ebook img

Wiener-Wintner for Hilbert Transform PDF

0.26 MB·English
Save to my drive
Quick download
Download
Most books are stored in the elastic cloud where traffic is expensive. For this reason, we have a limit on daily download.

Preview Wiener-Wintner for Hilbert Transform

WIENER-WINTNER FOR HILBERT TRANSFORM MICHAEL LACEY AND ERIN TERWILLEGER 6 0 0 Abstract. We prove the following extension of the Wiener–Wintner Theorem and the 2 Carleson Theorem on pointwise convergence of Fourier series: For all measure preserving n flows (X,µ,T ) and f Lp(X,µ), there is a set X X of probability one, so that for all t f a ∈ ⊂ x X we have J ∈ f dt 9 lim eiθtf(T x) exists for all θ. s↓0Zs<|t|<1/s t t ] The proof is by way of establishing an appropriate oscillation inequality which is itself an A extension of Carleson’s theorem. C . h t a m [ 1. The Main Theorem 1 v We are concerned with quantitative inequalities related to the pointwise convergence of 2 9 singular integrals that are uniform with respect to modulation. To state our results, define 1 dilation and modulation operators by 1 0 (1.1) Dil(p)f(x) d=ef s−1/pf(x/s), 0 < s,p < . 6 s ∞ 0 / Dil(∞)f(x) d=ef f(x/s), 0 < s < . h s ∞ at (1.2) Mod f(x) d=ef eixξf(x), ξ R. ξ m ∈ : v i Let K be a distribution. The most important example will be K (y) d=ef ζ(y)1, where ζ X H y is a smooth, symmetric, compactly supported function. This is a distribution associated to r a a truncation of the Hilbert transform kernel. Our principal concern is the convergence of terms (Dil(1)K) f(x) in a pointwise sense, s ∗ and in one that is, in addition, uniform over all modulations. To do this, we use the following definition. ∞ (1.3) Osc (K; f)2 d=ef sup [(Dil(1) K) (Dil(1) K)] f 2. n 2l/n − 2l′/n ∗ j=1 kj≤l<l′<kj+1 X (cid:12) (cid:12) (cid:12) (cid:12) Research supported in part by the National Science Foundation. The author is a Guggenheim Fellow. 1 2 MICHAEL LACEY ANDERIN TERWILLEGER This definition depends upon a choice of an increasing sequence of integers k Z, a depen- j ∈ dence that we suppress as relevant constants are independent of the choice of k . It also j { } depends upon a choice of positive integer n, which we have incorporated into the notation. This only permits dilations of the form 2l/n for integers l. 1.4. Theorem. Fix a smooth, symmetric, compactly supported function ζ. For integers n > 0 and 1 < p < there is a constant C so that we have the inequality n,p,ζ ∞ (1.5) supOsc (K ; Mod f) C f . n H N p ≤ n,p,ζk kp N (cid:13) (cid:13) The inequality holds for all choices of increasing sequences k : j 1 satisfying k (cid:13) (cid:13) j j+1 { ≥ } ≥ k +n. j Our primary interest in this theorem is the corollary below, which is a Hilbert transform counterpart to the well known Wiener–Wintner theorem for ergodic averages. Deriving the corollary below is a standard part of the literature, with the roots of the argument going back to Calder´on [8]. The use of an oscillation inequality to establish convergence was introduced by Bourgain [7]. Also see the papers of Campbell et al. [10], and Jones et al. [14]. 