• Tidak ada hasil yang ditemukan

Sublevel Set Estimates and the Van der Corput Lemma

Dalam dokumen Graduate Texts in Mathematics (Halaman 164-173)

2.6 Oscillatory Integrals

2.6.2 Sublevel Set Estimates and the Van der Corput Lemma

We discuss a sharp decay estimate for one-dimensional oscillatory integrals. This estimate is obtained as a consequence of delicate size estimates for the Lebesgue measures of the sublevel sets{|u| ≤α}for a functionu. In what follows,u(k)de-

notes thekth derivative of a function u(t)defined on R, andCk the space of all functions whosekth derivative exists and is continuous.

Lemma 2.6.5.Let k≥1and suppose that a0, . . . ,akare distinct real numbers. Let a=min(aj)and b=max(aj)and let f be a real-valuedCk−1function on[a,b]that isCkon(a,b). Then there exists a point y in(a,b)such that

k m=0

cmf(am) =f(k)(y),

where cm= (−1)kk! ∏k

`6=m`=0

(a`−am)−1.

Proof. Suppose we could find a polynomialpk(x) =∑kj=0bjxjsuch that the function ϕ(x) = f(x)−pk(x) (2.6.8) satisfiesϕ(am) =0 for all 0≤m≤k. Since theaj are distinct, we apply Rolle’s theoremktimes to find a pointyin(a,b)such that f(k)(y) =k!bk.

The existence of a polynomialpksuch that (2.6.8) is satisfied is equivalent to the existence of a solution to the matrix equation

ak0 ak−10 . . . a0 1 ak1 ak−11 . . . a1 1 ... ... ... ... ... akk−1ak−1k−1. . . ak−11 akk ak−1k . . . ak 1

 bk bk−1

... b1 b0

=

 f(a0) f(a1)

... f(ak−1)

f(ak)

 .

The determinant of the square matrix on the left is called theVandermonde determi- nantand is equal to

k−1

`=0 k j=`+1

(a`−aj)6=0.

Since theaj are distinct, it follows that the system has a unique solution. Using Cramer’s rule, we solve this system to obtain

bk =

k m=0

(−1)mf(am)

k−1

`6=m`=0 k

j=`+1 j6=m

(a`−aj)

k−1

`=0 k

j=`+1

(a`−aj)

=

k

m=0

(−1)mf(am)

k

`6=m`=0

(a`−am)−1(−1)k−m.

The required conclusion now follows withcmas claimed.

Lemma 2.6.6.Let E be a measurable subset ofRwith finite nonzero Lebesgue mea- sure and let k∈Z+. Then there exist a0, . . . ,akin E such that for all`=0,1, . . . ,k we have

k

j=0 j6=`

|aj−a`| ≥(|E|/2e)k. (2.6.9)

Proof. Given a measurable setE with finite measure, pick a compact subsetE0of E such that|E\E0|<δ, for someδ >0. Forx∈RdefineT(x) =|(−∞,x)∩E0|.

ThenT enjoys the distance-decreasing property

|T(x)−T(y)| ≤ |x−y|

for allx,y∈E0; consequently, by the intermediate value theorem,T is a surjective map from E0 to[0,|E0|]. Letaj be points in E0 such that T(aj) = kj|E0| for j= 0,1, . . . ,k. Forkan even integer, we have

k

j=0 j6=`

|aj−a`| ≥

k

j=0 j6=`

j k|E0| −`

k|E0| ≥

k

j=0 j6=k2

j k−1

2 |E0|k=

k 2−1 r=0

r−k2 k

2

|E0|k,

and it is easily shown that (k/2)!2

k−k≥(2e)−k. Forkan odd integer we have

k

j=0 j6=`

|aj−a`| ≥

k

j=0 j6=`

j k|E0| −`

k|E0| ≥

k

j=0 j6=k+12

j k−k+1

2k |E0|k,

while the last product is at least n1

k·2 k· · ·

k−1 2

k

o2k+1

2k ≥(2e)−k.

We have therefore proved (2.6.9) withE0replacingE. Since|E\E0|<δ andδ>0 is arbitrarily small, the required conclusion follows.

