A stationary set method
for estimating oscillatory integrals
Abstract.
We propose a new method of estimating oscillatory integrals, which we call a “stationary set” method. We use it to obtain the sharp convergence exponents of Tarry’s problems in dimension two for every degree . As a consequence, we obtain sharp Fourier extension estimates for a family of monomial surfaces.
2020 Mathematics Subject Classification:
42B20 (Primary) 03C10 (Secondary)1. Statement of the main results
Our main result, Theorem 1.1, gives upper bounds on certain oscillatory integrals of bounded complexity.
Theorem 1.1 (“Stationary set” estimate).
Let . Let be a bounded semialgebraic function in variables with complexity . Then
(1.1) |
The implicit constant depends only on and .
Theorem 1.1 relies mainly on tools from real algebraic geometry and model theory, in particular, o-minimal geometry.
We will apply Theorem 1.1 to obtain new and often sharp results on the convergence exponent of Tarry’s problem in dimension and Fourier restriction for two-dimensional Parsell-Vinogradov surfaces. These applications only require Theorem 1.1 with functions being polynomials of bounded degrees, but our proof does not simplify in this special case. The more general setting of semialgebraic functions may also be of independent interest.
1.1. Sharpness of Theorem 1.1 in an average sense.
The argument in this subsection was shared to us by Wright, who further attributed it to Stein in the real setting and to Hua (see [Hua65, page 15]) in the -adic setting. A similar form of it was used for instance in [Wri20]. Their argument also works for general valued fields, Archimedean or non-Archimedian. Here we only present this argument in the real setting. It shows that the upper bound (1.1) is tight in an average sense (up to a log factor). Let us be more precise. Denote
(1.2) |
and
(1.3) |
We will prove that
(1.4) |
where is a bounded semialgebraic function of complexity , , and the implicit constants are allowed to depend on and .
The first inequality in (1.4) follows directly from Theorem 1.1. We only need to prove the second inequality. Let be a measurable function. Let be a Schwartz function satisfying for every whose Fourier transform is compactly supported. We first show that
(1.5) |
To see this, we write
(1.6) |
By the definition of , this is bounded by
(1.7) |
The bound (1.5) now follows from the triangle inequality.
1.2. Comparison to previously known tools for estimating oscillatory integrals
A well-known method of estimating oscillatory integrals is through van der Corput estimates, which assume a lower bound for for some multi-index . This method is very efficient in dimension one, see for instance Stein [Ste93], or Arkhipov, Chubarikov and Karatsuba [AKC79, Section 1.1] for a more quantitative, but perhaps lesser-known, formulation. In [CCW99], Carbery, Christ, and Wright established van der Corput estimates in dimensions greater than one with very mild assumptions on the relevant phase functions. We also refer to Wright [Wri11a], [Wri11b] and [Wri20] for estimating multi-dimensional oscillatory integrals in settings other than (for instance the -adic setting).
The results in the literature that are perhaps closest to ours were obtained by Bruna, Nagel and Wainger [BNW88] (see also Nagel, Seeger and Wainger [NSW93] for a related result). To describe this result, let be a smooth, compact hypersurface that is convex and of finite type. Let be the surface measure on . For , let denote the outward unit normal to at . Bruna, Nagel, and Wainger [BNW88, Theorem B] proved that, for large , we have
(1.10) |
The left-hand side of (1.10) is the Fourier transform of the surface-carried measure . The neighborhood of on the right-hand side of (1.10) plays a similar role to our stationary sets in Theorem 1.1.
In a slightly different direction, Varchenko [Var76] obtained asymptotic estimates for oscillatory integrals
(1.11) |
where is a large parameter, is an analytic phase function, and is a function supported near one point, say the origin. The asymptotic expansion involves terms in the Newton polyhedron of the phase function. Instead of asking a “global” condition of the form as in [CCW99], Varchenko assumed the support of to be small enough; he used the resolution of singularities for analytic functions, and his results are more “local”. For further work along this line, we refer to Karpushkin [Kar80] and [Kar83], Phong, Stein and Sturm [PSS99] and Greenblatt [Gre10], and the references therein. Operator-valued oscillatory integrals in a similar spirit have also been extensively studied. It is beyond our capability to give a review of even a small part of it, and therefore we refer interested readers to more classical works including Phong and Stein [PS97] and Seeger, Sogge and Stein [SSS91], and more recent works including Ikromov, Kempe and Müller [IKM10] and Ikromov and Müller [IM16].
1.3. Applications of the stationary set estimate
As an application of our stationary set estimate, we obtain the sharp convergence exponent of Tarry’s problem in dimension two for an arbitrary degree. Let and with . Set
(1.12) |
For a multi-index , denote . Let denote the vector . Let denote the -dimensional manifold
(1.13) |
which is sometimes referred to as a Parsell-Vinogradov manifold. For , define the Fourier extension operator associated to to be
(1.14) |
The convergence exponent of Tarry’s problem is defined to be
(1.15) |
Theorem 1.1 will serve as a main ingredient in the proof of the following theorem, in which we obtain the sharp convergence exponent of Tarry’s problem in dimension two.