1.6. Corollary. For all measure preserving flows T : t R on a probability space (X,µ) t { ∈ } and functions f Lp(µ), there is a set X X of probability one, so that for all x X we f f ∈ ⊂ ∈ have dt lim eiθtf(T x) exists for all θ. t s↓0 t Zs<|t|<1/s This is a common extension of two classical theorems: Carleson’s Theorem [11] on Fourier series with Hunt’s extension [13], and the Wiener–Wintner Theorem [22] on ergodic averages. Carleson’s Theorem. We have the inequality dy sup Mod f(x y) . f , 1 < p < . N p N − y p k k ∞ (cid:13) (cid:12)Z (cid:12)(cid:13) Wiener–Wintner(cid:13)Th(cid:12)eorem. Forall measur(cid:12)e(cid:13)preservingflows T : t R on a probability (cid:13) (cid:12) (cid:12)(cid:13) t { ∈ } space (X,µ) and functions f Lp(X,µ), there is a set X X of probability one, so that f ∈ ⊂ for all x X we have f ∈ s lim s−1 eiθtf(T x) dt exists for all θ t s→∞ Z−s s lims−1 eiθtf(T x) dt exists for all θ. t s→0 Z−s The Wiener–Wintner Theorem can been seen as an extension of the Birkhoff Ergodic Theorem. The Carleson Theorem is a deep result from the 60’s, and since then several proofs have been offered. An extensive survey and bibliography on this subject can be found in [15]. WIENER-WINTNER FOR HILBERT TRANSFORM 3 The possibility of extending the Wiener–Wintner Theorem to the setting of the Hilbert transform was first raised in the paper of Campbell and Petersen [9]. The specific result proved there was essentially Carleson’s Theorem on the integers, with a transference to measure preserving systems. Part of this was contained in a prior work of M´at´e [18], a work that was overlooked until much later. Assani [1,2] proved our Corollary 1.6 on different classes of dynamical systems. Indeed, he formulated the concept of a Wiener–Wintner system. In this nomenclature, our corollary states that all measure preserving systems are Wiener–Wintner systems. Our tool to prove convergence in the Hilbert transform setting is the oscillation inequality (1.5), an idea first employed in ergodic theory in the pioneering work of Bourgain on the ergodic theorem along arithmetic sequences [7]. The use of oscillation has subsequently been systematically studied in e.g. [10,14] and in references therein. The main goal of the paper is a proof of Theorem 1.4. Clearly, we follow the lines of a proof of Carleson’s Theorem. In particular we employ the Lacey–Thiele approach [17] and refine one part of it to deduce our main theorem. We will also appeal to the ‘restricted weak type argument’ of C. Muscalu, T. Tao, and C. Thiele [20] and L. Grafakos, T. Tao, and E. Terwilleger [12]. Acknowledgment. The authors have benefited from conversations with Jim Campbell, An- thony Quas and Mate Wierdl. Part of this research was completed at the Schr¨odinger Insti- tute, Vienna Austria. For one of us (ML), discussions with Karl Petersen about this question formed our introduction to Carleson’s theorem, for which we have been indebted to him ever since. 2. Deduction of Theorem 1.4 There are two more technical estimates that we prove. Specifically, let ψ be some Schwartz function which satisfies (2.1) 0 ψ(ξ) C , 0 ≤ ≤ (2.2) ψ is supported in [ 2, 1], b − −2 (2.3) ψ(y) C min( y −ν, y ν). 1 b| | ≤ | | | | Here, ν will be a large constant whose exact value we need not specify. And we will not have complete freedom in precisely which Schwartz function ψ we can take here. It should arise in a particular way described in the proof of Proposition 2.4, and will be nonzero! The purpose of this section is to describe how a particular result for any choice of non zero ψ as above will lead to a proof of our main theorem. 4 MICHAEL LACEY ANDERIN TERWILLEGER Consider the distribution ∞ def (1) Ψ = Dil ψ. 2−v v=1 X We will prove the following two propositions in the next section. 