The following is the main result of this section.

Proposition 2.6.7.(a) Let u be a real-valued Ck function, k∈Z+, that satisfies u(k)(t)≥1for all t∈R. Then the following estimate is valid for allα>0:

t∈R: |u(t)| ≤α ≤(2e)((k+1)!)1kα

1

k. (2.6.10)

(b) For all k≥2, for every real-valued Ck function u on the line that satisfies u(k)(t)≥1, for any−∞<a<b<∞, and everyλ>0, the following is valid:

Z b a

eu(t)dt

≤12k|λ|1k. (2.6.11) (c) If k=1, u0(t)is monotonic on(a,b), and u0(t)≥1for all t∈(a,b), then

Z b a

eu(t)dt

≤3|λ|−1. (2.6.12)

Proof. Part (a): LetE ={t ∈R: |u(t)| ≤α}. If|E| is nonzero, then by Lemma 2.6.6 there exista0,a1, . . . ,akinEsuch that for all`we have

|E|k≤(2e)k

k

j=0 j6=`

|aj−a`|. (2.6.13)

Lemma 2.6.5 implies that there existsy∈ minaj,maxaj

such that u(k)(y) = (−1)kk!

k

m=0

u(am)

k

`6=m`=0

(a`−am)−1. (2.6.14)

Using (2.6.13), we obtain that the expression on the right in (2.6.14) is in absolute value at most

(k+1)! max

0≤j≤k|u(aj)|(2e)k|E|−k≤(k+1)!α(2e)k|E|−k, sinceaj∈E. The boundu(k)(t)≥1 now implies

|E|k≤(k+1)!(2e)kα as claimed. This proves (2.6.10).

Part (b): We now takek≥2 and we split the interval(a,b)in (2.6.11) into the sets

R1 = {t∈(a,b): |u0(t)| ≤β}, R2 = {t∈(a,b): |u0(t)|>β},

for some parameter β to be chosen momentarily. The function v=u0 satisfies v(k−1)≥1 andk−1≥1. It follows from part (a) that

Z

R1

eiλu(t)dt

≤ |R1| ≤2e(k!)k−11 β

1

k−1 ≤6kβk−11 .

To obtain the corresponding estimate overR2, we note that ifu(k)≥1, then the set {|u0|>β}is the union of at most 2k−2 intervals on each of whichu0is monotone.

Let(c,d)be one of these intervals on whichu0is monotone. Thenu0has a fixed sign

on(c,d)and we have

Z d c

eu(t)dt

=

Z d c

eu(t)0 1 λu0(t)dt

Z d c

eiλu(t) 1 λu0(t)

0

dt

+ 1

|λ|

eu(d)

u0(d) −eiλu(c) u0(c)

≤ 1

|λ| Z d

c

1 u0(t)

0 dt+ 2

|λ|β

= 1

|λ|

Z d c

1 u0(t)

0

dt

+ 2

|λ|β

≤ 1

|λ|

1 u0(d)− 1

u0(c) + 2

|λ|β ≤ 3

|λ|β,

where we use the monotonicity of 1/u0(t)in moving the absolute value from inside the integral to outside. It follows that

Z

R2

eiλu(t)dt

≤ 6k

|λ|β.

Choosingβ =|λ|−(k−1)/k to optimize and adding the corresponding estimates for R1andR2, we deduce the claimed estimate (2.6.11).

Part (c): Repeat the argument in part (b) settingβ =1 and replacing the interval

(c,d)by(a,b).

Corollary 2.6.8.Let(a,b), u(t),λ >0, and k be as in Proposition 2.6.7. Then for any functionψon(a,b)with an integrable derivative and k≥2, we have

Z b a

eu(t)ψ(t)dt

≤12kλ−1/k

|ψ(b)|+ Z b

a

0(s)|ds

.

We also have

Z b a

eu(t)ψ(t)dt

≤3λ−1

|ψ(b)|+ Z b

a

0(s)|ds

,

when k=1and u0is monotonic on(a,b).

