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!
0 and inequality (4.2) 3, if 1
e Co C. It will suffice to prove the
f
1 /2
o
AIIfII p
Rn
for PO = 2n + 32, and f e 5; the case 1 < p < p0 will then follow by
interpolation.* By covering the support of V1 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 Now with dµ = V'da we support o f Vi) as a graph: en have that
J )?(e)I2dµ = J f(e)f(()d, = J T(f)(x)13dx where (Tf) (x) _ (f * K) (x), with
K(x) =
fe
*In 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.
327
OSCILLATORY INTEGRALS IN FOURIER ANALYSIS
Thus (4.2) follows from Holder's inequality if we can show that (4.3)
IIT(f )Ilpo < AIIfIIpo
where po is the dual exponent to p0. To prove (4.3) we consider the function Ks (initially defined for Re(s) > 0 ) by
(4.4) Ks(x) =
e s2 1'(s,'2)
fe
1+s
2nix
)I
7
W(e')de
n(
Rn
by e'; we have set (C') =
Here we have abbreviated
+Iph(e')I2)I/2 , so that (C')dC'= dµ; also function which equals
ri
is a Co(R)
near the origin. Now the change of variables en -+ en + 0(') in the above integral 1
shows that it equals
('
e 2ni(x'-e1+xn0(e'))Yf(e')de'
s(xn)
= Cs(xn)K(x)
Rn-I with 00
2
Cs(xn)
1'(s/2) ./
e
2nix nenlfnl-l+s n(fn)den
-00
Now it is well known that 4s has an analytic continuation in s which is an entire function; also Co = 1 ; and I4 (xn)I < clxnl-Re(s), where Ixnl > 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
E. M. STEIN
328
(a) KO(x) = K(x) ,
(b) IK-n/2+it(x)I < A, all x e R, all real (c) IK1+it(e)I < A, all
t
£ Rn, 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 T5(f) _ f * Ks' From (b) one has IIT_n/2+it(f )IIL00 < AIIf1ILI , all real t
(4.5)
and from (c) and Plancherel's theorem one gets (4.6)
IITI+it(f )IIL2 < AIIfIIL2
,
all real t
,
An application of a known convexity property of operators (see [281) then shows that IIT0(f )II LP°, < AIIfJJ PO , with PO0 = 2n + 2 , and the proof of n+3
Theorem 3 is complete. REMARKS:
(i)
For hypersurfaces with non-zero Gaussian curvature this theorem is the best possible, only insofar as it is of the form (4.1) with q > 2 . If q is not required to be 2 or greater, then it may be
conjectured that a restriction theorem holds for such hypersurfaces in the wider range 1 < p < 2n/(n+1). This is known to be true when n = 2 (see also §7 below). (ii) For hypersurfaces for which only k principal curvatures are nonvanishing, Greenleaf [121 has shown that then the corresponding results hold with 1 < p < 2k 2 , giving an extension of +
Theorem 3.
(iii) In the case of dim(S) = 1 (i.e. in the case of a curve) there are a series of results extending our knowledge of the case n = 2 alluded to above. For further details one should consult the references cited below.
OSCILLATORY INTEGRALS IN FOURIER ANALYSIS
329
It would of course be of interest to know what are the exponents p and q (if any) for which the restriction holds if we are dealing with a given sub-manifold S. This problem is highlighted by the fact quite general sub-manifolds S (those which are of finite type in the sense described in §3) have the restriction property: THEOREM 4. Suppose S is a smooth m-dimensional sub-manifold of Rn of finite type. Then there exists a po = po(S), 1 < po, so that S has the LP restriction property (4.1) with q = 2 , and 1 < p < po. (In fact if
the type of S is k, we can take po = 2nk/(2nk-1 .) COROLLARY. Suppose S is real analytic and does not lie in any affine hyperplane. Then S has the LP restriction property for 1 < p < po, for some po > 1 .
Proof. As we saw above, it suffices to prove (4.3). However Tf = f *K, and K(x) = dµ(-x), Theorem 2 tells us that IK(x)l < AIxI-11k, So according to the theorem of fractional integration, (see [26], Chapter V), where a=n-1/k, and this we therefore get (4.3) with __ = P -n,
0
0
relation among exponents is the same as PO = 2nknk1
'
Q.E.D.
Further bibliographic remarks. The initial restriction theorem dates from 1967 but was unpublished. The sharp result for n = 2 was observed by C. Fefferman and the author and can be found essentially in [9]; see also Zygmund [33]. Further results are in Thomas [30], [31], Strichartz [29], Prestini [24], Christ [4], and Drury [7].
Oscillatory integrals of the first kind related to singular integrals A key oscillatory integral used in the theory of Hilbert transforms along curves is the following: 5.
00
(5.1)
P.V.
Jr
elpa(t) dt t,
E. M. STEIN
330
d
where Pa(t) is a real polynomial in t of degree d, Pa(t) = F a.O. j=o
It
was proved by Wainger and the author in [27], that the integral is bounded with a bound depending only on the degree d and independent of the
coefficients ao,al,...,ad. The relevance of such integrals can be better understood by consulting Wainger's lectures [32]. We shall be interested here in giving an n=dimensional generalization of this result. We formulate it as follows. Let K(x) be a homogeneous function of degree -n; suppose also that IK(x)I < AIxI-n (i.e. K is bounded on the unit sphere); moreover, we assume the usual cancellation property: f x,l-1 K(x') d a (x') = 0. We let P(x) _
I aaxa be any real polynomial of degree d. Ial
THEOREM 5:
P.V.
(5.2)
f
eiP(x)K(x)dx < Ad
Rn
with the bound Ad that depends only on K and d, and not on the coefficients as . Nagel and Wainger observed that if K were odd, one could prove (5.2) from the one-dimensional form (5.1) by the method of rotations (passage to polar coordinates). To deal with the general case we need two lemmas. d
Let Pa(t) = F a tJ denote a real polynomial on R1 , and write also j=1
d
Pb(t) = F j=1
j
OSCILLATORY INTEGRALS IN FOURIER ANALYSIS
331
LEMMA 1:
Jipa(t)
(e
(5.3)
-e
1Pb(t) dt ) t
d
log`-II,
e
with Ad independent of (aj f ,
I bj {
and e > 0.
F aaxa be a homogeneous polynomial of
LEMMA 2. Let P(x) =
Iaj=d
degree d on R'. Write mp
=f
IP(x')I da(x')
.
Ix k=1
Then,
f
(5.4)
Ilog
(I P 1.)) da(x') < B
,
Ix'I=1
with Bd independent of P.*
One can if one wishes give an elementary (but complicated) proof of Lemma 2. It may however be more interesting to obtain it as a conse-
quence of a general property of polynomials in Rn related to the class of functions of bounded mean oscillation. This property can be stated as follows. It is very well known that the function log jxj is in B.M.O., and this is usually the first example discussed in that theory. It is surprising therefore that the following natural generalization seems to be been overlooked.
Of course in writing (5.4) we assume that P is not identically zero, i.e. mp > 0.
E. M. STEIN
332
THEOREM 6. Let P(x) be any polynomial of degree < d in Rn logIP(x)I is in B.M.O. and in addition
.
Then
11 log IP(x)1 IIBMO < Bd
where Ba depends only on d, and not otherwise on P.
The proofs of Theorem 6 and Lemma 2 will be given in an appendix. We now pass to the proof of Lemma 1. We prove it by induction on d.
The case d = 1 , i.e. the estimate 00
5
(eiat_eibt) dt
t
)
E
is classical. Let us now assume (5.3) for polynomials of degree d - 1 , and observe that the estimates (5.3) we wish to prove are unchanged if we replace t by bt , 5 ;E 0. Thus we may assume that bd = 1 and Iadi < 1 . Now write 00
r (e1Pa(t)
- e iPb(t))
dt
J
=
5+ r J
t E
E
00
1
and we treat these two integrals separately. (If e'> 1 , we have only f"0
and that integral is estimate like
f00
.) Let us consider the second
integral. It equals 00
00
f eiPa(t) dt _ (' eiPb(t) dt J t t
f
Now since bd = 1, we see that (d/dt)d Pb(t) = d!, and hence
OSCILLATORY INTEGRALS IN FOURIER ANALYSIS
333
(' eipb(t) dt
1
by the corollary of Proposition 2 in §1. Next, by a change of variables t lad-1 /d t , the integral 00
t 1
becomes 00
e iVa(t)
dt
1
=
ladll/d
f +f ladll/d 1
where Pa(t) is a polynomial of degree d , with td having coefficient one. Again e"Va(t) dt t
while 1
<
r
dt_1
t d log l(Tl=a to
J
1
ladl1/d
since bd = 1 . Next
f(e Pa(t)
IPb(t)
i
E
a
d_ )t-J E
1
(e
)dtt+ 0('dt J t,
1Qa(t) e 'Qb(t)
E
E. M. STEIN
334 with d-1
d-1
Qa(t) = I ajtj ,
4 bjti
Qb(t) =
J=1
j=1
since IPa(t)-Qa(t)I < Iti and IPb(t)-Qb(t)I < Iti . However, by induction hypothesis (using (5.3) for E' = E , and E =1, and d - 1 ), 1
r (elQa(t) _ e iQb(t)
J
dt
Ad-1
t
Ilog
1+ J=1
E
Gathering all these terms together then proves (5.3). Armed with Lemmas 1 and 2 we can now prove Theorem 5. We may assume that P(x) = F axaa has no constant term, and using polar IaI
d
coordinates x = tx', t > 0, Ix'I = 1, we write P(x)
j=1
P-(x')ti
,
j
where Pj(x') are restrictions to the unit sphere of homogeneous polynomials of degree j . Let us also set mj = f x'I-I I Pj(x')I dv(x) , and write K(x) = t-n1Z(x'), with 1 bounded and f
x'I-11Z(x')dv(x') = 0.
Then to prove (5.2) it suffices to show that
el1(x)K(x)dx
with Ad independent of written as
('
J
E1 ,
E2 and P. The above integral can be
(fE2 e
Ix l=1
1FPj(x')tj
SZ(x")do(x')
E1
Since 0 has vanishing mean-value this integral may be rewritten as
OSCILLATORY INTEGRALS IN FOURIER ANALYSIS
f f
E
335
2
Ix'I=1 \ E1
However by Lemma 1 the inner integral is bounded by d +
Aa 1
m)
and so an appeal to Lemma 2 shows that fE2
exp (i
f Ix'1=1
mit) dt (sup IH(x')I)da(x') < Ad
Pj(x')ti) - exp (i
EI
proving (5.2) and the theorem. 6. Oscillatory integrals of the second kind: an example related to the
Heisenberg group
To motivate the interest in this example we recall the definition of the Heisenberg group Hm. The underlying space of Hm is Cm x R, i.e. Hm = #(z,t){, with z e Cm, t e R; the multiplication here is (z,t)(w,s) _ M
(z+w, t+s+
Now on the Heisenberg group one can consider two types of dilations and their corresponding quasi-distances. The first are the usual dilations (z,t) (pz, pt), p > 0, and the metric could be defined in terms of the
usual distance. The second are the dilations (z,t) -+ (pz,p2t), and the appropriate quasi-distance (from the origin) is then (Izl4+t2)1 /4. 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 Cn+1 . This point of view, as well as related generalizations, is elaborated in Nagel's lectures [211.
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 R2m+1 . However the two conflicting types of dilations (and related metrics) occur in e.g. the solutions of Ju = f . (One sees this for example in Krantz's lectures [17], where in the formula of Henkin we have a kernel made of products of functions each belonging 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 (6.1)
Tf =f*K
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)S(t) .
L(z) is a standard Calder6n-Zygmund kernel in Cm = R2m, i.e. L(pz) _ p-2m L(z), L is smooth away from the origin, and L has vanishing mean value on the unit sphere. Here 6(t) is the Dirac delta function in the t-variable, and in an obvious sense is homogeneous 3(pt) = p-1 S(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 L2(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 TA defined by
*In particular the terms AI and A2 that appear in §6 of [17].