2.4. Proposition. With the assumptions (2.2)—(2.3), the inequality (1.5) holds with n = 1 and the distribution K replaced by Ψ. H 2.5. Proposition. We have the inequality ∞ 1/2 (2.6) sup Dil(1)ψ Mod (f) 2 . f , 1 < p < . N |{ 2j }∗ N | p k kp ∞ (cid:13) hj=X−∞ i (cid:13) (cid:13) (cid:13) (cid:13) (cid:13) Notethatforfixedmodulation,(2.6)isaLittlewoodPaleyinequality, makingtheinequality above a “Carlesonized Littlewood Paley” inequality. Inequalities like this have been proved by Prestini and Sj¨olin [21]. They also follow from the method of Lacey and Thiele. Both propositions follow from our Proposition 3.9 of the next section, which is phrased in a language conducive to the methods of Lacey and Thiele [17]. These methods have been applied in a number of variants of Carleson’s theorem, see e.g. Pramanik and Terwilleger [19] and Grafakos, Tao and Terwilleger [12]. We turn to the deduction of Theorem 1.4. Observe that the two previous propositions immediately prove that when we consider dilations which are powers of 21/n we have supOsc (Ψ;Mod f) . n f , n N, 1 < p < . n N p p k k k k ∈ ∞ N Thus we need not concern ourselves with this feature of Theorem 1.4 and Corollary 1.6. For a distribution K, set K = sup supOsc (K;Mod f) . ∗,p 1 N p k k k k kfkp=1 N Note that since our definition incorporates differences, this is a seminorm on distributions K. That is, it obeys the triangle inequality (which we use), but can be zero for non zero distributions. In particular, for a Dirac point mass δ we have δ = 0, and similarly for ∗,p k k the distribution K with K = 1 . [0,∞) Our task is to show that K < , where K (y) = y−1ζ(y) for some smooth symmet- b H ∗,p H k k ∞ ric, compactly supported Schwartz function. Our Proposition 2.4 is, with this notation, the assertion that Ψ < . The same inequality will hold for a kernel which can be obtained ∗,p k k ∞ as a convex combination of dilations of ψ and Ψ. Thus, set 1 ds def (1) Ψ = Dil Ψ . 0 2s s Z0 WIENER-WINTNER FOR HILBERT TRANSFORM 5 In this integral, we are careful to integrate against the measure ds, which is the Haar measure s for the positive reals under multiplication, the underlying group for the dilation operators. In particular, it follows that Ψ is a distribution whose Fourier transform is a nonzero constant 0 on ( , 1) and is 0 on ( 1, ). Thus by Proposition 2.4, we clearly have Ψ < . −∞ − −2 ∞ k 0k∗,p ∞ Now we will show that D < for the distribution 0 ∗,p k k ∞ D (y) = y−1ζ(y) c(Ψ (y) Ψ (y)), 0 0 0 − − where we choose the complex constant c so that lim D (ξ) = 0. In fact, it is a well known ξ→∞ 0 elementary fact that for c = iπ, b dy (2.7) ζ(y)eiξy = c+O( ξ −1). y | | Z We will decompose the distribution D into a sum which can be treated with Proposition 2.5. 0 Then using that Ψ < and D < , we obtain the desired inequality for K(y) = 0 ∗,p 0 ∗,p k k ∞ k k ∞ y−1ζ(y). Choose χ to be a smooth function supported on 1 ξ 2 so that 2 ≤ | | ≤ ∞ Dil∞ χ = 1 , 2−k R−{0} k=−∞ X and set ∆ = D Dil∞ χ. The following lemma finishes the proof of Theorem 1.4. k 0 2−k 2.8. Lemma. We have b b ∆ . 2−|k|, k Z. k ∗,p k k ∈ Proof. We will verify that (2.9) ∆ . 2−|k|, k Z, k ∞ k k ∈ (2.10) ∆ is supported on 2−k−1 ξ 2−k+1, k b ≤ | | ≤ (2.11) ∆ (y) . 