Proof. Set

F(x) = Z x

a

eu(t)dt and use integration by parts to write

Z b a

eiλu(t)ψ(t)dt=F(b)ψ(b)− Z b

a

F(t)ψ0(t)dt.

The conclusion easily follows.

Example 2.6.9.TheBessel functionof ordermis defined as Jm(r) = 1

2π Z

0

eirsinθe−imθdθ.

Here we take bothrandmto be real numbers, and we suppose thatm>−12; we refer to Appendix B for an introduction to Bessel functions and their basic properties.

We use Corollary 2.6.8 to calculate the decay of the Bessel functionJm(r)as r→∞. Set

ϕ(θ) =sin(θ)

and note thatϕ0(θ)vanishes only atθ=π/2 and 3π/2 inside the interval[0,2π]and

thatϕ00(π/2) =−1, whileϕ00(3π/2) =1.We now write 1=ψ123, where

ψ1is smooth and compactly supported in a small neighborhood ofπ/2, andψ2is smooth and compactly supported in a small neighborhood of 3π/2. For j=1,2, Corollary 2.6.8 yields

Z 0

eirsin(θ) ψj(θ)e−imθ

≤C m r−1/2

for some constantC, while the corresponding integral containingψ3has arbitrary decay inrin view of estimate (2.6.3) (or Proposition 2.6.4 whenn=1).

Exercises

2.6.1.Suppose thatuis aCkfunction on the line that satisfies|u(k)(t)| ≥c0>0 for somek≥2 and allt∈(a,b). Prove that forλ >0 we have

Z b a

eu(t)dt

≤12k(λc0)−1/k

and that the same conclusion is valid whenk=1, providedu0is monotonic.

2.6.2.Show that ifu0 is not monotonic in part (c) of Proposition 2.6.7, then the conclusion may fail.

Hint:Letϕ(t)be a smooth function on the real line that is equal to 10ton intervals [2πk+ε,2π(k+1

2)−ε]and equal tot on intervals[2π(k+1

2) +ε,2π(k+1)−ε].

Show that the imaginary part of the oscillatory integral in question may tend to infinity over the union of several such intervals.

2.6.3.Prove that the dependence onkof the constant in part (b) of Proposition 2.6.7 is indeed linear.

Hint:Takeu(t) =tk/k! over the interval(0,k!).

2.6.4.Follow the steps below to give an alternative proof of part (b) of Proposition 2.6.7. Assume that the statement is known for somek≥2 and some constantC(k)

for all intervals[a,b]and allCkfunctions satisfyingu(k)≥1 on[a,b]. Letcbe the unique point at which the functionu(k)attains its minimum in[a,b].

(a) Ifu(k)(c) =0, then for allδ >0 we haveu(k)(t)≥δ in the complement of the interval(c−δ,c+δ)and derive the bound

Z b a

eu(t)dt

≤2C(k)(λ δ)−1/k+2δ. (b) Ifu(k)(c)6=0, then we must havec∈ {a,b}. Obtain the bound

Z b a

eiλu(t)dt

≤C(k)(λ δ)−1/k+δ.

(c) Choose a suitableδ to optimize and deduce the validity of the statement fork+1 withC(k+1) =2C(k) +2=5·2k−2. (Note thatC(1) =3.)

2.6.5.(a) Prove that for some constantCand allλ∈Randε∈(0,1)we have

Z

ε≤|t|≤1eiλtdt t

≤C.

(b) Prove that for someC0<∞, allλ∈R,k>0, andε∈(0,1)we have

Z

ε≤|t|≤1eiλt±tkdt t

≤C0.

(c) Show that there is a constantC00such that for any 0<ε<N<∞, for allξ12

inR, and for all integersk≥2, we have

Z

ε≤|s|≤Nei(ξ1s+ξ2sk)ds s

≤C00.

