
r
~ .~ ~ (
~
s 2 "- 1
d~
g(f )(x)
Cg{f)(x).
So if we assume 29), we have llsup Ml(x)ll 2 ~ Cilfl\ 2 • L L We turn now to the proof of 79).
396
STEPHEN WAINGER
m(r)
1
=
n-2 1n-2(r) .
-T
r 2
Here
J n-2
is the usual Bessel function. We shall need to know
I..
0
J~
j lm(t\e!) \f(g)j
J J lfO
r
ldiLh *f(x,y)-ifih *f(x,y)l 2 dh
I~
J e
diLh *f(x,y)dh\
0
~ g(f )(x,y) +sup 11/!h *f(x,y)l . h>O
402
STEPHEN WAINGER
A classica 1 argument (see [R]) shows
I sup r/lh * £11 h>O
L
p
:S Cpll £11 p · L
Thus by 95), we see
suP'I\
e>o
Je
dllh*f(x,y)dhl
~ ~ Cl!£11 L2
0
If
L
2 ·
f~O.
}J e
dllh * f(x,y)dh
0
kI
2h
f(x-t,y-t 2 )dtdh
h
>
!.Je
- e
0
~ !-
J'
J
e
f(x-t,y-t 2 )
1 dh
11
E/2
f(x-t,y-t')dt ,
0
So from 96) we infer
!sup~ fe e>o
0
It remains to prove 96).
~
f(x-y,y-t 2 )dt
All£11 L
L2
2 •
403
AVERAGES AND SINGULAR INTEGRALS 00
Jj(gf)(x,y)j 2 dxdy =
f
~ Jjdp.h ... f-1/lh *fj2dxdy
0
=I""'~ I ~h((•.,>-¢h 0 we may write the expression in 98) as
f
00
~ l~(,\h(,,\2h2q}-~(,\h(,,\2h2q)j2
.
0
By choosing ,\ so that ..\2( 2 +..\4 71 2 == 1, we see that it suffices to estimate 99) when ( 2 + q 2 = 1 . In this case we see
J~ 1
99)
dp.(h(.h 2.,)-¢(h(.h 271)' 2
0
A
~c
f~ 1
0 ~
since dp.(O) = 1/1(0) = 1 .
dh
~c
404
STEPHEN WAINGER
Then from 27)
for some 8 > 0. So
r
'\\! if.;(he,h'•ll ~ { h'~~g ~ c .
100)
1
~
Also r/J((,TJ) S.
( (1 +
1
c
2 N 2 N for any N, so
+., )
f "" d: l~(h(,h2 .,)1
101)
<S
c.
1
Now we obtain 98) and hence 96) by combining 100), 101) and 102). In this section we have emphasized L 2 methods. LP results for p > 1, can be obtained by combining the L 2 estimates presented here with the techniques of section 2. Altogether one can prove the following theorems: THEOREM 4. If y(t) satisfies y' = ~ y(t) where all the eigenvalues of A have positive real part,
kJ jf(x-y(t))\dtiiLp <S ClifiiLP' h
II
sup o