OSCILLATORY INTEGRALS IN FOURIER ANALYSIS
(6.3)
337
(TA)(F)(z) = i L(z-w)e1A
(with
in A, -DO«
,
>_
B( , ) , and F = f. Then the operators TA have the form
(6.4)
(Tf)(x) =
K(x-y)eiB(x,Y)f(y)dy
I
Rn
We shall suppose B is a real bilinear form, but we shall not suppose that B is necessarily anti-symmetric nor that K is homogeneous of
degree -n. THEOREM 7. Suppose K is homogeneous of degree -µ, 0 < g < n, smooth away from the origin, and with vanishing mean-value when p = n. (a) If B is non-degenerate, then the operator T given by (6.4) is
bounded on L2(Rn) to itself, for 0 < p < n ; when 1 < p < oo , the operator is bounded on Lp(Rn) to itself if -11 I < 2n I1
.
(b) If we drop the assumption that B is non-degenerate but require that
g = n, then T is bounded on LP(Rn) to itself for 1 < p < -o. The bound of T can then be taken to be independent of B. We shall give only the highlights of the proof, leaving the details, further variants, and applications to the papers cited below. Let us con-
sider first the L2 part of assertion (a) when n/2 < g < n. Suppose
ri
*For further details see Mauceri, Picardello and Ricci [19] and Geller and Stein [10].
E. M. STEIN
338
is a Co function, with 77(x)=l for IxI < 1/2, and ii(x) = 0, for 1x1 > 1 .
To is defined as in (6.4), but
We write T =
with K replaced by Ko = riK, and T with K replaced by K = (1-77)K. Observe first that since Ko(x-y) is supported where Ix-yl < 1 , estimating T0(f)(x) in the ball IxI < 1 involves only f(y) in the ball IYI < 2. We claim
5Ixl
(6.5)
ITo(f)(x)12dx < A
f
If(Y)I2dy .
IYI<_2
In fact when IxI < 1 ,
fKo(x_y)ei'cf(y)dy -J
Ko(x-Y)eis(y,Y)f(y)dy
Ix-YI-'`+'If(Y)IdY
,
Ix-YI
and thus the L2 theory for f
Ko (which is non-trivial only when
µ = n ) proves (6.5). While operators of the type (6.4) are not translation invariant they do
satisfy r_hTrhf = e1B(h,h)e1B(x,h)T(eiB(h*)f( ))
(6.6)
with rh(f)(x) = f(x-h). Applying this to To gives the following generalization of (6.5)
J Ix-hi
ITo(f)(x)I2dx < A
r Iy-hI<2
If(Y)I2dY
OSCILLATORY INTEGRALS IN FOURIER ANALYSIS
339
and an integration in h shows that as a consequence
f IT0(f)(x)I2dx < A2n J
If(y)12dy
Rn
Rn
We now turn to the proof of
J IT f(x)I2dx < A Rn
f
If(x)I2dx
Rn
This will be done by proving the corresponding result for the operator
T,*,T. The kernel L of this operator is given by L(x,y) = J e-iB(z,x-Y)1c(z-x)K.(z-y)dz
Now since K. is in L2(Rn) (here the assumption n/2 < µ is used), Schwarz's inequality implies IL(x,y)I < A .
that
We next integrate by parts in the definition of L(x,y), using the fact (Dz)Ne-iB(z,x-y) = e-iB(z,x-Y), where Dz = i(a,Vz)/Ix-yI , with
a = B-I
(Ix yI ` x-
,
and B denotes the matrix so that B(x,y) _ (Bx,y).
The result is IL(x,y)I < ANIx-yI-N, for every N > 0, and hence (6.7)
IL(x,y)I < AN(1 + Ix-yI)-N
,
N>0.
This shows that L is the kernel of a bounded operator on L2 proving the boundedness of T ,T and thus of T.. The proofs of the L2 boundedness when 0 < µ < n/2 (in part (a) of the theorem), and the L2 boundedness when µ = n but when B is not assumed to be nondegenerate, are refinements of the above argument.
E. M. STEIN
340
Let us now describe the main idea in proving the LP inequalities stated in (a) and (b) above. We shall need a generalization of BMO (and of H1 ) which may be of interest in its own right. Suppose E = leQI is a mapping from the collection of cubes Q in Rn to complex-valued functions on Rn so that IeQ(x)I =
,
Q(x),
all x
where XQ denotes the characteristic function of the cube Q. Let us define on "E-atom" to be a function a so that for some cube Q (i) a is supported in Q (ii) Ia(x)I < 1/IQI (iii) ,f a(x) Q(x) dx = 0
The space HE is then given by If if = I Ajaj , with each aj an E atom, and I IAil < ool. In a similar vein the function fE will be defined as
(6.8)
fE(x)
su IQI
I
!f -fQldx
,
Q
where
fQ
IQI
f
f U_Q
Q
and we take BMOE = IfIfE E L°°I .
Some of the basic facts about the standard H1 and BMO spaces* go through for HE and BMOE, and sometimes these come free of charge.
One such case is the following assertion: Suppose f c Lpo , 1 < p < p0 < oo, and fE f LP. Then f c LP and *The standard situation arises of course when eQ = yQ, all Q.
OSCILLATORY INTEGRALS IN FOURIER ANALYSIS
Lp
OilLp
(6.9)
341
To prove this we need only observe that (if I)# < 2f0 , and use the result (see [10]) for the standard # function. The point of all of this is that for operators of the form (6.4), there is a naturally associated HE and BMOE theory, and it is given by choosing eQ(x) = e
(6.10)
-iB(x,cQ) ,
where cQ is the center of the cube Q. The basic step in the LP theory (besides an appropriate interpolation which goes via (6.9)), is the proof that when u = n our operator T maps L°° to BMOE . Let us give the proof in the case (a). We may assume that ilf 11 °° < 1 , and
suppose first that Q is a cube centered at the origin. Then we have to show that there exists a constant yQ , so that
(6.11)
1
Q,
f
JTf -yQ dx < A
Q
The corresponding inequality for a cube centered at another point, say CQ , then follows from the translation formula (6.6), (and this is the reason for defining eQ as we do). Turning to (6.11), the argument is not exactly the same as in the standard case (see e.g. Coifman's lectures [5] or [91), since we must split f into three parts to take into account the oscillations of e1B(x,y) . Suppose Q = QS , has side-lenghts S , then write
f = fl+f2+f3, where
Q28' fl = 0 elsewhere, CQ25 n QS-1
,
f2 = 0 elsewhere,
SQ2S n cQ 5-1 , f3 = 0 elsewhere.
E. M. STEIN
342
(Note that f2 occurs only when S < x/2/2 .) We have F = T(f) =
F1+F2+ F3, where FJ =T(fj). For F1 we make the usual estimate, using the fact that T is bounded on L2. Next observe that IK(x-y)eis(x,Y)- K(-Y)I < cS
if
1
1
+
IYIn+1
IYIn-1
'
x cQs and y e'-Q2s. Thus if yQ = f K(-y)f2(y)dy, we get that for
xCQ3
IF2(x)-yQI
f
dy
dy
+S
l
IYI°-1
IYIn+1
Qs-1
Q2S
Finally
K(x-Y)eiB(x,Y)f3(Y)dY
F3(x) = I
= J (K(x-Y)-K(-Y))eiB(x,Y)f3(Y)dY + J K(-Y)f3(Y)eiB(x,Y)dY
For F3 we again make the standard estimates, and for F3 we use Plancherel's formula (which we may since we have assumed that B(x,y) is non-degenerate). The result is
5 IF(x)I2dx < J IF3(x)I2dx = A J 3 Q
Rn
Rn
den
IK(-Y)f3(Y)I2dy < Al J
IYI
cQ S-1
Combining these estimates proves (6.11), and hence the fact that T takes L°° to BMOE .
=
A5"-
OSCILLATORY INTEGRALS IN FOURIER ANALYSIS
343
We shall now state a generalization of this theorem which also includes the oscillatory integral result given in Theorem 5 (in §5). Suppose P(x,y) is a real polynomial on Rn xRn of total degree d . Consider the operator (6.12)
(Tf) (x) = p.v.
I
e'p(x'y) K(x-y) f (y) dy
Rn
where K is homogeneous of degree -n, smooth away from the origin and with vanishing mean-value.
THEOREM 8. The operator T given by (6.12) is bounded on L2(Rn) to itself, with a bound that can be taken to depend only on K and the degree d of P, and is otherwise independent of P.
This is a recent result obtained jointly with F. Ricci. The proof is based in part on a combination of ideas used in the proof Theorems 5 and 7. This result has also many variants, and we now state some of these: (i)
One may also show that the operators (6.12) are bounded on LP,
1
is homogeneous of degree - g, n - E < µ < n , T is still bounded on LP. However now the bounds may depend on P, and in addition one must assume that P(x,y) is not of the term P(x,y) = P0(x) + Pl(y). (iii) One can replace K(x-y) in (6.12) by a more general "Calder6nZygmund kernel" K(x,y), a distribution for which the operator when P e 0 is bounded in L2 , and which in addition is a function (when x y ) which satisfies IK(x,y)l < Alx-yl-n, lpxK(x,y)l + IoyK(x,y)I <
Alx-yl-n-I
References. For the detailed proof of Theorem 7, other variants, and applications to the a Neumann problem see the papers of Phong and the author [22], [231.
344
E. M. STEIN
7. Further oscillatory integrals related to restriction theorems and
Bochner-Riesz summability
We have seen that if S is a hypersurface in Rn with non-vanishing Gaussian curvature, then
fei(x)du(x) =
0(I6I-(n-I)/2)
as e
00
S
whenever 1 e Co , and this is a typical oscillatory integral of the first kind. We may pass to an oscillatory integral of the second kind when we replace the Co function l by an Lr function f , and consider the resulting linear operator on f . The resulting operator, as is easy to observe, is in fact the dual to the restriction operator considered in §4. Hence by Theorem 3 we can state that operator is in fact bounded from L2(da) to L p (Rn) , where p' is the dual exponent to 2n + 2 We shall
n+3
now describe the sharper result in this setting that can be obtained for n = 2. We fix a curve t -Y(t) = (t, y2(t)) , 0< t < 1, lying in R2 , with y(t) a C2 , and having non-vanishing curvature, i.e. Iy 2(t)I > c > 0. Con-
sider the transformation T, which maps function on the interval [0,1] to functions on R2 , given by
(7.1)
(Tf)( ) - f e4'y(t)f(t)dt
.
0
THEOREM 9. Under the assumption above T is bounded from Lp[0,1]
to Lq(R2), whenever 3/q + 1/p = 1 and 1 < p < 4. (Note that when p -. 4
and q .)
,
then q -. 4 in the above relation between p
OSCILLATORY INTEGRALS IN FOURIER ANALYSIS
345
r(7.2) f
Proof. Write
1
F() = ((Tf)( ))2
0
J0
e e-(Y(s)+Y(t))f(s)f(t)ds dt ,
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 , which divides [0,11 x [0,1] into the union of two regions R1 and R2 . We then consider the mapping of R1 R2 given by x = y(s) + y(t), i.e. x1 = S + t , X2 = y2(s) + y2(t). It is easy to verify on the basis of our assump-
tions that this mapping is one-one, and its Jacobian j satisfies IJI = Iy2(s)-y2(t)I > cls-tI . Therefore
(7.3)
r e'6'(Y(s)+y(t))f(s)f(t)dsdt
=
f e'C'xf(x1,x2)dxldx2
R1
R2
with f(x1,x2) = f(s)f(t)1J1_1
the quantity appearing in (7.3) then when-
So if we denote by
ever 1 < r < 2 , and 1/r'+ 1/r = 1 , we know that (7.4)
IIF111Lr(R2) <- cOf1ILr(R2)
However
IIfIILr(R2)
5f12t12 =
f
If(t)Ir If(t)I`IJI I-r ds dt
f
If(t)Ir If(t)Ir Is-tl I-rds dt
346
E. M. STEIN
To estimate the last integral we need to invoke the theorem of fractional integration in one dimension in the form
fg(s)g(t)(st)_1+adsdt < AIIgflu , u - 1 = a, 0 < a < 1
.