2−k−|k|(1+2−k y )−ν, k Z, y R, k | b| | | ∈ ∈ with implied constants independent of k Z and ν the large, unspecified constant that ∈ appears in (2.3). With decay in k in both (2.9) and (2.11), the lemma then follows from a | | trivial change of scale and from Proposition 2.5. Let us recall the trivial estimate which follows from the symmetry of ζ, eiξy (2.12) K (ξ) = ζ(y) dy . ξ . H | | y | | (cid:12)Z (cid:12) (cid:12) (cid:12) d (cid:12) (cid:12) 6 MICHAEL LACEY ANDERIN TERWILLEGER In addition we have the estimate below, applied for ξ 1 | | ≤ ξ w even (2.13) dw K (ξ) = ζ(y)yw−1eiξy dy . | | dξw H 1 w odd. ( (cid:12) (cid:12) (cid:12)Z (cid:12) (cid:12) (cid:12) (cid:12) (cid:12) Whereas for ξ 1,(cid:12)wedhave (cid:12) (cid:12) (cid:12) | | ≥ (2.14) dw K (ξ) . ξ −ν, ξ > 1, 0 < w ν. dξw H | | | | ≤ (cid:12) (cid:12) That is, we have very ra(cid:12)pid decay(cid:12)in a large number of derivatives. d (cid:12) (cid:12) Now, (2.10) is true by definition of ∆ . To see (2.9) for k 2, note that this is only k ≥ determined by the Fourier transform of K since Ψ and Ψ are zero. The result easily H 0 0 follows by the inequality (2.12) and property (2.10). For k 2, the inequality follows from ≤ the construction of D , and in particular the propertcy in (2.7c). 0 We turn to the last condition, (2.11). It is well known that decay of order ν in spatial variables is implied by differentiability of a function in frequency variables. Observe that yν∆ (y) (ξ) = i−ν dν ∆ (ξ) k dξν k (cid:0) (cid:1) = i−ν dν Dil(∞)χ(ξ)K (ξ). b dξν c 2−k H Hence, d ν yν∆ (y) (ξ) 2kw dν−w sup K (ξ) . k ≤ dξν−w H 2−k−1≤|ξ|≤2−k+1 (cid:12)(cid:0) (cid:1) (cid:12) Xw=0 (cid:12) (cid:12) (cid:12) (cid:12) (cid:12) b (cid:12) (cid:12) d (cid:12) For k 1, this sum is dominated by the last two terms. To control them, use (2.13), ≥ supplying the estimate . 2(ν−1)k. This is better by a factor of 2−k than the trivial estimate, so that Fourier inversion proves (2.11) in this case. The case of k 0 is easier, due to the rapid decay in (2.14). (cid:3) ≤ 3. Decomposition and Main Proposition We state the definitions needed for the main proposition and conclude this section with the argument of how this proposition proves the results of the previous section, namely Proposition 2.4 and Proposition 2.5. Inaddition to the modulation anddilation operatorsin (1.1) and (1.2), we need translation operators (3.1) Tran f(x) d=ef f(x y), y R. y − ∈ WIENER-WINTNER FOR HILBERT TRANSFORM 7 We set to be the dyadic grid and say that I ω is a tile iff ω I = 1. Let D × ∈ D ×D | |·| | T denote the set of all tiles. We think of ω as a frequency interval and I as a spatial interval; our definition of a tile is a reflection of the uncertainty principle for the Fourier transform. We will plot frequency intervals in the vertical direction. Each dyadic interval ω is a union of two dyadic intervals of half the length of ω. We call them ω and view ω as above ω . ± + − We take a fixed Schwartz function ϕ with frequency support in the interval [ 1/ν,1/ν]. − For a tile s = I ω , define s s × (3.2) ϕ d=ef Mod Tran Dil2 ϕ. s c(ωs−) c(Is) |Is| Here, c(J) is the center of the interval J, and ω is the lower half of the interval ω . Thus, s− s this function is localized to be supported in the time frequency plane close to the rectangle I ω . s s− × There are companion functions which depend on different choices of certain measurable functions. These functions should be thought of as those choices of modulation and indices that will achieve, up to a constant multiple, the supremums in the oscillation function. To linearize the modulation, let N : R R be a measurable function (a modulation parameter). −→ We define another function related to the rectangle I ω which tells us when the linearized s s+ × modulation parameter is at a certain frequency. Let def (3.3) φ (x) = 1 (N(x))ϕ (x). s ωs+ s Now