Theorem 1.2.
For every , we have
(1.16) |
The quantity appears in the leading coefficient of the asymptotic expansion of Vinogradov’s mean value, as was first shown by Hua for large exponents [Hua47, Chapter 10.3], and in the full range of in [Woo19, Corollary 1.3]. The problem of determining the convergence exponent (1.15) goes back to Hua’s works [Hua52, Hua59]. In our notation, Hua proved that , and posed the problem of determining . This problem was resolved by Arkhipov, Chubarikov and Karatsuba [AKC79] (see also [ACK04, Theorem 1.3] in their book), where they proved that
(1.17) |
for every . For the connection to the original Tarry problem, we refer to [Woo19, Section 13].
In the same paper [AKC79], the authors also studied Tarry’s problem in higher dimensions, and obtained that
(1.18) |
for every . At the end of their paper, they also stated as an open problem to find the exact value of in higher dimensions. Later, Ikromov [Ikr97] obtained the lower bound
(1.19) |
In the quadratic case, that is, the case , the problem was resolved by Mockenhaupt [Moc96], who proved that for every .
For more partial results and for results on other systems of monomials, we refer to Chakhkiev [Cha03], Ikromov and Safarov [IS13], Safarov [Saf19], and Dzhabbarov [Dzh19, Dzh20].
Remark 1.3.
In [Dzh20], Dzhabbarov considered the case and claimed that The strategy in that paper is to show that . To bound the norm, the author there expanded the 12-th power of and applied some delicate change of variables, which further required some careful computations of Jacobians. Although this approach looks very natural, our results below (see Section 5, in particular, Remark 5.3) indicates that narrowly misses being integrable. Indeed,
(1.20) |
for every large where is the ball of radius centered at the origin.
As a consequence of Theorem 1.2, we obtain the following Fourier restriction estimates.
Corollary 1.4.
Let . For , let denote the smallest even number that is greater than or equal to . Then
(1.21) |
for every . Moreover, when is even, which is the same as saying , the above estimate is sharp in the sense that it fails for every .
When , it is reasonable to believe that (1.21) also holds for every . Here our result covers the range .
Recall that Theorem 1.2 states that (1.21) holds with whenever . That such an estimate implies (1.21) for a general for an even is standard, and is exactly how Mockenhaupt [Moc96] proved (1.21) for every and every . In the case of Mockenhaupt, the exponent is always even, which allows him to obtain sharp estimates in . In our case, our exponent may be odd, and by doing this reduction we may be off the sharp exponent of by 1.
Because of its applications to areas like partial differential equations, analytic number theory, and combinatorics, the Fourier restriction theory has received significant amount of attentions in the past few decades. A number of methods, including the bilinear method (see for instance Tao, Vargas and Vega [TVV98], Wolff [Wol01] and Tao [Tao03]), the multilinear method (see for instance Bennett, Carbery and Tao [BCT06] and Guth [Gut10]), the broad-narrow analysis of Bourgain and Guth [BG11] and the polynomial method (see for instance Guth [Gut16]), have been introduced to study the Fourier restriction phenomenon. Perhaps it is not exaggerating to say that each of the above mentioned papers generated an area of active research. Our method in the current paper seems to be disjoint from the above mentioned ones, and it would be very interesting to see how they can be connected.
1.4. A few remarks on Theorem 1.2 and Corollary 1.4
Let us first remark on the methods of proof of Theorem 1.2, and compare with those in the literature.
First of all, in the quadratic case in [Moc96], in order to bound the norm of , Mockenhaupt computed the function directly. To be more precise, he first smoothed out the function to , took the advantage of having a quadratic phase function and computed a Fourier transform directly (roughly speaking via completing squares). It seems difficult to generalize this approach to surfaces of cubic degree or higher. The result of Mockenhaupt [Moc96] has recently been recovered in [GOZZK20] via the broad-narrow analysis of Bourgain and Guth and decoupling inequalities for quadratic forms, and the approach there does not work either for degrees .
Next, we turn to the upper bound (1.18) established in [AKC79]. The authors of [AKC79] first derived an accurate pointwise bound for a -dimensional oscillatory integral, and then applied this bound to obtain a good pointwise bound for in terms of . Their pointwise bound is sharp for some, but not all, , resulting in an integrability exponent that is not sharp.
In our paper, we propose an entirely different approach, which we call a stationary set method. This method, combined with a careful study of the geometry of stationary sets in our problem (see Subsection 4.2), and certain rigidity properties of stationary sets (see Subsection 4.3), allows us to prove sharp integribility of for and every .