So we take g(t) = If(t)Ir, then Ilgllu = IIfIIp when p = ur. Then if we fix
a so that -1 +a =1-r, then
3r
2 rr = u , and
The limitation
0 < a becomes r < 2, and with q = 2r' we obtain from (7.4) that 1
X2/P
IIFiIILr (R2) < c' 0
with a similar estimate for F2(e) which is the analogue of (7.4), but taken over R2 . Since F = FI +F2 and F = (Tf )2 we obtain IIT(f )IILq(R2) <- Ai1fIILp(0,ii
Note that Q =_i=32r3=1-P , so 1 < r < 2 is equivalent with 1 < p = 3
q r
+P =1, and the limitation < 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 L2 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 A , A > 0, defined by
TA(f)(e) = J Rn
ei1'(x,4)&(x,e)f(x)dx
OSCILLATORY INTEGRALS IN FOURIER ANALYSIS
347
where 0 is a fixed Co (Rn x Rn) cut-off function; (D is a real-valued C°° phase function which we assume satisfies the assumption that its Hessian is non-vanishing, i.e.
axi°j det a2L(x, e)
(7.6)
0
PROPOSITION: (7.7)
[ITA(f)IIL2(Rn) <
A)C-n/2IIfJIL2(R')
.
COROLLARY: (7.8)
IITA(f )II
Lp (Rn)
< AAn/P IIfff
LP(Rn)
,
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 - oc. This decrease is consistent with the special case when t(x,6) 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 the M. 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 AK-n. T*TA is bounded by Now this operator has as its kernel the function KA(e,n) given by
(7.9)
J Rn
348
E. M. STEIN
Now since
L
(a,Vx)[c(x,n)-F(x,e)] =l _ _ a,17-e'\ + OIn-ej2 OXOV
we can find a = (a,,...,an), so that the aj depend smoothly on x and (a,px). IA(x,e,n) >- cIe-nI on the support of KA(e,n). Set Dx = 0 = ei\(((x,n)-D(x,e)) , we can Then since integrate by parts N times in (7.9) and obtain (Dx)NeA(t(x,n}-'P(x,6))
(7.10)
IKA(e,n)I < AN(1
+ale-nI)-N , N > 0 .
It follows from (7.10) with N = n+1 , that the operator TX*TA which has kernel KA has a norm bounded by AKn and the proposition is proved. We shall now formulate some theorems for oscillatory integrals of the form (7.11)
ei
(TAf)(e) = J
(t.4)c&(t,e)f(t)dt,
6 E Rn
Rn-I 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 cF are as follows: We consider for each fixed (t°,e 0) the associated bilinear form B(u,v) defined by B(u,v) = (v,Vt) (u, Ve) (c) (to,4 0), with u E v e Rn . Our first assumption is that (7.12a)
B is of rank n - 1
Rn-1
.
Thus there exists (an essentially unique), u E R n, IuI = 1 , so that the scalar function t - (u,V,(D(t,e)) has a critical point at (t°,e°). 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:
349
OSCILLATORY INTEGRALS IN FOURIER ANALYSIS
det (pt (G,Oe1)) (t°, e 0) - 0
(7.12b)
These assumptions will be supposed to hold at all (t°,6 °) in the support of qi(t,e), where i/i is a fixed function in Cc (Rn-1 xRn). THEOREM 10. Under the assumptions above the operator (7.11) satisfies IITx(f )IIq < Ak n/gllf IIp
(7.13)
with q = n-1 p',
(n+1)
p +p 1 1,
if 1
REMARKS:
(1) When
(D(t,6)=t161+t262+...+tn-1en-1+c(ti,...,tn-1).en
,
and
(V)(0) = 0, then the conditions (7.12) are near the origin equivalent with the non-vanishing Gaussian curvature of the graph tn 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 Hormander [15]. Since part (b) has not appeared before, we will outline its proof. This will also serve as a good review of many of 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)
LIT, (F)IIL2(Rn-1) < AX n/r'IIFIILT(Rn)
350
E. M. STEIN
with r = 2(n+
,
where
(Ta)(F)(t) = J e-iAD(t14)0(t,e)F(e)de, t e
Rn-1
Rn We can calculate
J
TX*(F)Tj*(F)dt
Rn-I and write as
KA(6,r!)F(e)F(n)d6drl
J RnxRn
with
(7.15)
KA(e,r1) = J
e'
Rn-1
It suffices therefore to see that K,\ is the kernel of a bounded operator from Lr(Rn) to Lr(Rn), with norm not exceeding AA 2n/r' Because of our assumptions on 4) we can construct a phase function on Rn x Rn so that the following holds: we will write x e Rn, as Rn-1 . we can construct will The (t,xn) with t = satisfy: (i) $(x,e)_O(t,e)+(Dp(6)xn
(ii) the determinant of the nxn matrix pxVe$ is non-vanishing.
OSCILLATORY INTEGRALS IN FOURIER ANALYSIS
351
In fact Vtpe(b already has rank n-1 by assumption (7.12a), and so we need only choose (Do(e) so that (u,V,)Io(e) 0 to increase the rank of px e to n . Now, as in the proof of Theorem 3 in §4, we form Ka defined by es2
f eia( (x,n
I'(s/2) J Rn
with dx = dtdxn , and where v is a Co function which equals
1
near
the origin. We easily verify (7.16)
since when x = (t, 0) .
(V(x,
Next (7.17)
Kx+it is the kernel of a bounded operator from L2(Rn) to itself with norm < AX -n/2.
This follows by applying the estimate (7.7) of the proposition above and
using the non-degeneracy of the Hessian of $(x,e). Finally we claim that (7.18)
I K-n /2+1 /2+it(e ,l)I < A A
To see this write KAs(6,rq) as KA(6,r?)v(k(4)o(rl)-(Do(6))) where 00
2
vs(u) = F(ss/2)
r eixnuv(xn)+xnl-1+sdxn -Do
.
352
E. M. STEIN
Then since Iv_n/2+1/2+it(u)1-< cluln/2-1/2 , as u
00 we see that to
prove (7.18) it suffices to show that (7.19)
IKX(e,q)I
In 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 = to , (for otherwise we can write it as the sum of
finitely many such terms). When we write I(t,ri) = I(t,e) = (VeO)(t,77) ('i-e) + 0(77-e)2 we see that these are two cases to consider as in the
proof of Theorem 1 in §3: 1° when the directions n - 6 or e - n 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). In the second case, we actually get 0(AI,7-e I)-N , for every N > 0 as an estimate, by Proposition 4. This completes the proof of (7.18), and shows that Kin/2+1 /2+it is the kernel of a bounded operator from L1(Rn) to L°°(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 §5. First let `f'd denote the linear space of polynomials in Rn of degree < d . We claim that there is a constant Ad , so that 1 /2
(8.1)
IQI J (1 Q
IP(x)I2 dx
< Ad
IQI
f IP(x)I dx
holds for all P c 'd , and all cubes Q. The space Td is invariant under translations and dilations, and so a moment's reflection shows that to prove (8.1) for all P t Td , it suffices to prove it for Q = Q0 , the unit cube centered at the origin. However
OSCILLATORY INTEGRALS IN FOURIER ANALYSIS
(f
IP(x)I2 dx)1
2 and f
Q0
353
IP(x)ldx are two (equivalent) norms on the
Q0
finite-dimensional space 5d , so (8.1) holds for Q = Q0, and then for general Q. Now it is well known (see e.g. [6]) that a function which satisfies a "reverse Holder" inequality belongs to the weight space A.. Examining the proof of this fact one obtains an r = r(d), 0 < r < co, and a constant Cd , so that (8.2)
(7Q,1 f IP(x)I dx
f
(fQ_o I
Q
1Jr
IP(x)I"dx
<< Cd
for all cubes Q. From (8.2) and Jensen's inequality, Theorem 6 follows easily.
Let us now assume that P is homogeneous of degree d. Observe also that since (8.2) holds, if we normalize P by the condition that mp = fIXI_1 I P(x)l da(x) = 1 , we can conclude that
f
(8.3)
I P(x)I-r da(x) < ca
.
IXI=1
However, when u > 0, Ilog ul = log+u + log+u < u + u -r. Therefore (8.3) implies (5.4) whenever mp = 1 , and so that result also holds in
i
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.
3S4 [3]
[4] [5]
[6]
E. M. STEIN
L. Carleson and P. Sjolin, "Oscillatory integrals and a multiplier problem for the disc," Studia Math. 44(1972), 287-299. M. Christ, "On the restriction of the Fourier transform to curves," Trans. Amer. Math. Soc., 287(1985), 223-238. R. Coifman and Y. Meyer, in these proceedings. R. Coifman and C. Fefferman, "Weighted norm inequalities for maximal functions and singular integrals," Studia Math. 51 (1979), 241-250.
[7]
[8] [9]
S. Drury, "Restrictions of Fourier transforms to curves," preprint. A. Erdelyi, "Asymptotics Expansions," 1956, Dover Publication. C. Fefferman, "Inequalities for strongly singular convolution operators," Acta Math. 124 (1970), 9-36.
[10] C. Fefferman and E.M. Stein, "HU spaces of several variables," Acta Math. 129 (1972), 137-193.
[11] D. Geller and E. M. Stein, "Estimates for singular 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 Integrale auf konvexen Korper. I," Monatsh. Math. 54 (1950), 1-36.
[15] L. Hormander, "Oscillatory integrals and multipliers on FLU," Ark. Mat. 11 (1973), 1-11.
, "The analysis of linear partial differential operators. I,"
[16]
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 #51, 1963, Princeton University Press.
[21] A. Nagel, "Vector fields and nonisotropic metrics," in these proceedings.
OSCILLATORY INTEGRALS 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.
, "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 of a convex set," Trans. Amer. Math. Soc. 139(1%9), 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. [29] R. S. Strichartz, "Restrictions of Fourier transforms to quadratic surfaces and decay of solutions of wave equations," Duke Math. J. [23]
44 (1977), 705-713.
[30] P.A. Tomas, A restriction theorem for the Fourier transform," Bull. A.M.S. 81 (1975), 477-478.
, "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
[31]
sets," in these proceedings.
[33] 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(l) 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
I f f(x-y)dy
f(x) = limo IQ r
a.e.
(where Qr is the square, Q. =jxcRnIsupIxil
f(x) = lim 1
r
r-+0 JBr) J
f(x-y)dt
Br
(where B. is the ball, Br = {xI lxI
*Supported in part by a grant from the National Science Foundation.
357
a.e.,
358
STEPHEN WAINGER
Our first problems are the following: Problem IA: Does
1)
f(x-y)dar(y) = f(x)
lim
a.e.?
r-+0
aQr
Here aQr denotes the boundary of Q. and dar is n-1 dimensional Lebesque measure on aQr normalized so that dar(aQr) = 1 . Problem IB:
Does
lim J
2)
r-O
f(x-y)dp = f(x)
a.e.?
aBr
Here di 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-, L2,
or LI. In questions IA and IB, we are considering certain averages
3)
MQ f(x) = r
r
,J
f(x-y)dar(y)
aQr and
4)
MB
r
f(x) = J
f(x-y)dkr(y)
aBr
We are asking if 5)
MQr f(x)
f(x)
a.e.
AVERAGES AND SINGULAR INTEGRALS
359
MBrf(X) - f(x)
a.e.
and 6)
The standard approach to this type of problem involves consideration of appropriate maximal functions. We define the maximal functions )lIQf(x) = sup MQ (If1)(x) r>0 r
7)
and
8)
)RBf(x) = sup MB (IfI)(x) r r> O
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 < p < °o , if we could show 9fIB
9)
11XIM Qf(x)>AlI <--9 1'fI'p
for every f in LP, and 2) would follow if we could show 10)
I1xI)RBf(x)>x1I <-L IIfIIP
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°° functions. In our case this means the following;
Let E be a measurable set and XE its characteristic function. Then if (11)
I1xIRQXE(x)>a1I < C(a) I E I
360
STEPHEN WAINGER
where C(a) may depend on A but not on E, then 1) holds for every f in L°°. If 12)
I1xI911BXE(x)>a1I < C(A)JEJ
then 2) holds for every f in
,
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<x
[f(x+r)+f(x-r)]
MQrf(x) = MBrf(x) = 2
So if we take f(x) to be log L near x = 0, have compact support, Ix
and be in C°° away from the origin, we would have a function f in every LP class such that )IIQf(x) =)IIBf(x) = oo for every x . We can also see that 11) and 12) are false in one dimension. We just take EE = Ix 10<x<E} .