define a tile variant of the oscillation operator by ∞ 2 1/2 def (3.4) Tile-osc(f) = sup f,ϕ φ . s s h i hXj=1 kj≤l<l′<kj+1(cid:12)(cid:12)2l≤Xs|I∈sT|≤2l′ (cid:12)(cid:12) i (cid:12) (cid:12) Here, an increasing sequence of integers k : j 1 are specified in advance. We make the j { ≥ } definition for clarity’s sake, as we will not explicitly work with it. Rather we prefer to fully linearize this maximal operator. This requires the additional choices of functions ∞ (3.5) α : R R, α (x) 2 1, for all x , j j −→ | | ≤ j=1 X (3.6) ℓ ,ℓ : R Z, k ℓ < ℓ < k . j− j+ j j− j+ j+1 −→ ≤ And we set (3.7) F d=ef x : 2ℓj−(x) I < 2ℓj+(x) , s,j s { ≤ | | } def (3.8) f (x) = 1 (x)α (x)φ (x). s,j Fs,j j s 8 MICHAEL LACEY ANDERIN TERWILLEGER The sequences of functions ℓ are selecting the level at which the maximal difference occurs. j± The α are chosen to realize the ℓ2 norm in the definition of oscillation. We make all of these j choices in order to linearize the oscillation operator. Our main proposition is 3.9. Proposition. For all choices of N(x) and increasing sequence of integers k , the j { } operator Tile-osc extends to a bounded sub linear operator on Lp, 1 < p < . In particular, for sets G,H R of finite measure, we have ∞ ⊂ (3.10) 1 ,ϕ 1 ,f . min( G , H ) 1+ log |G| . |h G sih H s,ji| | | | | |H| s∈T Xj∈N (cid:0) (cid:12) (cid:12)(cid:1) (cid:12) (cid:12) Note that the inequality above implies that Tile-osc(1 ),1 . G 1/p H 1−1/p, 1 < p < . G H |h i| | | | | ∞ That is, we have the restricted weak type inequality for all 1 < p < .1 Hence, an interpo- ∞ lation argument will give us the estimate (3.11) Tile-osc(f) . f , 1 < p < . p p k k k k ∞ The Deduction of Proposition 2.4 and Proposition 2.5. Forξ R and ℓ Z, consider ∈ ∈ the operators def A f = 1 f,ϕ ϕ . ξ,l ξ∈ωs+h si s |IXs|=2l The tile oscillation operator is built up from these operators. Observe that these operators enjoy the properties (3.12) A Trans = Trans A , n Z, ξ,l n2l n2l ξ,l ∈ (3.13) A Dil(2) = Dil(2) A , l′ Z, ξ,l 2−l′ 2−l′ ξ2−l′,l+l′ ∈ (3.14) A Mod = Mod A , θ R. ξ,l −θ −θ ξ+θ,l ∈ Notice that these conditions tell us that the operators A have a near translation invariance, ξ,l a certain modulation invariance, and are related to each other through dilations. In addition, these operators are bounded on L2 uniformly in ξ and l, a fact well represented in the literature. We will now define 1 K L def B f = lim Mod Trans A (Trans Mod f)dydθ. ξ,l −θ −y (ξ+θ),l y θ K→∞ 4KL L→∞ Z−KZ−L 1In fact, the estimate (3.10) gives a favorable upper bound on the behavior of the constant with respect to p, namely that they are no more than max(p, p ). See [12] p−1 WIENER-WINTNER FOR HILBERT TRANSFORM 9 By periodicity of the integrand in y and θ, for all Schwartz functions f, the averages on the right hand side converge pointwise to B f(x) as K,L . ξ,l → ∞ Let us make some observations about the operators B . First, (3.12) and periodicity of ξ,l the integrand in y imply B commutes with translations. Second, it is a bounded, positive, ξ,l semidefinite operator, as is easy to see. Hence, it is given by convolution. Indeed, (3.14) implies that B f = Mod β (Mod f). ξ,l ξ l −ξ ∗ (1) for a function β that we turn to next. The equality (3.13) implies that β = Dil β where l l 2l β is given such that β is a smooth Schwartz function satisfying the conditions (2.1)—(2.3), 0 a routine exercise to verify. Assuming Proposition 3.9, it follows that we can conclude Proposition 2.4 and Proposi- tion 2.5 for nonzero functions ψ = β . Our proof is complete. 