Remark 1.5.
We expect our stationary set method to be useful in obtaining the sharp convergence exponents of Fourier extension operators for more general semialgebraic sets. In a follow up paper, we are planning to show that the stationary set approach is often “tight” in the following sense. Let be a semialgebraic function and
be the size of the stationary set associated to . Define the convergence exponent and the stationary exponent by
We will show that . The same principle also applies to Fourier transforms of surface measures of semialgebraic sets.
Moreover, via the method of the current paper, we will also show that, for a semialgebraic set of dimension , it holds that if and only if the intersection of with every hyperplane of has dimension .
Although the problem of finding can be reduced to finding whenever , we remark that already for three dimensional Parsell-Vinogradov surfaces (for which the condition holds), we encountered significant technical difficulties when estimating the relevant stationary sets. We refer to Remark 4.11 for a more detailed discussion.
In the end, we comment on Corollary 1.4.
The estimate (1.21) with a sharp range of when was proven by Drury [Dru85]; indeed, he proved a stronger result with replaced by for an optimal range of as well.
As a consequence, Drury [Dru85] recovered the result of Arkhipov, Chubarikov and Karatsuba [AKC79] and gave an alternative proof for (1.17).
It is worth mentioning that Drury’s result can be partially recovered via the broad-narrow analysis of Bourgain and Guth [BG11] and the multi-linear restriction estimates of Bennett, Carbery and Tao [BCT06]; to be more precise, via these more recent tools, one can prove (1.21) for an optimal range of , and therefore obtain another proof of (1.17). Moreover, as was shown in [GOZZK20], if one combines the broad-narrow analysis of Bourgain and Guth, with the multi-linear Fourier restriction estimates of Bennett, Bez, Flock and Lee [BBFL18] and the decoupling inequalities for quadratic Parsell-Vinogradov surfaces in Guo and Zhang [GZ19], then one can also recover the result of Mockenhaupt [Moc96] (i.e. the case where and ).
We tried to prove Corollary 1.4 by this approach, but did not succeed.
For and , we encountered problems in the narrow part of the analysis; while for , we encountered significant difficulties in both the broad and the narrow part.
Acknowledgements. S.B. was supported in part by the NSF grant CCF-1910441. S.G. was supported in part by the NSF grant DMS-1800274. R.Z. was supported by the NSF grant DMS-1856541, DMS-1926686 and by the Ky Fan and Yu-Fen Fan Endowment Fund at the Institute for Advanced Study. P.Z. was supported in part by DFG (German Research Foundation) – EXC-2047/1 – 390685813. The authors would like to thank Xianghong Chen, Ziyang Gao, Phil Gressman, Jonathan Pila, Chieu-Minh Tran, Shaozhen Xu, Trevor Wooley and James Wright for valuable discussions at various stages of this project.
2. Bounding oscillatory integrals by the size of stationary sets
In this section, we prove Theorem 1.1, which says that we can bound the oscillatory integral
(2.1) |
by a constant times the measure of the largest “stationary set”
if the real phase function is semialgebraic with bounded complexity.
In order to prove the main theorem, we will change the integral to a one-dimensional integral of the form
(2.2) |
Then we will see that the conclusion follows once we prove that changes monotonicity finitely many times.
To prove the latter property, we have to use some tools from real algebraic, and more generally, o-minimal geometry. Recall that a semi-algebraic subset is, by definition, a finite union of subsets, each of which is defined by a formula of the form , where each . If the total number of polynomials occurring in the above formulas is bounded by , and the maximum degree of these polynomials is bounded by , we will say that the complexity of (as well as that of the formula) is bounded by . A semi-algebraic function is a function whose graph is a semi-algebraic set. We will say that the complexity of is bounded by if the complexity of the semi-algebraic set is bounded by .
We will need to define generalizations of the notion of semi-algebraic sets and functions. For this purpose, it is convenient to introduce some basic terminology from model theory, where these concepts originated.
2.1. A detour into model theory
In this subsection, we introduce some terminologies and tools in model theory. They will be useful when we prove Lemma 2.9, which is a crucial ingredient in the proof of Theorem 1.1.
A language is a triple of tuples
where each is a constant symbol, each is a function symbol of arity , and each is a relation symbol of arity .
An -structure is a tuple
where is a set, for each , , for each , is a function , and for each , is a -ary relation on . We call
interpretations in of the corresponding symbols in . It is a common abuse of notation to denote the structure by the underlying set , and we will take this liberty.
An -formula, is obtained using the logic connectives (denoting disjunction, conjunction and negation, respectively), and the quantifiers , from atomic formulas. An atomic formula is a formula either of the form , where are terms built out of the function and constant symbols and variables, or of the form where is an -ary relation symbol and are terms. A variable occurring in an -formula is called a bound variable of the formula, if it appears as or in . Otherwise, is called a free variable of . If are the free variables of a formula , we will denote this by writing as .