Then
J EEl
0 but
)IlQf(x) = 911Bf(x) > 1
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 [SI] of 9), 10), 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 < p < 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°° functions. See [SW]. Thus we obtain negative results in one dimension. Similar reasoning gives the same negative conclusion for question IA in any number of dimensions.
361
AVERAGES AND SINGULAR INTEGRALS
One need only consider
f
is as
above and h is a nice function. So there are no interesting positive results in problem IA. However as we shall see later there are positive results for problem IB 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 boundary 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 of 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 h J f(x-y(t))dt = f(x) a.e., for
f in
L'° or L2 or LI ?
Problem III: Let v(x) be a smooth vector field in Rn. Does h
lim
1
h-'0 h _
1
f(x-tv(x))dt = f(x)
a.e.
0
for f in
L°° or L2 or LI ?
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 a
13)
Hyf(x) =
J-a
f (x-y(t)) Lt .
(where sometimes we wish to think of a as finite and sometimes as and
00 ),
362
STEPHEN WAINGER
1
14)
Hvf(x)
f(x-tv(x)) dt
-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)
IIHyf 11
LP
< CPIIIIILP
for some p's ? Problem III': Can we have an estimate 16)
IIHyfIILp <_ CpIIfilLP
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
Tf = K*f
where K is a kernel having "standard homogeneity" that is 17)
K(Ax) = A-nK(x)
h>0
AVERAGES AND SINGULAR INTEGRALS
363
(K(x) is a function on Rn ) to the one-dimensional Hilbert transform
ff(x-t) dt
Hf(x) =
(f a function on R1 ). We will explain how the method of rotations can lead to problem II'.
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 K(Ax.X2y) = 3 K(x,y)
18)
X3
We wish to consider the LP boundedness of the transformation 00
00
Tf(u,v) =
19)
f(u-x,v-y)K(x,y)dx dy .
I
We now introduce parabolic polar coordinates into 19) x = rcos 0 y = r2sinO and find 00
20)
277
Tf(u,v) = J
J
0
0
f(u-rcos 0, v-resin 0)
K(rcos 0, resin 0)r2N(0)drd0
where r2N(0) is the Jacobian factor in the change of variables. N(0) is smooth and N(0+n) = N(0). By 18 we see that
364
STEPHEN WAINGER o0
27r
r
21) Tf(u,v) =
N(O)K(cos d,sin O)dO
0
r C f(u-rcos 0,v-rsin 0) dr 0
2rr
00
-J
N(O)K(cos(O+n),sin(O+n))dd f
0
0
=
i f(u-rcos 0, v-rsin(O)) dr
since K is odd. Thus 217
r
fTf(u,v)
i f(u+rcos O,v+rsin 0) dr
N(O)K(cos 6,s in 0) dd
0
0
since
N(6+n) = N(6) .
Finally 277
r
21A) Tf(u,v) =
00
N(O)K(cost,sinO)dO
r r f(u-rcos6;v-r2sinO)dr
Now adding 21) and 21A) we find that 2n
f
Tf(u,v) = 2
ao
N(O)K(cos O,sin 0)dd f -00
0
i (u-rcos6,v-r2sin6)dr f
If K(cos O,sin 0) is in LT of [0,2n] , we can apply Minkowski's inequality n
IITfIILp < C f de IIHIILp 0
where
AVERAGES AND SINGULAR INTEGRALS
Hef =
f
365
00
f(x-yg(r))
dt
,
-00
with
ye(r) = (rcos O,r2sin O) .
Now we prove IITfIILp
by showing IIHefIILP
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
MEf(x,Y) =
E2
ff
f(x-r,Y-s)
1
drd s
(is)(i+ E2
(") If K(x,y) were dominated by a decreasing, radial, LI function, the classical theory would imply lim MEf(x,y) = f(x,y) E-0
see [SWE]. However K(x,Y) =
1
(1+x2)(1+y2)
a. e.
366
STEPHEN WAINGER
so the smallest radial majorant of K is
1
which is not integrable.
1+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 Cn. 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 B1 and B2 of the same radius, r, intersect, then B2 , the ball having the same
center as B2 but having radius 3r contains B1 . To see how badly this property fails for our problems let us suppose we were considering averages h
1
ff 0
t2+E
f(x-r,y-s) dr ds
t2
over slightly thickened parabolas or balls BE,h = i(r,s)I0
.
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 r ).
We can also see the difficulty of using the Calderon-Zygmund theory to study Problem 11'.
Suppose y(t) is the parabola (t,t2) in R2 and
Tf(x,y) = J f(x-t,y-t2)dt Then we may formally write
Tf(x,y) =
fff(x-t,y-st2)-
= JJ
dsdt
f(x-t,y-v) . S (1- -1 dv dt
0)
Or
22)
Tf=K*f,
where 22A)
K(x,Y)=XS(1--X
),
368
STEPHEN WAINGER
The Calderon-Zygmund theory deals with convolution operators with
kernels K(x,y), but in their theory K(x+h,y+k)- K(x,y) should be much
less than 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 = x2 and (x+h,y+k) is not 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. The Calder6n-Zygmund Theory is based on 4) tools a) The Fourier transform (The Fourier transform is even used in the L1 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 Calder6n-Zygmund decomposition. Some progress in finding covering lemmas for related problems was made by Stromberg [Str] and [STRO] and Cordoba [COR1], [COR2], Cordoba and Fefferman [CF1], [CF2], [CF3], 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], [CS], 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 ,,
r 0
O(t,O)dt .
AVERAGES AND SINGULAR INTEGRALS
369
u is supported on a straight line, namely the x-axis, and 1
g(e'exel)?y)
=
r eietdt 0
which is independent of n and hence cannot decay at infinity along the Y7-axis. Now let us consider a measure supported on a parabola, 00
23)
v(q) =
r
e_t2rb(t,t2)dt
.
_ Then
v(e exei7y) 00
fe_t2eietet2dt. This integral may be computed exactly by completing the square, and it
is easy to see that e2
Ce 1+1771
Thus v(e,q) tends to zero at infinity. Another example is afforded by rotationally invariant Lebesgue measure on the n-1 dimensional sphere Ixl = 1 in Rn. If we denote
this measure by dµ, we have for e c Rn
= CnJ(lei) Ii
2
(v).
370
STEPHEN WAINGER
See [SWE]. Thus
^
_
Idu(4)I < Cn(1 + I.I)
(n
1
2
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.
VAN DER CORPUT'S LEMMA. Let h(t) be a real function. For some j , assume Jh(1)(t)I > A in an interval a < t < b. If j = 1, assume also that h'(t) is monotone, then b
r
exp (ih(t))dt
a
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
24)
du(O)
f
2
(k(t,t2)dt.
1
Then
2
el4telnt dt
du(e,i) = 1
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 t4 rt2 . First we use the fact that h"(t) = rt . Thus by Van Der Corput's lemma with j = 2 , we see
AVERAGES AND SINGULAR INTEGRALS
n
ldu(, ,
25)
C
1
(1 + Inl)I12
Now if Ir1I < igl,
Ih'(t)I = Ie+2,itI > e/8
.
So by Van Der Corput's lemma with j = 1 , 26)
n .I
I
if 161 > 8lnl
Putting 25) and 26) together we have 27)
C
for some 8 > 0. Stein pointed out in retrospect that we can already see from an estimate like 27) that du has interesting properties from the point of harmonic analysis - namely even though du is singular,
Tf=du*f maps LP into L2 continuously for some p < 2. For
371
STEPHEN WAINGER
372
f(Tf)2
=
=
ITf(6,r/)12 dtd77
f
Idu(e,17)12If(e, )I2 de d,1
=f
If( ,ii)I 2
+e 2 +1771)2,5
(1+e 2+I17I If(e,,7)12
(1+e2+1,,127
2q ) /q
J
If(e,rl)I
1 /q
(Ll+e2+'2')
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 e LP where
P +2q=1 II. The Hilbert transform along 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 of as a multiplier transformation Hyf(e) = my(e) f (e )
28)
where
29)
my(e) =
f
Lt
_,p
To see that 29) is true we may either substitute the formula
AVERAGES AND SINGULAR INTEGRALS
f(x)
373
je-e'x f(e)dC
into 13) or recognize the fact that
where D is a distribution
So
and b 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 L2 in 2-dimensions for the curve y(t) = (t, ltlasgnt) ,
a > 0.
So Fabes' proof consisted in showing that the integral
m((,q) =
r
exp(ite+iIt1a(sgnt)rt) Lt
-00
is uniformly bounded in 6 and q. 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
f 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,t2,t3), one would have to understand the zero set of
Real Part
uniformly in 61 , e2 and 63. So it is hard to imagine using the method of steepest descents, and for the curve t, t2, t3, t4, is it would seem close to impossible. Fabes' result was very important in that it gave the first clue that problems such as II and III 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 show that if y(t) = (t,tal,ta2'...'tan-1), 1
(here t 1 can mean either Itlal or Itlalsgnt ) then Hy was bounded on L2(Rn) [SWA]. Here we had to prove the boundedness of the integral 00
r
dt
J 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)
I
dt < C(n) .
E
We shall prove 30) by where C(n) does not depend on induction on n. By changing variables, replacing t by 1/n we
41
375
AVERAGES AND SINGULAR INTEGRALS
may assume en = ± 1 in 30). Then by using Van Der Corput's lemma with j = n , we find t
exp (iels + ... +
31)
ien-
sn-1 +
iSn)
< C(n) .
An integration by parts together with 31) shows that
32)
1 1<JtJ
Now
f exp (ieit+...+ien_ltn-1 +itn) dt
I exp (ieit+... +
ien-ltn-1) Lt
t
<
t
Jr to dt < C(n) . E
E
Hence we have reduced the proof 30) for n to proving 30) for n - 1. Also the case of n = 1 is easy. So we are done by induction. We now turn to the LP theory. We wish to emphasize how curvature and Fourier Transform are joining together to help us. So we shall compare the case of the parabola (t,t2) to the straight line (t,t). In the case of the parabola we are studying 34)
Hpf = DP * f
where DP is a distribution. For a test function 0
376
STEPHEN WAINGER
35)
Dpi
dt
In the case of a straight line we are studying
HLf=DL*f,
36)
where
37)
DL95 = J o(t.t) dt
00
38)
Dp(e,rl) = Dp(e1eXei"ly) = F
eieteirlt2 dt
-00
and
39)
DL(e,rl) = DL(e exe"7y)
f00ete'it. -00
t
We can calculate DL explicitly, and we find 40)
DL(e,rl) =c sgn((+n) .
We shall Notice that DL is discontinuous along the line show in contrast that Dp(e,rl) is continuous away from the origin. It is very easy to see that Dp(e,rl) is C°° away from the line i = 0 by complex integration. If for example rl > 0, we think of t as a complex z variable and integrate along the line lmt = Ret. Then the factor eirlt
decays as fast as e-c7ItI2 , and one can easily justify differentiation under the integral sign as long as it > e , for some positive E.
AVERAGES AND SINGULAR INTEGRALS
377
We shall now show that Dp(6,0 is continuous near 161 = 1 . What we must show is that l m0 Dp(4,-q) exists for 6 near ±1 . We shall show
41)
lim ,q-to
00
r
00
J
t
ei6t dt
ei 'tei77t2 dt _ t -00
Assume for simplicity that 71 > 0. Of course 1
1/3 00
77
42)
eiCt dt =
r
lim
r ei4t
J
7,-0
Lt
t
-00 ,17 1 /3
and
I ,71 /3
t
eiet(ei??t2_1) dt
43) 1
.771 /3
1
711r/ 3
< 271
J
tdt < 171/3 - 0
as 17 ->0.
0
In view of 42) and 43) we can show 41) by showing
44)
I
f
1/3
e'eteigt2 dt
o.