0 4. Main Lemmas Toprove (3.10), wesplit thesumover s into thesumover ssuch thatI M1 > λ s G ∈ T ⊂ { } and the sum over s such that I M1 > λ . The former sum can be taken care of by an s G 6⊂ { } argument of M. Lacey and C. Thiele [16] which also appears, slightly modified, in the paper of L. Grafakos, T. Tao, and E. Terwilleger [12]. Thus we restrict our attention to the tiles s where I M1 > λ . s G 6⊂ { } We begin with some concepts needed to phrase the proof. There is a natural partial order on tiles. We say that s < s′ iff ω ω and I I . Note that the time variable of s s s′ s s′ ⊃ ⊂ is localized to that of s′, and the frequency variable of s is similarly localized, up to the variability allowed by the uncertainty principle. Note that two tiles are incomparable with respect to the ‘<’ partial order iff the tiles, as rectangles in the time frequency plane, do not intersect. A “maximal tile” will be one that is maximal with respect to this partial order. Let denote an arbitrary set of tiles. We call a set of tiles T a tree if there is a S ⊂ S tile IT ωT, called the top of the tree, such that for all s T, s < IT ωT. We note that × ∈ × the top is not uniquely defined. An important point is that a tree top specifies a location in time variable for the tiles in the tree, namely inside IT, and localizes the frequency variables, identifying ωT as a nominal origin. We say that the count of is at most A iff = T, where each T is a tree S S T⊂S ⊂ T which is maximal with respect to inclusion and S def Count( ) = IT A. S | | ≤ T⊂S X 10 MICHAEL LACEY ANDERIN TERWILLEGER Fix χ(x) = (1 + x )−ν, where ν is, as before, a large constant whose exact value is | | unimportant to us. Define (1) (4.1) χ := Trans Dil χ, I c(I) |I| (4.2) dense(s) := sup χ dx, Is′ s<s′ZN−1(ωs′)∩H dense( ) := supdense(s), . S S ⊂ T s∈S The first andmost naturaldefinition of a “density” of a tile, would be I −1 N−1(ω ) I . s s+ s | | | ∩ | However ϕ is supported on the whole real line, although it does decay faster than the inverse of any polynomial. We refer to this as a “Schwartz tails problem.” The definition of density as χ dx, as it turns out, is still not adequate. That we should take the supremum N−1(ωs) Is over s < s′ only becomes evident in the proof of the “Tree Lemma” below. R The “Density Lemma” is 4.3. Lemma. Any subset is a union of and for which heavy light S ⊂ T S S dense( ) < 1 dense( ), Slight 2 S and the collection satisfies heavy S (4.4) Count( ) . dense( )−1 H . heavy S S | | What is significant is that this relatively simple lemma admits a non-trivial variant inti- mately linked to the tree structure and orthogonality. We should refine the notion of a tree. Calla tree T with topIT ωT a tree iff for each s T, aside fromthetop, IT ωT Is ωs± × ± ∈ × ∩ × is not empty. Any tree is a union of a +tree and a tree. If T is a +tree, observe that the − rectangles I ω : s T are disjoint. We see that s s− { × ∈ } f,ϕ 2 . f 2. |h si| k k2 s∈T X This motivates the definition (4.5) size( ) := sup IT −1/2 f,ϕs 2 : T , T is a +tree . S {| | |h i| ⊂ S } s∈T X The “Size Lemma” is 4.6. Lemma. Assume that f = 1 . Any subset is a union of and for which G big small S ⊂ T S S size( ) < 1 size( ), Ssmall 2 S and the collection satisfies big S (4.7) Count( ) . size( )−2 G . big S S | |

See more

The list of books you might like

Most books are stored in the elastic cloud where traffic is expensive. For this reason, we have a limit on daily download.