Given an -formula , an -structure , and , one can define in the obvious way (using induction on the structure of the formula ) whether is true in the structure (denoted ). Thus, every -formula defines a subset, , of , defined by
Example 2.1.
If denotes the language of ordered fields, where are constant symbols, are binary function symbols, and is a binary relation symbol, then are variables, are terms, and
(2.3) |
is a formula with free variables , and a bound variable .
Denoting by the -structure whose underlying set is the real numbers, and with interpreted in the usual way, we have
Given an -structure , a formula with free variables
and a tuple , we call an -formula (-formula with parameters in ) with free variables .
An -theory is a set of of -formulas without free variables (also called -sentences). An -structure is a model of (denoted ), if for each . An -theory admits quantifier elimination if for every -formula there exists a quantifier-free -formula such that for every model of , (here and elsewhere, is the standard abbreviation for the formula ).
The theory of the -structure , denoted , is the set of -sentences which are true in . Two -formulas and are said to be equivalent modulo , if . The definable sets of an -structure are sets of the form
where is an -formula. The definable functions are those whose graphs are definable sets.
For the rest of the article, we will denote by the language of ordered fields.
The set with the usual interpretations of is an -structure. Notice that, by definition, semi-algebraic sets are definable sets in the -structure . The Tarski-Seidenberg theorem [Tar51] states that the the -theory, , of real closed ordered fields admits quantifier elimination. Here, is the set of -sentences axiomatizing ordered fields, in which every non-negative element is a square and every polynomial having odd degree has a root. By definition, we see that the field (with the usual interpretations of ) is a model of . As a consequence of the Tarski-Seidenberg theorem [Tar51], one immediately obtains that every definable set of the -structure is semi-algebraic.
As an illustration of Tarski-Seidenberg theorem, observe that the -formula in (2.3) (which has quantifiers) is equivalent modulo the theory to the quantifier-free -formula
As a consequence,
Another simple geometric consequence of the Tarski-Seidenberg theorem is that if and are a semialgebraic sets in , then the set is also semialgebraic.
While the Tarski-Seidenberg theorem is not quantitative as stated above, there are effective versions with complexity estimates. In particular, given any -formula , there exists a quantifier-free -formula equivalent to modulo , such that the number and degrees of the polynomials appearing in is bounded by a function of the number and degrees of the polynomials appearing in , as well as the number of quantified and free variables. This is in fact a consequence of the original proof of Tarski [Tar51] of the Tarski-Seidenberg theorem, and we omit detailed explanation here. We remark that the above mentioned function can be made explicit, see for instance [BPR06, Theorem 14.16].
2.1.1. o-minimal structures
Let be any language containing the binary relation , and let be an -structure such that the relation defines a total order on . The structure is called o-minimal if the set of definable subsets of is the smallest possible – namely, the set whose elements are finite unions of points and intervals (such subsets are clearly definable by quantifier-free -formulas).
Example 2.2.
Recall that is the language of ordered fields. The -structure is an example of an o-minimal structure.
Suppose that is a language containing (we call an expansion of ). An -structure is called an o-minimal expansion of if the underlying set of the structure is , and the structure is o-minimal (i.e., the definable subsets of are precisely the semi-algebraic subsets of ).
Let us look at an example to better digest this notion. In the following example, we expand the language with one additional function symbol.
Example 2.3.
Let be the expansion of by an additional -ary function symbol , and consider the -structure whose underlying set is , with the symbols of interpreted as before, and the -ary function symbol interpreted by the trigonometric function . Then, is not an o-minimal expansion of , since the definable subset of defined by the formula is the set , which cannot be expressed as a finite union of points and intervals of .
2.1.2. The o-minimal structures and
The above example shows that the real analytic functions cannot all be definable functions in any o-minimal expansion of . For instance, the function is never definable in any o-minimal expansion of .
However, notice that if we consider a function , which is equal to on , and defined to be outside of the , then the set of zeros of is a semi-algebraic set and hence definable in every o-minimal expansion of .
Motivated by this example, we call a restricted analytic function, if there is an open subset containing in , and an analytic function , such that
(2.4) |
Let denote the expansion of obtained by including a function symbol (of appropriate arity) for each restricted analytic function. We denote by the structure with underlying set , with the usual interpretations of the symbols of , and where we interpret the new symbols by the corresponding restricted analytic functions. It is a theorem of Gabrielov [Gab68] that the -structure is an o-minimal expansion of .
The (unrestricted) exponential function is not a restricted analytic function. We expand the language further and include a symbol for the exponential function and denote the new language by . Denote by the -structure obtained by interpreting the new symbol by the exponential function, and interpreting the symbols of as before. It is a theorem of van den Dries, Macintyre, and Marker that:
Theorem 2.4 ([vdDMM94]).