Hint:Part (a): For |λ| small use the inequality |eiλt−1| ≤ |λt|. If |λ| is large, split the domains of integration into the regions|t| ≤ |λ|−1and|t| ≥ |λ|−1and use integration by parts in the second case. Part (b): Write

ei(λt+±tk)−1

t =eiλte±itk−1 t +eiλt

t

and use part (a). Part (c): Whenξ12=0 it is trivial. Ifξ2=0,ξ16=0, change variablest=ξ1sand then split the domain of integration into the sets|t| ≤1 and

|t| ≥1. In the interval over the set|t| ≤1 apply part (b) and over the set|t| ≥1 use integration by parts. In the caseξ26=0, change variablest=|ξ2|1/ksand split the domain of integration into the sets|t| ≥1 and|t| ≤1. When|t| ≤1 use part (b) and in the case|t| ≥1 use Corollary 2.6.8, noting thatdk12|dt−1/kt±tk) =k!≥1.

2.6.6.(a) Show that for alla≥1 andλ>0 the following is valid:

Z

|t|≤aλelogtdt

≤6a.

(b) Prove that there is a constantc>0 such that for allb>λ>10 we have

Z b 0

eiλtlogtdt

≤ c

λlogλ.

Hint:Part (b): Consider the intervals(0,δ)and[δ,b)for someδ. Apply Propo- sition 2.6.7 withk=1 on one of these intervals and withk=2 on the other. Then optimize overδ.

2.6.7.Show that there is a constantC<∞such that for all nonintegersγ>1 and allλ,b>1 we have

Z b 0

eiλtγdt

≤ C λγ.

Hint:On the interval (0,δ)apply Proposition 2.6.7 with k= [γ] +1 and on the interval(δ,b)withk= [γ]. Then optimize by choosingδ =λ−1/γ.

HISTORICAL NOTES

The one-dimensional maximal function originated in the work of Hardy and Littlewood [123].

Itsn-dimensional analogue was introduced by Wiener [291], who used Lemma 2.1.5, a variant of the Vitali covering lemma, to derive itsLpboundedness. One may consult the books of de Guzm´an [72], [73] for extensions and other variants of such covering lemmas. The actual covering lemma proved by Vitali [285] says that if a family of closed cubes inRnhas the property that for every pointxARnthere exists a sequence of cubes in the family that tends tox, then it is always possible to extract a sequence of pairwise disjoint cubesEjfrom the family such that

|A\SjEj|=0. We refer to Saks [233] for details and extensions of this theorem.

The classLlogLwas introduced by Zygmund to give a sufficient condition on the local integra- bility of the Hardy–Littlewood maximal operator. The necessity of this condition was observed by Stein [255]. Stein [259] also showed that theLp(Rn)norm of the centered Hardy–Littlewood maxi- mal operatorMis bounded above by some dimension-free constant; see also Stein and Str¨omberg [262]. Analogous results for maximal operators associated with convex bodies are contained in Bourgain [29], Carbery [42], and M¨uller [204]. The situation for the uncentered maximal operator Mis different, since given any 1<p<there existsCp>1 such thatkMkLp(Rn)→Lp(Rn)Cnp(see Exercise 2.1.8 for a value of such a constantCpand also the article of Grafakos and Montgomery- Smith [109] for a larger value). The centered maximal functionMµwith respect to a general inner regular locally finite positive measureµ onRnis bounded onLp(Rn,µ)without the additional hypothesis that the measure is doubling; see Fefferman [93]. The proof of this result requires the following covering lemma, obtained by Besicovitch [23]: Given any family of closed balls whose centers form a bounded subset ofRn, there exists an at most countable subfamily of balls that covers the set of centers and has bounded overlap, i.e., no point inRnbelongs to more than a finite number (depending on the dimension) of the balls in the subfamily. A similar version of this lemma was obtained independently by Morse [202]. See also Ziemer [300] for an alternative formulation.

The uncentered maximal operatorMµof Exercise 2.1.1 may not be weak type(1,1)if the mea- sureµ is nondoubling, as shown by Sj¨ogren [243]; related positive weak type(1,1)results are

contained in the article of Vargas [283]. The precise value of the operator norm of the uncentered Hardy–Littlewood maximal function onLp(R)was shown by Grafakos and Montgomery-Smith [109] to be the unique positive solution of the equation(p1)xppxp−11=0. This con- stant raised to the powernis the operator norm of the strong maximal functionMsonLp(Rn)for 1<p∞. The best weak type(1,1)constant for the centered Hardy–Littlewood maximal opera- tor was shown by Melas [193] to be the largest root of the quadratic equation 12x222x+5=0.