378
STEPHEN WAINGER
We shall prove 44) by using Van Der Dorput's lemma. By using Van Der Corput's lemma with j = 2, we see
f
t
eiese"gs
2
ds < c XFII
-2/3
(Let h(s) = es + qs2 , then h"(s) = 277 .) So an integration by parts shows
00
eieteir7t2 d t
45)
00
2/3
<
+
(' dt
1
77
,J
1/2
t2
77
77-
< C 77 2/3171/2
2/3
< C171/6
Note that if t < 77-2 /3 ie+217tl > iei - 2,71/3 So
-2/3 77
e tese l'7s 2 ds
1/3 77
is bounded if 6 is close to minus one. Hence an integration by parts shows 1
13 77
46)
f
00
t
eieteil7t2 dt
1
1/3 77
dt +77 1/3
AVERAGES AND SINGULAR INTEGRALS
379
45) and 46) together with similar estimates for negative t prove 44) and hence the continuity of Dp(e,-9) away from the origin. If one is a little more careful in the above argument, one can prove that Dp(C,.9) satisfies a Lipschitz condition away from the origin. One may then use Riviere's [R] version of Hermander's multiplier theorem [H] to obtain some LP results for p 2 . In fact if y(t) = (t,t2) one obtains 47)
IIHyfIILpCC IIfllLp,
3
For quite a while we tried to prove the range of p, 3 < p < 4 , in 47) was optimal - with no success. Also, there was a suspicion that the use of the Hormander Riviere theorem lost something. In the Hormander argument, one wishes to estimate an expression of the form
f
J=
IK(x+h)-K(x)Idx
R
where K is a kernel, in terms of K. One does this by using Schwartz's inequality and Plancherel's theorem.
f
J=
IxIaIK(x+h)-K(x)ldx I
la
R
<(
f
IXI2a 1
d)' x
IxI2aI(x+h)-K(x)I2dx)
(J
/
R
,_n
<
R
fI1ei2i
K( )12
2
where the sum is over all a'th derivatives of K.
380
STEPHEN WAINGER
Now there was the feeling that a use of Schwartz inequality like that above lost too much, and that more careful estimates for j for particular kernels might lead to better results. To get an idea of what to do we calculated D(e,n) very precisely by the method of steepest descents. We found 2
48)
Dp(6.77) = sgne + C
I77I
e1e2/n
`- 2/ + Better terms ,
where i/i is a Co function on R1 which is one near the origin. Hence,
the crux of the matter was to study the transformation Tf given by 49)
Tf (e,11) = m(e,rl)f((,rl) ,
where
50)
m(,n) =
1
Ir111/2
r)2
1
e
This suggests introducing an analytic family of operators in the sense of [SI] as follows: 51)
T (e,rl) = mz(e,rl)f(e,i)
where
52)
mz(e,rl) =
(171
2)
m(e,v)
Tzf is clearly bounded on L2 if Re z = -1/2 . So if we could prove that the kernel Kz corresponding to Tz for Re z positive satisfied a condition of the Calderon-Zygmund type, we could prove Tz was bounded in each LP if Re z > 0. Hence by Stein's interpolation theorem we would know that To = T was bounded in all LP, p < 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 Tz 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 To. T is essentially HY y = (t,t2),
and so the kernel K0 of To is essentially K0(x,Y)=X8(1-z )
from 22) and 22A). If we introduce "parabolic polar coordinates"
y =resin 0 x =rcos 0 , we see K0(x,y) = a S(9-90) r3
where
sin0 0 cos200
We might expect if Re z = e > 0 Kz to be
E
better than KO. So we
might expect Kz(x,Y) =
1
1
r3
Now we would like to explain why a
19-0011-e
1
10-0011-E
singularity is better than
a S(0-90) singularity. To see the situation more clearly, let us examine the analogous situation for the standard polar coordinates with 00 = 0. Suppose S3)
K0(x,Y) _
O r
where x = rcos 0 and y = rsin 0 , and
382
STEPHEN WAINGER
54)
KE(x.y) =
1
r 21011-E
for ordinary polar coordinates plays the same role as
(The factor
r
for parabolic polar coordinates.) We are trying to see whether
55)
jK(x,y)-K(x-h,y-k)j < C
1
r
.
x2+y2>c(h2+k2)
For either K = KO or K = KE . Let us take h = 0 and k = 1 and consider first K = Ko. Let us look at the contribution from y's which are very close to 0. We have
J J
r20
I-Dr2
)Id9rdr
r>C nearo
If y is very close to 0 161 '" r y r 1
0
y-1 r
So the left-hand side of 55) is at least fC i dr = 00 . So 55) can't
hold. Let us put the matter a little differently. If we consider the 0's with 0 > 0 where the difference K0(r, 0) - Ko(r , 0)
offers no cancellation, we find there is only one bad 0, 0 = 0. But still
AVERAGES AND SINGULAR INTEGRALS
383
n/4
f
JK0(r, 6)- K0(r', O')j dO = 1
.
J0
Let us consider now what happens with
We should expect no
help from the difference K(x,y) - K(x,y-1) when y > 0, if y-1 < 0. But this can only happen if y < 1 or 0 < 1/r . But over this set KE(x,y) is integrable at infinity 00
1 /r
KE(r, 0)d0rdr
1 50
5
1/r
00
Irfderdr
el-F
2
0
5
f1 00
<
,J
rl+e
dr .
5
It is not difficult to complete the argument and to show 55). After a laborious calculation one could prove that the kernel Kz corresponding to Tz of 51) was for Re z > 0 an operator of CalderonZygmund type. This proved that Hy was bounded in LP 1 < p < oo if y = (t,t2). However it would be extremely difficult to carry over this proof to a three dimensional curve. For example it would be hard to derive an analogue of 48) for the curve (t,t2,t3). Essentially one needed a way to define a suitable analytic family Tz without using the asymptotic formula 48). Recall that Hyf = DP * f where
56)
Dp(e,rl) =
reieteint2 dt
384
STEPHEN WAINGER
So one might be tempted to define Hy,f = DP * f
57) w here
e
a irlt2 dt (1+t2)z/2
Dz(e,rl) p
58)
It turns out that 58) is not a good idea for a very important reason. By changing variables in formula 56) we see that Dp(Ac, A271) = Dp(c,rl) , for
any A > 0. Note that also the function mz(e,r/) defined in 52) also has this type of homogeneity, namely mz(Ac,A2r7) = mz((,77), for A > 0.
Now experience has shown that homobeneity is a powerful friend not
to be tossed away lightly. However Dp does not have this homogeneity. This situation can be remedied by defining
Hyf=Dzxf
59)
where 00
60)
Dp(c,rl) =
I
(1
+r72t4)-z /4 Xteir7t2 dt
-00
Note that for A > 0 61)
Dp(Ac, A2rl) = DP(c ,77) .
Let us see how formula 61) can help us. We would like to show 62)
C(z)
if Re z > -2 . By formula 61) we may assume r/ = ±1 , let us say Then by Van Der Carput's lemma with } = 2 , we see that
77
=1.
AVERAGES AND SINGULAR INTEGRALS
t
eXsel1
2
ds
1
ft d ese's2 dsI
So an integration by parts shows 00
J
(1+t4)-z/4eieteirlt2 dt < C(z) t
1
Now 1
j
(1+t4)-z /4 Xt eirlt2 dt
t
-1
f
1
e'et
t2 dt t
-1
1
r t2dtt
J-1
But we already know that
r J
-1
1
dt T
385
STEPHEN WAINGER
386
One may finally deal with -00 < t < -1 in the same manner that one treated 1 < t < 00. Hence 62) is proved. Let us calculate the kernel, Kz , corresponding to Hy if z = E > 0. Then
eiCx e myDp(e,rl)dCdrl
KE(x,Y) =fv
00
f f Lt t
00 +,72t4)-E/4
e-177Y (1
eirlt2
- 00
f
00
00
e-'6(x-t)dx
r
=
J
dt
J -00
00
t
f
i?7(t2-Y)(1+772t4) -E/4 ei 7t2&(x -t)dt
00
00
-.1
ir7(x2-y)(1 + 772x4)-E/4 drl
X
_00
0o
x3 f e
Ir7(x2y) 1
(1+,72)1/4
dr7
-00
1p
x3
x2'-Y2 e/2
(
x2
where PE/2 is a modified Poisson kernel. PE/2 decays exponentially
fast at oo and PE/2(u) ti JuJICE/2 as u -, 0. See
[SWE].
Thus KE(x,y) has a singularity near the curve (t,t2) of the form which is just the improvement over the 8-function that we 1
1e_eo1i-E/2
seek. A modification of these ideas worked for curves
AVERAGES AND SINGULAR INTEGRALS
a' Y(t) = (t
,ta2,...
,
387
tan)
an. See [NRW]. al < However, there is a natural generalization of these curves. All of
these curves satisfy an equation of the form Ye(t) = A Y(t)
63)
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,t2)
A curve satisfying 63), where all the eigenvalues of A have positive real part is called a homogeneous curve. A will generate a group of transformations TX = exp (A loga) . Then
Hyf - Dy*f
64)
where 65)
Dy(Tx) = Dy(e )
Moreover there is a distance pA(x) defined on Rn such that P(TAX) _ AP(x)
In the case of the cure (t,t2) we may, as we said before take
38$
STEPHEN WAINGER 0
A =C1 0
2
)
Then 0 ),
T-
,
0
and p(x,y) = (x4+y2)1 /4
It turns out that in the case of a general homogeneous cure, we can obtain a satisfactory analytic family of operators by defining 66)
Hzf=Dz*f
where 67)
z(e) =pA*(() j ItI-zexp(ie'y(t))dt
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
p * A
y(t) = (tacos (j3logt), tasin(/3logt))
is an example. So one could believe it might be rather messy to prove integrals involving exp to be bounded. It might be difficult to show that at each t some derivative of , y(t) would be non-zero. However, if one makes a change of variables t = eu , we would be led to consideration of integrals involving n(u) where n(u) = y(eu). If A y'(t) = y(t), n(u) satisfies 68)
n'(u) = Ai(u)
.
AVERAGES AND SINGULAR INTEGRALS
389
We shall show that if rj(u) is a curve in Rn satisfying 68) where the
eigenvalues of A have positive real part, then either r)(u) lies in a 0 and u there is a j proper subspace of Rn or for every
1<j
0.
rl(u)
du)
From 68) we see that dI+17
Ajrl
(u)
(u)
du1+1
By the Cayley-Hamilton theorem, we can find numbers aj , 0 < j < n, such that
jAj=0. j=0
So n
Yd j+t
aJ duI+1 '7(u) = 0 j=O and n
j=0
a j
aj
.
rr'(u)
=0.
duJ
In other words rJ,(u) e satisfies an nth order constant coefficient differential equation. So if for some u and d, n'(u) .
=0
j = 1 ,2,...,n-1
du3
rj'(u)
6 = 0 for all u. Thus r)(u) 6 is a constant. But r!(u) 6 , 0
STEPHEN WAINGER
390
as u -, -oo since the eigenvalues of A have positive real part. Hence rt(u)
is in the subspace of Rn orthogonal to e.
We shall conclude this section with the statement of some theorems that follow from the reasoning discussed above.
THEOREM 1. Let y(t) satisfy A Y '(t) -
Y(t)
Suppose the span in Rn of y(t) for positive t and the span in Rn of y(t) for negative t agree. Then IIHyf1I
<
LP
Cp/Y
1
IIf1ILp
We say that a curve y(t) in Rn is well curved if dJy(t)
dyj(t)
j = 1,2,... t= O
span Rn. It turns out that well curved curves can be approximated by homogeneous curves. We can then prove
THEOREM 2. Let y(t) be well curved then,
f f(x-y(t))
dt
<ApyIIfIILp,
1
A general theorem in L2 for curves which are approximately homogeneous was obtained by Weinberg (We].
III. Maximal functions and g-functions We turn now to a discussion of maximal functions. We are especially interested in how the Fourier transforms and g-functions may be used as a tool to relate our maximal functions to more classical ones. The story
AVERAGES AND SINGULAR INTEGRALS
391
began with the study of maximal functions along the curve (t,t2). Thus we wish to consider averages h
70)
r f(x-t,y-t2)dt
Mhf(x,y) --
F
.
0
After much frustration it was decided to take Fourier Transforms and try to see if anything could be learned. It is easy to see that Mhf(e,r1) = mh(e,r))f(e,rl)
where h eit e eit2 'Idt
mh(e,r1) = h 0
Now mh(e,rl) cannot be evaluated explicitly. If one hopes to gain some insight by staring at a formula, one should have a formula that is as explicit as possible. Now a similar situation arose in the path integral approach to Quantum Mechanics. See [FHI. Feynman and Hibbs wished to have an explicit expression for the probability amplitude that a particle lies in a sphere of radius t. In essence, they had to consider an integral in Rn of the form
f exp Q(x)dx I. J
where Q(x) was a quadratic function of x .