The -structure is an o-minimal expansion of .
Remark 2.5.
Notice that the logarithm function is definable in , since its graph
is definable in .
Let denote an o-minimal expansion of . If we denote by for each the set of subsets of which are definable in , then the tuple satisfy the following properties, which follow directly from the definition of an o-minimal expansion.
-
(a)
If and , then . If and , then .
-
(b)
If is the projection map and , then .
-
(c)
Elements of are precisely the finite unions of points and open intervals.
Remark 2.6.
While the set of definable sets in an o-minimal expansion of is stable under finite set operations and projections (Properties (a) and (b) above), the set of definable functions is not necessarily closed under taking parametric integrals. More precisely, if is a definable function in , then the function
(2.5) |
is not necessarily definable in . It suffices to take the example of the function
(2.6) |
whose graph is a semi-algebraic set and so is clearly definable in the o-minimal structure . However, the parametric integral is the logarithm function whose graph is not a semi-algebraic and hence is not definable in . On the other hand, notice that is definable in the o-minimal expansion (cf. Remark 2.5).
In fact, it is proved in [CM11, Theorem 1.3], that if is a constructible function, then the function defined in (2.5) is constructible (cf. Definition 1.2 in [CM11]). We omit the definition of constructible functions in the sense of [CM11], but list two of their properties, which are the only facts that we will need in this paper.
-
(1)
Constructible functions are definable in the o-minimal expansion of . This follows easily from the definition of constructible functions and the fact that the logarithm function is definable in .
-
(2)
Semi-algebraic functions are constructible.
An important consequence of the properties (a), (b) and (c) is the topological tameness of definable sets in an o-minimal expansion of . Let be a -structure which is an o-minimal expansion of . Then the definable sets of can only have a finite number of connected components (this is a consequence of [vdD98, Theorem 2.9]).
Another important property of the definable sets is that of topological tameness in families.
Proposition 2.7 ([vdD98, Chapter 9, Section 2]).
Let be a -structure which is an o-minimal expansion of . Let and be definable sets and let be a definable function. Then, there are finitely many homeomorphism types amongst the fibers . In particular, there exists a constant , such that the number of connected components of is bounded by .
As an application of Proposition 2.7, we have:
Proposition 2.8.
Let be a -structure which is an o-minimal expansion of . Let be a definable function, and for , let denote the number of times the function changes monotonicity. Then, .
Proof.
First observe that for each , the function is a definable function. Using the monotonicity theorem for definable functions [vdD98, Theorem 1.2, Chapter 3], for each , there exists a finite partition of into points and open intervals, such that over each open interval the function is either constant, strictly increasing, or strictly decreasing. Hence, the set of points at which the function changes monotonicity is contained in the set of end points of these intervals, and in particular is a finite subset of .
Now, let . Since changing monotonicity of a function is expressible using a first-order formula (involving quantifiers) in the language , it follows that the set is definable in o-minimal structure .
Now, applying Proposition 2.7, we get that there are finitely many homeomorphism types amongst the fibers of the map restricted to , noting that this map is clearly definable in . Moreover, each fiber has finitely many points using the observation in the first paragraph of the proof. The proposition follows. ∎
2.2. Proof of Theorem 1.1
With the preparation above, let us prove a lemma that will play a crucial role in the proof of Theorem 1.1.
Lemma 2.9.
Let and let be a semialgebraic set in with complexity . Assume further that is bounded. For each , we define
(2.7) |
Then the function changes monotonicity times.
Proof of Lemma 2.9.
First observe that for each there exists , and a finite set
where each is a quantifier-free -formula, such that if is a semi-algebraic set of complexity bounded by , then there exists and such that
Intuitively, the above property holds because by definition, our set can be described by an -formula involving many polynomials of degree , parameters in , logic connectives and constant symbols.
For each , let be the set defined by , and let be the projection map to the first factor.
For , consider the functions on defined as follows: For and , we set
(2.8) |
whenever the above integral is finite, and otherwise.
The functions are semialgebraic and thus definable in . By Theorem 1.3 in [CM11] (as mentioned in Remark 2.6), is a constructible function on . Using the properties of constructible functions listed in Remark 2.6, we obtain that is definable in .
It now follows from Proposition 2.8 that there exists , such that the number of times the function changes monotonicity is bounded by . Let . It follows that the number of times the function changes monotonicity is bounded by , observing that there exists , , such that for . ∎
Proof of Theorem 1.1.
Fix a constant (in our proof will work). In the proof we will suppress the dependence on the constant . By a change of variables, we have
Here, denotes the Lebesgue measure. Note that is a nonzero constant (depending only on ) as . Hence, if we define a stationary set to be
(2.9) |
then we have
(2.10) |
Since is bounded, the function is compactly supported. We claim that it is piecewise monotonic and that it only changes monotonicity many times. This claim follows from Lemma 2.9 in a straightforward way, and we postpone its justification to the end of this proof.