The strong maximal operatorMsis not weak type(1,1), but it satisfies the substitute inequality dMs(f)(α)CRRn

|f(x)|

α (1+log+|f(x)|α )n−1dx. This result is due to Jessen, Marcinkiewicz, and Zygmund [140], but a geometric proof of it was obtained by C´ordoba and Fefferman [58].

The basic facts about the Fourier transform go back to Fourier [95]. The definition of distribu- tions used here is due to Schwartz [235]. For a concise introduction to the theory of distributions we refer to H¨ormander [130] and Yosida [296]. Homogeneous distributions were considered by Riesz [222] in the study of the Cauchy problem in partial differential equations, although some earlier accounts are found in the work of Hadamard. They were later systematically studied by Gelfand and ˇSilov [100], [101]. References on the uncertainty principle include the articles of Fefferman [90] and Folland and Sitaram [94]. The best possible constantBpin the Hausdorff–

Young inequality bf

Lp0(Rn)Bp f

Lp(Rn)when 1p2 was shown by Beckner [16] to be Bp= (p1/p(p0)−1/p0)n/2. This best constant was previously obtained by Babenko [13] in the case whenp0is an even integer.

A nice treatise of the spacesMp,qis found in H¨ormander [129]. This reference also contains Theorem 2.5.6, which is due to him. Theorem 2.5.16 is due to de Leeuw [74], but the proof pre- sented here is taken from Jodeit [142]. De Leeuw’s result in Exercise 2.5.9 says that periodic elements ofMp(Rn)can be isometrically identified with elements ofM(Tn), the latter being the space of all multipliers on`p(Zn).

Parts (b) and (c) of Proposition 2.6.7 are due to van der Corput [282] and are referred to in the literature as van der Corput’s lemma. The refinenment in part (a) was subsequently obtained by Arhipov, Karachuba, and ˇCubarikov [6]. The treatment of these results in the text is based on the article of Carbery, Christ, and Wright [44], which also investigates higher-dimensional analogues of the theory. Precise asymptotics can be obtained for a variety of oscillatory integrals via the method of stationary phase; see H¨ormander [130]. References on oscillatory integrals include the books of Titchmarsh [280], Erd´elyi [83], Zygmund [303], [304], Stein [261], and Sogge [248]. The latter provides a treatment of Fourier integral operators.

Chapter 3

Fourier Analysis on the Torus

Principles of Fourier series go back to ancient times. The attempts of the Pythagorean school to explain musical harmony in terms of whole numbers embrace early ele- ments of a trigonometric nature. The theory of epicycles in theAlmagestof Ptolemy, based on work related to the circles of Appolonius, contains ideas of astronomical periodicities that we would interpret today as harmonic analysis. Early studies of acoustical and optical phenomena, as well as periodic astronomical and geophysical occurrences, provided a stimulus of the physical sciences to the rigorous study of expansions of periodic functions. This study is carefully pursued in this chapter.

The modern theory of Fourier series begins with attempts to solve boundary value problems using trigonometric functions. The work of d’Alembert, Bernoulli, Euler, and Clairaut on the vibrating string led to the belief that it might be possible to rep- resent arbitrary periodic functions as sums of sines and cosines. Fourier announced belief in this possibility in his solution of the problem of heat distribution in spatial bodies (in particular, for the cubeT3) by expanding an arbitrary function of three variables as a triple sine series. Fourier’s approach, although heuristic, was appeal- ing and eventually attracted attention. It was carefully studied and further developed by many scientists, but most notably by Laplace and Dirichlet, who were the first to investigate the validity of the representation of a function in terms of its Fourier series. This is the main topic of study in this chapter.

Dalam dokumen Graduate Texts in Mathematics (Halaman 164-173)