(' a
Instead they considered
Ixl2
e J e R
F
+Q(x) dx
STEPHEN WAINGER
392
which could be calculated explicitly. This suggests to consider instead of 70) vh * f(x,y)
71)
f
exp( t2)f(x-t,y-t2)dt h/
.
Then
vh (e,q) = v(he,h2rl) f (C,rl)
72)
where v is the measure considered in 23). In particular 00
v(he,h277) = f e t2eiheteih2r)t2dt
,
and this integral can be computed explicitly by completing the square. We find
73)
v(h,h2) = (nice smoothly decaying function) exp i
2 h477 1 +h 4,772
One might guess that the appearance of the oscillatory factor exp i h42 't is a reflection of the fact that
is a singular measure.
1+h477 2
On the other hand we see that if h2r) is large
expii h27,expie77 which is independent of h. Thus one might try to write (from 72) vh* f (e,n) _ (i(h,h27l)exp(_ l
2
rl //
((exp
i
)f
One might now hope that if one defines a measure vh by the formula
AVERAGES AND SINGULAR INTEGRALS
h(6,r!) = v(he,h2rl) exp - i
393
r7
vh could be dealt with by classical arguments, while
{exp g
(
'I ) f( ,14 =
is another L2 function having the same norm as f. So one
could hope that lsup vh * fllL2 = IISUp vh * gllL2
74)
< CIIgIIL2 = CIIfIIL2
Roughly speaking this works out. See [NRWM].
The proof of 74) was a hint on how to proceed. However, it depended (as had happened before) on very special computations. What was needed was a way to compare averages like vh to more classical averages by using only the decay of vh and not so much the explicit expression of vh as was used above. Stein [Ssp] and [SH] succeeded in doing this by introducting appropriate g-functions. Stein's first argument with g-functions dealt with the averages MB f of equation 4). Recall r
75)
MB r
f(x) = J f(x-y)dgr(y) Br
where Br is the ball of radius r centered at the origin, and dur is the unit rotationally invariant measure on Br. We set, as before, 76)
911f(x) = sup IMB f(x)I r>O
r
STEPHEN WAINGER
394
Stein used g-functions to prove THEOREM 3: Cp,n 110
77)
LP
if p>nn-1 and n>3. Simple examples of the form n-1
IxI O log' IXI where
1
near the origin and has compact support show that p > nn1
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
g(f) (x) _
78)
f
1 /2
00
tI dt Mtf(x)I2 dt
Assume that we could prove IIg(f )IIL2 < C(n)IIf1IL2 '
79)
and let us see how 77) would follow. Now
rnMrf(x)
CT-
snMsf(x)ds
0
=n
r sn-1 Msf(x)ds +
r sn ds Msf(x)ds
AVERAGES AND SINGULAR INTEGRALS
395
Thus r
Mrf(x) < n
r
r sn'1 Msf(x) ds + 0
n
r
sn d Msf(x)ds
0
= I(r) + 11(r)
.
Now l(r) is dominated by the Hardy-Littlewood Maximal function and
II(r) < n
f
sn-1 /2 s1 /2 Msf(x)ds
rJ
0
r
fr -rn J <
1
1 /2
s2n-1 ds
g(f) (x)
0
We turn now to the proof of 79). W
JRn
Ig(f)(x)I2dx= J t J
IdtMtf(x)I2dxdt
Rn
0
0 =
r t Jr
J0
Mtf(6)12d6dt
Rn
00
tJ
_ 0
Rn
Imtr(6)I2dedt
.
STEPHEN WAINGER
396
But Mtf(e) = f(()m(tlel), where m(r) = n12 Jn-2(r) r 2
2
Here Jn_2 is the usual Bessel function. We shall need to know 2
dmr I < C r
and
dorm
=C
1
1
n-1
n-2
dr
2 r1/2
r 2
is bounded. See [SWE]. Thus
f f Jtm(t()Ide 00
f lg(f)(x)I2dA <
t
Rn
0
00
< f If( )12 5 t at m(t 1j1)2 dt Rn
0
Thus to obtain 79) we need to show 00
r tIdm(tIibI2dt
First since
is bounded,
f
1 /IeI
1/161
dt
0
0
Next, since Im'(t)I <
<1e12f
C
n-1 t 2
,
tdt
AVERAGES AND SINGULAR INTEGRALS
397
00
t
dt m(tIeI)I2 dt
/IeI 00
Ifi2 f t
dt < C , (t lel
)n-1
1 /ICI
if n>4. Let us be more precise about the counterexample in 2-dimensions. We take 1
x very near 0
IxIlogI1
in Co
away from 0 .
We can disprove 77) for p = 2, n = 2 by showing 80)
MBIxIf(x) _ 00
for all small x. Because of rotational symmetry it suffices to prove 80) for points (a,0) with a small. In that case
dO
MIxlf(x) _
(a2(1-cos)2+a2sin2ej1 /2 log -n
a2[(l-cos)2+sin2O] E
dO
ti
-E
1
IeI In IeI
= 00 .
398
STEPHEN WAINGER
A similar argument shows sup IMB f(x)I = 00
81)
1
r
a.e. for an appropriate f . We wish now to prove a theorem indicating how bad 81) fails in L2(R2). We set 82)
)Rkf(x,Y) =
sup
f(x)YI
IMB
2k<j<2k+1 2k
where f is in L2(R2). We shall show 83)
Il)KkfpL2 < CkIIfIIL2
To prove 83), it suffices to show that for each function j(x) taking the values 84)
IIMBI+j(x)2-k f(x)II < CkpfIIL2 ,
with C independent of the function j(x). We show 85) by induction on k. That is given a j(x) taking the values we shall define
a function j*(x) so that j*(x) takes values in the set and 85)
IIMBI+j(x)2-kf
_MB1+j*(x)2-k+l
fIIL2 < CIIfIIL2
85) provides the inductive step to prove 86). If j(x) is given define if j(x) is even
if j(x) is odd .
AVERAGES AND SINGULAR INTEGRALS
Then, if we set 2k
gf(x) _ I IMB
86)
J=0
f(x)-MB
k
1+
f(x)I2 1+11 2k
2
we see IMBfI < gf(x)
Thus to prove 86) and hence 85) it suffices to prove 119(fAL2 < CIIfIIL2
87)
We shall prove 87) by using the Fourier Transform. 2k
('
J Igf(x)I2 dx = I 1
fMf(x)-MB
l+2k
0
l+-
2k
=I
IMB
W) - MB
j=0
f where m(r) = J0(r). So to prove 87) it suffices to show
88)
2
399
400
STEPHEN WAINGER
But 88) follows because for s positive and
r
positive
IJ0(r)) < C/'
89)
and
IJ0(r+s)-Jo(r)I < Ct.
90)
To prove 88) we use 89) if ICI > 2k and 90) if 161 < 2k Finally we will show how Stein [SH] proved the maximal function along the parabola (t,t2) is bounded by using g-functions. We start with the measure dµ defined in 24) 2
qS(t,t2)dt
dµ(O) = J
.
1
We set 2
91)
dµh(qS) =
f
-O(ht,ht2) dt
.
I
Then 2
dµh * f(x,y) =
r
f(x-ht,y-ht2)dt
,
1
or
92)
dµh * f(x,y) _
f
2h
f(x-t,y-t2)dt
.
h
We choose a function &i(x,y) E C0m(R2) with %&(0) = 1. We set
AVERAGES AND SINGULAR INTEGRALS
93)
'Ph(x,Y) = h3 0 x , h)
and
94)
g(f)(x,Y) _
Idµh *f(x,Y)-Oh *f(x,Y)I
I
Let us first assume 95)
119(f)IIL2 5 C IIf1IL2
Note that E
sup
f Idµh*f(x,Y)-h*f(x,Y)Idh
E>0
0
sup
1f
E
Idµh *f(x,Y)-oh *f(x,Y)I2dh
E>0 E1/2
0
<
r
1 /2
*f(x,Y)I2 dh
Idµh
0
< g(f) (x,Y) So E
96)
sup I E f dµh *f(x,Y)dh1 t>o 0
<- Of )(x,Y) + sup 10h *f(x,Y)I h>0
401
STEPHEN WAINGER
402
A classical argument (see (RI) shows sup Oh *fp < C tIfIILp Lp h>0
Thus by 95), we see E
sup I e fd1zh*f(xY)dhl
< CIIf11L2
E>0
L2
0
If f>0, E
(' J dµh*f(x,Y)dh 0
2h
E
f
f(x-t,y-t2)dtdh h
0
E
E
>
f r f(x-t,y-t2)
f
dh
/2
0
('E
E >
f
f(x-t,y-t2)dt
J
.
0
So from 96) we infer E
sup If
E>0
eJ
0
It remains to prove 96).
f(x-y,y-t2)dt
L2
AVERAGES AND SINGULAR INTEGRALS
fI(gf)(x,y)I2dxdY
403
r T J Idµh*f-ch*fl2dxdy
=
0
ff 00
rd h(e,-0)h(e,7)I2 If( ,77)I2 d dii
0
f
a* .77)12
f0,1f(
dry.
0
0
So to prove 96) it suffices to prove 00
I h(
J
97)
,rl)I2 < C .
0
The integral on the left side of 97) is
JT
98)
0
Thus by replacing h by Ah A > 0 we may write the expression in 98) as
0
J0
µ(Xhe,)12h271)-0(Ahe,42h277)I2
By choosing A so that \2e2+X4772 = 1 , we see that it suffices to estimate 99) when e 2 +772 = 1 . In this case we see 1
1
(' 99)
JJ
dh dµ(he,h27)-cb(he,h27)I2 < C
0
since dµ(0) = 0(0) = 1 .
r h2 dh < C 0
404
STEPHEN WAINGER
Then from 27) Idµ(he,h27l)I < Ch-3
for some 6>0. So ('00
100)
('00
ld26 < C
T Idµ(he,h2rl)I < J
J
J
h
1
1
CN
Also k1;'"?) <
for any N, so
(1+e2+772)IJ 00
('
101)
1
dh Ii&(he,h277)I < C
.
1
Now we obtain 98) and hence 96) by combining 100), 101) and 102).
In this section we have emphasized L2 methods. LP results for p > 1, can be obtained by combining the L2 estimates presented here with the techniques of section 2. Altogether one can prove the following theorems :
THEOREM 4.
If y(t) satisfies y'= A y(t) where all the eigenvalues of
A have positive real part, h
IIo
THEOREM 5. Let y(t) be a curve in Rn. If the vectors y'(0), y"(0), y(3)(0) ... span R° ,
II
sup
0
f 0
h
If(x-y(t))IdtII
< CpIIfIILP LP
1
AVERAGES AND SINGULAR INTEGRALS
405
IV. Vector field problems We turn now to problems III and 1II'. To study problems III and III' it is convenient to make a change of variables. It is possible to make a change of variables so that the integral curves of the vector field v(x) become lines parallel to the x-axis. Under this change of variables the vectors v(x) transform into curves y which vary from point to point. So we are led to studying problems IV and IV'. We facilitate the statement of these problems with 3 definitions.
If y(x,t) is for each x a smooth curve in t with y(x,0) = 0, let !'1
102)
Hyf(x) =
f(x-y(t,x)) dt
f -1
let !'h
103)
f(x-y(t,x))dt ,
Myf(x) _ 0
and 104)
)11 f(x) = Y
sup Ahf(x)l 1>h>o
y
.
We are now ready for the statement of problem IV and W. Problem IV: When do we have UHyf(x)11Lp <- Cpilf11Lp ?
Problem IV': When do we have f(myf11Lp <- CpIIfhILp ?
The change of variables described above preserves tangency and
curvature conditions. So for example N It=o will be parallel to the x
406
STEPHEN WAINGER
axis, and if the curvature of the integral curves of v(x) never vanishes a2y
will not be zero. It turns out that one can prove the following
at2 t=0 theorems:
THEOREM 6. Let y(t,x) = (t,r(t,x)) be a smooth curve in R2 satisfying
ar
x
It=o = 0
and
a2rlt=ono. at
Then 105)
IIHyf IIL2(R2) <- Of I!L2(R2)
and
106)
IItyfIIL2(R2) <- CIIfllL2(R2)
A consequence of Theorem 6 for vector fields will then be THEOREM 7. If v(x) is a smooth vector field in R2 such that the
integral curves of v have nowhere vanishing curvature, IIHvf1IL2(R2) <_ IIfIIL2(R2)
and h
lim h1 h-00
f(x-tv(x))dt = f(x)
,
a.e.