We now explain how to get the desired conclusion from the above claim. We decompose into a union of many closed intervals such that the interiors of are disjoint and that the function is compactly supported and monotonic on each . For each , the total variation of on is bounded by , uniformly in . By integration by parts, we see
(2.11) |
uniformly in . Hence, we get
(2.12) |
by assumption. This is the desired conclusion.
Finally, we justify our claim that is piecewise monotonic and that it only changes monotonicity many times. Consider the set
(2.13) |
The set is bounded and can be described as a “vertical” neighborhood of the graph of .
Observe that since is a semi-algebraic function, is a semi-algebraic subset of and so defined by a -formula . Then is defined by the formula
Using an effective version of Tarski-Seidenberg theorem (that was mentioned in Subsection 2.1; see e.g. [BPR06, Theorem 14.16] for a reference), we obtain that is equivalent to a quantifier-free formula such that the number of polynomials and their degrees that occur in is . And so is semialgebraic with complexity . Now we apply Lemma 2.9 to the set , and note that in the notation of that lemma. By the conclusion of Lemma 2.9, we deduce the claim in the beginning of this paragraph. ∎
Remark 2.10.
In Theorem 1.1, the amplitude function of the oscillatory integral is taken to be the sharp cut-off . For a compactly supported smooth amplitude function, one can apply a Fourier series expansion and then apply Theorem 1.1. In applications, due to the rapid decay of Fourier coefficients, the extra summation that is brought in by Fourier series expansion usually will not cause any problem. For example, in Tarry’s problem, one can insert a compactly supported smooth amplitude function in (1.14), and ask what the convergence exponent is. It is elementary to see that it coincides with the one of sharp cut-off, that is, .
3. Theorem 1.2: Reduction to a spread out case
Sections 3 and 4 are devoted to the proof of Theorem 1.2. From now on, we will fix and abbreviate to . To simplify notation, we denote
(3.1) |
Under this notation, to prove Theorem 1.2, it is equivalent to prove that for every .
We will use to denote the function . Our goal is to show that
(3.2) |
for every , where
(3.3) |
This will imply the upper bound of that . We will use to denote the polynomial . For a polynomial and real numbers and , define
(3.4) |
For a given , since the measure of depends continuously on , we can choose such that
(3.5) |
To simplify notation, we denote
(3.6) |
From Theorem 1.1, we immediately obtain that there exists such that for every , it holds that
(3.7) |
where
We will show that there exists such that
(3.8) |
for every . This will be proven via an inductive argument. More precisely, we will show that there exists such that for every sufficiently large dyadic number (picking will be enough), we can find another constant such that
(3.9) |
Since , this implies our claim (3.8) after choosing so large that
and unrolling the recursion (3.9).
Remark 3.1.
Remark 3.2.
The choice of will be made in Lemma 4.7.
Let be large dyadic number to be determined. Let be the collection of all (vertical) dyadic rectangles of size . We have . Let be a large constant to be determined. We write as a disjoint union
(3.10) |
where
(3.11) |
We will control the sizes of and separately. For , we say that it is in the concentrated case if and that it is in the spread out case if .
Remark 3.3.
The concentrated/spread out dichotomy here is inspired by the broad-narrow dichotomy of Bourgain-Guth in [BG11]. We leave the comparison to the interested readers.
In the remaining part of this section we will control . Let us start with estimating for a fixed . Without loss of generality, take . For an , define
(3.12) |
with , so that . Then, for , we have
(3.13) |
By definition of , we have
(3.14) |
By scaling, this implies
(3.15) |
We sum over and obtain
(3.16) |
This finishes the argument for the concentrated case.
4. Theorem 1.2: The spread out case
In this section, we will show that
(4.1) |
thus finishing the proof of the desired estimate (3.9). Next we sketch the proof ideas before working out the details,
For , recall that the stationary set has measure at least . Let us think about what such a may look like. Note that has bounded complexity. In the second subsection, we will prove that a semialgebraic set of bounded complexity in that has measure at least is “not far from” a union of -cubes. This is essentially Lemma 4.9 in two dimensions.111There is a higher dimensional version that can be proved in an identical way. We omit that more general version here since it is irrelevant to our current problem. Before proving such a lemma we will need some preparations (done in the first subsection) using elementary real algebraic geometry.
Since has bounded complexity, the (quantitative) Tarski-Seidenberg theorem shows that its projection to the -axis is a union of boundedly many points and intervals. If we pretend that is a union of -squares, by the fact that it is spread out in the direction (thus hitting many vertical -strips), we see that the projection of to the -axis has to contain a “long interval” (with constant length) if we choose large enough. This means we can find an interval on the -axis of constant length such that for all dyadic interval of length in , there is a -square in whose projection to the -axis is . In real life, is not honestly a union of -squares but we still have something very similar to the conclusion above thanks to Lemma 4.7.
Finally, we fix a constant (, to be more accurate) many evenly-spaced and look at all possible tuples . For each tuple, if every -square in it is essentially contained in , we will see that has to be contained in a very specific rectangular box (depending on the particular tuple ). Geometrically, this is the dual box to the convex hull of all caps corresponding to all in that tuple. Therefore the union of all such rectangular boxes will contain . Simply by adding up the volume of all the rectangular boxes we get the desired upper bound of and finish the proof. This is carried out in the last subsection.
Remark 4.1.
For general when , we expect this proof framework to continue working but also expect a lot more new technical difficulties and hence do not pursue the more general problem here. See Remark 4.11 for a discussion.
4.1. Shifts of semialgebraic sets
Lemma 4.2 (Finite boundary).
For every there exists such that every semialgebraic subset of complexity has at most boundary points.
Proof of Lemma 4.2.
By definition, is a union of many subsets where each is a set defined by many polynomial equations or inequalities of degree . This implies that each is an intersection of many larger sets, each larger set being a union of many points or intervals. Hence each is a union of many points or intervals. Taking the union, we deduce that is also a union of many points or intervals and the conclusion follows. ∎
Corollary 4.3 (Shifts of a bounded complexity set).
Let be a semialgebraic set of complexity . Then, for every , we have
Proof of Corollary 4.3.
By Lemma 4.2, is the union of many intervals, which immediately implies the desired bound. ∎
Lemma 4.4.
For every , there exists a small constant such that, for every and every semialgebraic set with complexity and measure , we have
Proof of Lemma 4.4.
Let and be the unit vectors. Let be a small number that is to be chosen. For any , the set
will be a semialgebraic set of complexity at most . Using Corollary 4.3 in the first variable and integrating, we see that
for some large constant depending only on . If , this will be . Using Corollary 4.3 in the second variable and integrating, we see that
for some large constant depending only on . If , then this is . Summarizing, it suffices to take . ∎
Remark 4.5.
Lemma 4.4 is also true when the dimension is replaced by an arbitrary . One just needs to repeat the procedure in the proof times. We do not need that more general result in this paper but record it here as it may be of independent interest.
4.2. Large projections of stationary sets
In this subsection, we fix , and discuss some geometry of a particular stationary set .
We apply Lemma 4.4 to the set . Its complexity is bounded by a constant depending on . Therefore, we can find a small constant such that for the set
(4.2) |
we have . If we also know that , then we will see that the set has a “large” projection in the horizontal axis.
Lemma 4.6.
For every , it holds that
(4.3) |
Proof of Lemma 4.6.
Let be the set consisting of those strips that have non-empty intersection with . Using that for any , we obtain
Rearranging the terms gives the claim. ∎
Lemma 4.7.
If is large enough depending on , then, for every , the projection contains a dyadic interval of length . Here denote the projection to the first variable.
Proof of Lemma 4.7.
Using an effective version of quantifier-elimination in the theory of real closed fields (that was mentioned in Subsection 2.1; see e.g. [BPR06, Theorem 14.16] for a reference) is semialgebraic set having bounded complexity depending only on . By Lemma 4.2, it is the union of at most intervals, where depends only on . If , then it follows from Lemma 4.6 that contains at least one full dyadic interval of length . ∎
Remark 4.8.
From this point on, all estimates are allowed to depend implicitly on both and ; these implicit constants will accumulate to the constant in (3.9). To simplify notation, the dependence on and will often be dropped.
By Lemma 4.7 and a standard pigeonholing argument, we can find a subset and an interval of length such that
and for every we have . Let
(4.4) |
be a -separated set.
Let be the partition of into dyadic squares of side length . Next, for each , we will construct a sub-collection .
Roughly speaking, the purpose of this construction is that we want to be approximated by a union of certain lattice dyadic -squares. To this end, we choose sufficiently many such that the union of all approximates well. Moreover when , we want to have for every so that is in a slightly thickened stationary set. This is guaranteed by the following Lemma whose proof uses the Lagrange interpolation.
Lemma 4.9.
Let . Then, for each and , we have
(4.5) |
Proof of Lemma 4.9.
In the proof of this lemma we will see the motivation of introducing (4.2). We take one . By definition, we can find such that . By the definition of as in (4.2), we know that
(4.6) |
As a consequence, we obtain that
(4.7) |
The desired bound (4.5) follows from applying the interpolation polynomials in the Lagrange form. ∎
In particular,
(4.8) |
for every and every .
4.3. Rigidity of stationary sets
For , let be the closure of the convex hull of . By John’s ellipsoid lemma, we know that there exists a rectangular box such that
(4.9) |
where is a large constant that depends only on . Let denote the center of . Let denote
(4.10) |
where is the same as the one in Lemma 4.9. Notice that .