1 0
for f in L2(R2). These theorems are announced in NSWV. Here we shall give some
discussion of the ideas. Because y(t,x) depends on x, the Fourier Transform no longer seems like a good tool to study Hy and 59 y. We must find a different way to employ orthogonality. Here we're motivated
AVERAGES AND SINGULAR INTEGRALS
407
by work of Kolmogorov and Silveristov [Z] on the partial sums of Fourier Series and an approach to the Poisson Integral by Paley [P]. The idea is denotes composition to consider Hy H* and Mh(x) . (Mh(x))* where
of operators and * signifies Hilbert space adjoint. In order to gain some insight into this method we shall just discuss it in the case y(t) = (t,t2) where, of course, we already know the results by a different method.
Let us consider K = Hy H*. The kernel of K will have support on t(u,v)Iu=t-s,v=t2-s2J. Hence one might hope that K would have a much smoother kernel than Hy H. (Of course if K were bounded on L2 so would Hy.) One might hope then, that K would be a Calderon-Zygmund operator. Let us see if this could be the case. One can see that
H*f(x,y) = J f(x+t,y+y(t)) dt
and hence that
HYH*f(x,y)
f(x+t-s,y+t2-s2) dt
tss
=ff
Let u=s -t, v =s2-t2, and we find HYH*f(x,y)
=ff f(x-u,y-v)k(u,v)dudv
where C
k(u,v) = IuI
u-u u+u
.
Now k(u,v) does have its support spread out. However the singularities of k across the curves v = ±u2 are not locally integrable away from the
408
STEPHEN WAINGER
origin. Hence HYHY cannot be a Calderon-Zygmund operator. Let us try to see what goes wrong on the level of the Fourier Transform. Recall
HYf=Qp*f where
p(e,r/) _ f
e(iet+ir7t2)dt
-00
We know from 48) that
Qp((,ri) = sgn + C
IjII/2 e
77
6
+ better terms.
The multiplier for HY HY would be essentially IQp12 . Notice that IQp12 doesn't look any nicer than Q. However IQp(e,r7)-sgneI2 is
much nicer than Qp(e,rj) or Qp(e,rj) - sgne. Now sgne corresponds to the operator Lf =
ff(x-t,y) dt
L is known to be bounded in L2. This suggests that we try to consider M = (L-HY)(L*-HY) .
If M is bounded in L2 so will be L - HY . Hence so will be HY . This actually works and is the basic idea in the proof of Theorem 4. We turn now to the idea of the proof of Theorem 5. Paley showed that 107)
Ph(x) (Ph(x))*f(x) < CIPh(x)f(x)+P*(x)f(x)]
where Ph is the Poisson Kernel. It follows from 107) that
409
AVERAGES AND SINGULAR INTEGRALS
IIPh(x)f Il L2 < Clip h(x)f(X)ll
L2
and hence
IIPh(x)f1IL2
Since the function h(x) is arbitrary we have Ilsup Ph*f1IL2 <_ CIIf1IL2 h
The fact that we had to modify Hy suggests that we should not expect 107) to hold, but we might expect a variant to hold - perhaps involving an operator
h(x,y) Rh(x,Y)f
(x,Y) =
h(x,y) f
f(x-t,Y)dt
.
0
(We know
IIRh(x,y)f(x,Y)IIL2 <- Cllf(x,Y)lIL2 )
We might hope to prove 108)
Mh(x)(Mh(x))*f(x) < C(Rh(x)Mh(x)f(x)+Mh(x)Rh(x)f(x)+Bh(x)f(x))
where Bh(x) is some bounded operator. Even 107) is not quite right. We refer the reader to NSWV for the correct technical modification of 107). This concludes our discussion of the vector field problem.
V. Recent developments The positive results of Theorem 2 and Theorem 5 assume that the curve y(t) has some curvature at the origin. There have been a number of papers trying to understand what happens when this curvature condition is dropped. See [C], [CNVWW], [NVWW] 1), 2), 2), and [NE]. Let us note that we don't have positive results for all C°° curves.
410
STEPHEN WAINGER
Suppose for example y(t) is odd and y(t) _ (t,r(t)) where 0
for 0
t-1
for t>1
r(t) = Then
n
Hrf(6,71) = my(6,17)f(6,q)
where
r
my(4,i7) =
00
ei4teirir(t) dt
-00
=
r
+
00
1
sintt dt
0
at
1
In fact it is easy to see which is easily seen to be unbounded if the following: Suppose y(t) is odd and linear on a sequence of intervals ai < t < bi < 1 and assume that the linear extension of y on b [ai,bi] does not pass through the origin. Then if the ratios a are t
unbounded,
lim E--0
I
f(x-y(t)) dt
ItI<E
cannot exist in the L2 sense for every f in L2 . It is somewhat more difficult to produce counterexamples for the operator 1R y. For example in the case of the two dimensional curve
(t,r(t)) described above, it is easy to see that Vy is bounded in every LP, because it is dominated by
AVERAGES AND SINGULAR INTEGRALS h
!'h
sup
I
0
411
If(x-t,y)1 dt + sup h
f
f(x-t,y-t+I)dt t
.
0
Both of these operators are bounded by the classical theory. We refer to [SWI for an example to show that the maximal function along a Coo curve has no non-trivial positive results. The above authors have been investigating the behavior of the Hilbert Transform and maximal functions related to convex curves. Let y(t) _
(t, F(t)) a plane curve with r(t) convex for positive t. Let h(t) = tr'(t)- F(t). -h(t) represents the y-intercept of the line tangent to the curve y at y(t). We then have THEOREM 8. Let y(t) = (t,P(t)) with r(t) convex and increasing for t > 0. If r(t) is even IIHyfIIL2 <- CIIflIL2
if and only if
Ir'(ct)I > 21r'(t)I
t>0
for some C>0. If y(t) is odd IIHyfIIL2 <- CNfIIL2
if and only if h(Ct) > 2h(t) ,
t>0
for some C>0. There are generalizations of Theorem 8 to higher dimensions, and an investigation of the LP theory has begun. We know that
412
STEPHEN WAINGER
II
yfliL2 <- CIIfIIL2
if * holds for some C > 0, however the maximal functions can be
bounded on L2 (and in fact in LP for any p > 1) for some convex curves even if * fails. Recently Phong and Stein [PS] introduced a general problem of which
our problems are special cases. They consider at each point P in Rn a submanifold Mp of dimension say a and an a dimensional CalderonZygmund kernel K(P,Q). Then they consider
Tf(P) T (Q) K(P,Q)dm(Q)
where dm(Q) is a measure on Mp M. They show that if n > 3 and
k = n-1 IITfIILp <_ CIIfIILP
if Mp satisfies a kind of generalized curvature condition. See [PSI. Various authors have also considered multiple parameter problems which are essentially multiple Hilbert transforms on surfaces and multiparameter maximal functions on surfaces. See [NW2], [V], [STR1], and [CSS].
Appendix 1. An introduction to the method of steepest descents. Here we shall try to give an explanation of the main ideas of the method of steepest descent. The interested reader can find a more detailed description in [B]. Let us first consider the behavior of the integral 00
A-1)
1(I) = f e ax2 dx z
413
AVERAGES AND SINGULAR INTEGRALS
for large A. Of course we can make a change of variables
and observe
A-2)
I(A) _ V"X
where 00 e-t2
B=
A-3)
dt .
-00
The point we wish to make here is that if A is large most of the contribution to the integral I(A) comes from a small neighborhood of the origin. In fact
5e
At2 dt
<
e
f
('
At2
A__t2
dt + J e 2
2 dt
t>1
A21/5
It1,A21/5
<e-AI/5+e,\/2 -gyp very fast as A. o . Thus the main contribution to the integral I(A) comes from the small interval - 1 < t < 1 . Now if we perturb the integrand in I(A), we
A2/5 - -A2/5
can expand the integrand in a power series in that little interval. For example we might consider 00
JA) =
J
e Xh(t)dt
STEPHEN WAINGER
414
where h(t) > t2 for large t, h(t) = t2 +0(t3) for small t and h(t) > 0 for t > 0. Then one can easily see that
f
dt
Itl>A2/5
is exponentially small as before. Now e Ah(t)dt
J
= f e-At2 . (1+O(at3) Itl«
ItIU 2/5
00
e-at2 dt
e-'fit2
dt + Ea I /5
ltl«2/5 =
B + O(e-AI
f
-00
/5)+0
(1
l
.
In the method of steepest descents we try to choose a contour of integration so that on the new contour the situation would be essentially that of J. Thus for example, if we had
K=
r
eikt2+.k P(t)dt
-00
and P(t) were very negative at infinity and 0(t3) near t = 0 we would try to write t = a + it and integrate on the line a = r for It l < 1/, 2 /5 , and we would expand in a power series in this small interval.
AVERAGES AND SINGULAR INTEGRALS
415
More, generally, if we were concerned with the asymptotic behavior for large values of k of an integral of the form fg(z)eAh(z)dz
g(z) and h(z) are holomorphic, we would try to choose a contour on which Re h(z) had only a finite number of maxima, and argue that the main contribution to the integral should come from a small neighborhood of the largest maximum or perhaps an endpoint and we then expand g and
h in a Taylor series at such points. At such a maximum, e, h'(C) = 0. Thus the main contribution to our integral should come from a point where
h'(C) = 0 or an endpoint. This is also reflected in Van Der Corput's lemma. A variant of this principle for non-analytic functions is called the principle of stationary phase and is discussed in Professor Stein's lectures in these proceedings. Appendix 2. The method of stationary phase and quantum mechanics
The method of stationary phase lends itself to a formulation of quantum mechanics that is very appealing to at least some mathematicians. The principle of stationary phase asserts that the integral b
I=
r eiAf(t)dt a
with f(t) real gets most of its contributions for large A, near a, b, or a zero of f'(t). If we had no endpoints for example if f were periodic with period b - a or if the interval of integration was from -oc to oo and f oscillated very rapidly for large t , we would expect the main contribution to the integral to come from small neighborhoods of a zero of V.
416
STEPHEN WAINGER
Let us now turn to quantum mechanics. In particular let us consider a particle moving from a point xa to xb as time evolves from time ta to time tb. According to classical physics, the particle will follow a path for which the classical action is stationary. If x(t) is any path, and our particle has mass m and is moving under the influence of a potential V(x,t), the classical Lagrangian is defined by L = 2 [x(t)]2 - V(x(t),t) . The action along a path x
is defined by
tb
S(x) =
r
L(x(t))dt .
ta
The path on which the classical particle moves will be a path x(t) such that z(ta) = xa , z(tb) = xb , and such that S(x +Sx)IS=o
=
0.
One of the most important principles of quantum mechanics asserts that motion on a classical scale must be essentially described by the laws of classical mechanics. In our example it means the only paths that are im-
portant are paths near the path x of *. This principle is expressed in terms of a small number h. The principle says that the only classical paths that should be important are paths which differ from x only on an h scale of measuring. Now the Feynman path integral formulation is in terms of probability amplitudes of events. In our case the probability amplitude of passing from xa at time ta to xb at time tb is given by an integral
Fje"h
s(x)
Dx
AVERAGES AND SINGULAR INTEGRALS
417
where the integration is an integral over all paths x(t) such that x(ta) = xa and x(tb) = xb. Leaving aside the question of how such an integral can be precisely defined, let us try to guess what paths contribute the most to the integral F . Since h is very small, the principle of stationary phase would indicate that the main contribution to the integral F should come from a small neighborhood of a path of which some kind
of a derivative of S was zero. Thus we might expect the main contribution to the integral F to come from paths which are very close (on a
scale of h ) to the path x defined by *. There have been many papers in the mathematical and physical literature dealing with the problem of making sense out of the definition F. See for example [CS] and references cited there. However, the original definition in [FH] serves the purpose of making many formal calculations. We may imagine dividing the t interval into 21 subintervals, Ij , of equal length [b_ta]k
Ik =
l
+ ta, [tb-ta]2
nl
+ to
.
We consider only paths which are linear on Ik . Such paths are determined by xk = X (ta
[tb-ta])
+ 2
k=
We then take
1 S(x1 ... x2j-1)
F = lim A. j-,°°
J
eh
dx 1dx2,...,dxk
where S(x1,...,x2j_1) is the action along the polygonal path determined by x1'...,x2j_1 , and Aj is a normalizing factor. For more details see [FH]. We would like to make one final remark about the book [FH]. It's
418
STEPHEN WAINGER
great. It explains quantum mechanics in terms of mechanics and does not use notions of atomic physics as many of the standard books do. Thus it is accessible to many more mathematicians than standard quantum
mechanics texts. The book also elucidates the differences between the nature of physicists and mathematicians. If you don't want to know h to 3 significant figures in ergs/sec, whatever they are, you probably would rather be a mathematician than a physicist. Finally, many mathematicians could probably learn a great deal to improve themselves as mathematicians by reading the book. STEPHEN WAINGER DEPARTMENT OF MATHEMATICS UNIVERSITY OF WISCONSIN MADISON, WISCONSIN
REFERENCES [B]
N. DeBruijn, Asymptotic Methods in Analysis, North Holland Publishing Co., Amsterdam, 1958.
[BF]
H. Busemann and W. Feller, "Zur Differentiation des Lebesguesche Integrale," Fund. Math, Vol. 22, 1934,
[CS]
[CSS]
[CW]
[C] [CS]
[CORI]
pp. 226-256. R. Cameron and D. Storvick, A simple definition of the Feynman integral with applications, Amer. Math. Soc., Providence, 1983.
H. Carlsson, P. Sjogren, and J. Stromberg, "Multiparameter maximal functions along dilation-invariant hypersurfaces" to appear in Trans. of the A.M.S. H. Carlsson and S. Wainger, "Maximal functions related to convex polygonal lines," to appear. M. Christ, preprint. J. L. Clerc and E. M. Stein, "LP multipliers for non-compact symmetric spaces," Proc. Nat. Acad. Sci., U.S.A., Vol. 71, 1974, pp. 3911-3912. A. Cordoba, "The Kekeya maximal function and the spherical summation multipliers," Amer. J. of Math., Vol. 99, 1977, p. 1-22.
[COR2]
"Maximal functions, covering lemmas and Fourier multipliers," Proc. Symp. in Pure Math., Vol. XXXV, Part I, 1979, pp. 29-50.
AVERAGES AND SINGULAR INTEGRALS
[CF1] [CF2]
[CF3]
419
A. Cordoba and R. Fefferman, "A geometric proof of the strong maximal theorem," Annals of Math., Vol. 102, 1975, pp. 95-100. , "On differentiation of integrals," Proc. Nat. Acad. Sci., U.S.A., Vol. 74, 1977, pp. 2211-2213. "On the equivalence between the boundedness of certain classes of maximal and multiplier operators in Fourier Analysis," Proc. Nat. Acad. Sci., U.S.A., Vol. 74, 1977, pp. 423-425.
[CNVWW] A. Cordoba, A. Nagel, J. Vance, S. Wainger, and D. Weinberg,
"LP bounds for Hilbert Transforms along convex curves," preprint. [F]
E. B. Fabes, "Singular integrals and partial differential equations of parabolic type," Studia Math., Vol. 28, 1966, pp. 81-131.
[FEF]
[FH] [H] [KS]
[NRW] [NRWM]
R. Fefferman, "Covering lemmas, maximal functions, and multiplier operators in Fourier Analysis," Proc. Symp. in Pure Math., Vol. XXXV, Part 1, pp. 51-60. R. Feynman and A. Hibbs, Quantum Mechanics and Path Integrals, McGraw Hill, New York. L. HOrmander, "Estimates for translation invariant operators in LP spaces," Acta Math., Vol. 104, 1960, pp. 93-139. R. Kunze and E. Stein, "Uniformly bounded representations and harmonic analysis of the 2x2 unimodular group," Amer. J. of Math., Vol. 82, 1960, pp. 1-62. A. Nagel, N. Riviere, and S. Wainger, "On Hilbert transforms along curves, It, "Amer. J. Math., Vol. 98, 1976, pp. 395-403. A. Nagel, N. Riviere and S. Wainger, "A maximal function
associated to the curve (t,t2)," Proc. Nat. Acad. of Sci., U.S.A., Vol. 73, 1976, pp. 1416-1417. [NSWB]
A. Nagel, E. Stein, and S. Wainger, "Balls and metrics defined by vector fields I; Basic Properties," to appear in Acta Mathematica.
A. Nagel, E. Stein, and S. Wainger, "Boundary behavior of functions holomorphic in domains of finite type," Proc. Nat. Acad. Sci. U.S.A., Vol. 78, 1981, pp. 6595-6599. [NSWV] A. Nagel, E. Stein, S. Wainger, "Hilbert transforms and maximal functions related to variable curves," Proc. of Symposia in Pure Math., Vol. XXXV, part 1, 1979, pp. 95-98. [NVWW1] A. Nagel, J. Vance, S. Wainger, and D. Weinberg, "Hilbert transforms for convex curves," Duke Math. J., Vol. 50, 1983, [NSW]
pp. 735-744.
420
STEPHEN WAINGER
[NVWW2] A. Nagel, J. Vance, S. Wainger, and D. Weinberg, "The Hilbert transform for convex curves in Rn," to appear in Amer. J. of Math. [NVWW3] , "Maximal functions for convex curves," Preprint.
-
[NW]
[NW2]
[NE]
[P] [R]
[PS]
[S]
[SBC]
[SHI
[Ssp]
[SPL] [SI] [SWA]
A. Nagel and S. Wainger, "Hilbert transforms associated with plane curves," Trans. Amer. Math. Soc., Vol. 223, 1976, pp. 235-252. -, "L2 boundedness of Hilbert transforms along surfaces and convolution operators homogeneous with respect to a multi-parameter group," Amer. J. of Math., Vol. 99, 1977, pp. 761 785. W. Nestlerode, "Singular integrals and maximal functions associated with highly monotone curves," Trans. Amer. Math. Soc., Vol. 267, 1981, pp. 435-444. R. Paley, "A proof of a theorem on averages," Proc. Lond. Math. Soc., Vol. 31, 1930, pp. 289-300. N. Riviere, "Singular integrals and multiplier operators," Ark. Mat., Vol. 9, 1971, pp. 243-278. D. Phong and E. Stein, "Singular integrals related to the Radon transform and boundary value problems," Proc. Nat. Acad. Sci., U.S.A., Vol. 80, 1983, pp. 7697-7701. E. Stein, Singular Integrals and Differentiability Properties of Functions, Princeton University Press, 1970. , Boundary Behaviour of Holomorphic Functions of Several Complex Variables, Princeton University Press, Princeton, 1972. , "Maximal functions: Homogeneous curves," Proc. Nat. Acad. Sci., U.S.A., Vol. 73, 1976, pp. 2176-2177. , "Maximal functions: Spherical means," Proc. Nat. Acad. Sci., U.S.A., Vol. 73, 1976, pp. 2174-2175. , Topics in Harmonic Analysis related to the LittlewoodPaley Theory, Princeton University Press, Princeton, 1970. , "Interpolation of linear operators," Trans. Amer. Math. Soc., Vol. 88, 1958, pp. 359-376. E. Stein and S. Wainger, "The estimation of an integral arising in multiplier transformations," Studia Math., Vol. 35, 1970, pp. 101-104.
[SW]
, "Problems in harmonic analysis related to curvature," Bulletin of the A.M.S., Vol. 84, 1978, pp. 1239-1295.
AVERAGES AND SINGULAR INTEGRALS [SWE]
[STR1] [STR]
[STRO]
[V]
421
E. Stein and G. Weiss, Introduction to Fourier Analysis on Euclidean Spaces, Princeton University Press, Princeton. R. Strichartz, "Singular integrals supported on submanifolds," Studia Math., Vol. 74, 1982, pp. 137-151. J. Stromberg, "Weak estimates on maximal functions with rectangles in certain directions," Ark. Mat., Vol. 15, 1977, pp. 229-240. J. Stromberg, "Maximal functions associated to rectangles with uniformly distributed directions," Ann. of Math., Vol. 107, 1978, pp. 399-402. J. Vance, "LP boundedness of the multiple Hilbert transform along a surface," Pacific J. of Math., Vol. 108, 1983, pp. 221-241.
[WE]
[Z]
D. Weinberg, "The Hilbert transform and maximal function for approximately homogeneous curves," Trans. Amer. Math. Soc., Vol. 267, 1981, pp. 295-306. A. Zygmund, Trigonometric Series, Vols. I & II, Cambridge University Press, London, 1959.
INDEX
approach regions admissible, 245 non-isotropic, 261 non-tangential, 244
Ap classes, 73, 353 area integral, 92 atomic decomposition, 114, 156, 159, 340
convex curves, 411 curvature, 321, 361
and Fourier transform, 321, 325, 268, 375 DeGiori-Nash regularity theory, 144, 158
Dirichlet problem, 132, 243 domain of holomorphy, 211
Bergman kernel, 230
duality of H1 and BMO, 114, 340
B.M.O. (Bounded mean oscillation), 9, 94, 331, 340 BMO(R+xR+), 101
electrostatics, 163
Bochner-Martinelli formula, 196
Bochner-Riesz summability, 344 Calder6n-Zygmund decomposition, 48
Campbell-Hausdorff formula, 267
exponential mapping, 267, 275
Fatou's theorem, 245 finite type, 269, 283, 324 Fornaess imbedding theorem, 226
functional calculus, 27
canonical coordinates, 275 Carleson measure, 16, 94 Cauchy-Fantappie formula, 195 Cauchy integral, 5, 8, 186 on a Lipschitz curve, 143 Cauchy-Riemann equation, 201 convergenge of averages over spheres, 358, 361 along curves, 361 along vector fields, 361, 406 covering lemmas, 60
g-functions, 393
Hardy space, 144
Hp spaces, 89 HP(R+ x R+)
,
101
harmonic measure, 141
Hartogs extension phenomenon, 207
heat operator, 254 Henkin integral formula 219, 336
423
424
INDEX
Heisenoerg group, 257, 335 Hilbert transforms along curves, 361, 372 along vector fields, 362,406
hydrostatics, 177 hypoelliptic differential operators, 281 Kohn (canonical) solution, 209 Laplacian, 266
non-commuting vector fields, 267
oscillatory integrals (first kind),308 (second kind), 335, 344, 349 Poisson integral, 187, 243 bi-Poisson integral, 102 Rellich-type formulas, 150, 166
restriction theorems, 325, 344
Korn-type inequalities, 145 Korteweg-de Vries equation, 25
Laplace equation, 132, 242 Leray form, 194 Levi-pseudoconvex, 214, 257 polynomial, 227 Lipschitz domain, 133, 145 Littlewood-Paley-Stein theory, 48, 53 local singular function, 210 Lu Qi-Keng conjecture, 235 maximal functions, 48, 49, 245 strong, 60 spherical, 359, 393 on curves, 391, 400 on vector fields, 406 method of layer potentials, 133, 143 Mobius transformation, 186
multilinear Fourier analysis, 18 multiparameter differentiation theory, 57 multipliers, 72
Sobolev estimates, 145 space of holomorphy, 42
space of homogeneous type, 251 stationary phase and quantum mechanics, 415
Stein-Weiss spaces, 96, 160 steepest descents, 412 Stokes theorem, 189 surfaces (non-zero curvature), 321 systems of elliptic equations, 133, 163
Szego kernel, 193, 265, 296, 297 Transference theorem, 39 Van der Corput lemmas, 309, 370
weight norm inequalities, 72 Zygmund conjecture, 67
Library of Congress Cataloging-in-Publication Data Beijing lectures in harmonic analysis.
(Annals of mathematics studies ; no. 112) Bibliography: p. Includes index.
1. Harmonic analysis. 1931- . H. Series. QA403.B34 1986 ISBN 0-691-08418-1
I. Stein, Elias M.,
515'.2433
86-91452
ISBN 0-691-08419-X (pbk.)
Elias M. Stein is Professor of Mathematics at Princeton University