Recall from (4.5) that, for every , we have
(4.11) |
Denote
(4.12) |
We obtain that
(4.13) |
for every , and therefore
(4.14) |
for every , which further means .
4.4. Estimating dual boxes
In this section, we will bound . Recall the choice of from previous subsections. Let denote
(4.15) |
For a fixed tuple
(4.16) |
denote
(4.17) |
Therefore we can write
(4.18) |
Let us focus on estimating . Recall the discussion in Subsection 4.3. Let us use to denote . We have
(4.19) |
We will get a good lower bound on by choosing a transverse set of line segments inside this convex set. A standard geometric observation is that each set contains line segments of length in the direction of -th derivatives of :
Lemma 4.10.
For every , every , and every multiindex with , the convex hull contains a line segment of length in the direction .
From now on, we order the monomials in in the increasing lexicographic order and write
(4.20) |
We also order the ’s in a convenient way. By the Vandermonde determinant formula, we know that
Let be the vector whose Hodge dual is
It follows that . By rearranging if necessary, we may assume that the last entry of is . Doing this in the reverse order of , we may assume that this holds for every .
With this order of ’s, we will use the following lower bound:
Here, we have chosen a total of
vectors in the difference set of the convex set , of which the first come from differences between and , while the remaining ones come from individual ’s via Lemma 4.10.
Writing these vectors as rows of a matrix, we see that, up to a -dependent constant, the latter wedge product is the determinant of the block upper triangular matrix
with
Since is a Vandermonde matrix, its determinant is . Consider next
(4.21) |
We fix some with
(4.22) |
and select for recursively in such a way that (4.22) holds and the absolute value of (4.21) is maximized. By our arrangement of ’s, the last entry of is . Therefore, given , the set of ’s for which is the union of at most two intervals of length . It follows that, given and , there are many for which , and there are no other possibilities.
Remark 4.11.
If we follow the proof of Theorem 1.2 in higher dimensions , the problem of estimating the spread out set becomes much more difficult. In dimension , we greatly benefited from a factorization of the upper bound (4.19) into linear factors given by (4.21). Unless a similar upper bound with good factorization can be found in higher dimensions, we would face complicated quantitative real algebraic problems. We feel it may be difficult to find such an upper bound at this moment, and hence do not pursue the problem for , , in the current paper.
5. Lower bounds
In this section, we prove lower bounds for . The idea is not difficult to explain. We will show that the function is “large” (see Lemma 5.2) on a disjoint union of rectangular boxes (see Lemma 5.1). This idea has already been used by Arkhipov, Chubarikov and Karatsuba [AKC79, ACK04] and by Ikromov [Ikr97]. The lower bound in Lemma 5.2 is essentially the same as in the above mentioned papers; the key difference is that we are able to find more rectangular boxes where such a lower bounds holds. Roughly speaking, this is achieved via applying the “rotation symmetry” of Parsell-Vinogradov surfaces.
Let be a large number. Let , . Let , . The choice of and is such that
(5.1) |
and what we will need is that is close to 0, and is away from both 0 and 1.
Recall that . Let us write
(5.2) |
For given and , we let and write
(5.3) |
where for each , and , the function is a polynomial of degree in . In particular, is a constant function that is independent of and ; indeed, , which follows from the normalization in (5.2). Define
(5.4) |
Here is a sufficiently small parameter that will be picked later, and for a polynomial , we use to denote the sum of all its coefficients.
Lemma 5.1.
If (depending only on ) is chosen to be sufficiently small, then for each , it holds that
(5.5) |
and for , we have .
Lemma 5.2.
For each and each , it holds that
(5.6) |
provided that is chosen to be sufficiently small, depending only on .
The desired lower bound follows immediately from these two lemmas.
Remark 5.3.
In the end, we will prove Lemma 5.1 and Lemma 5.2. We will start with the latter one. Its proof relies on the following lemma in Ikromov [Ikr97], a slightly different form of which also already appeared in Arkhipov, Chubarikov and Karatsuba [AKC79, ACK04].
Lemma 5.4 ([Ikr97]).
Let and . Suppose that the polynomial satisfies
(5.7) |
for some large and small . Then we have the following asymptotic representation
(5.8) |
where is a constant depending on satisfying
(5.9) |
with depending only on .
Proof of Lemma 5.2.
Let us write our integral as
(5.10) |
After the change of variable
(5.11) |
for each fixed , the integral becomes
(5.12) |
In the case , since , Lemma 5.4 gives a uniform lower bound for the inner integral, provided that is chosen to be sufficiently small. In the case , we complete squares in the phase and use the fact that the Fresnel integral is bounded below by a positive constant as soon as is bounded away from , while the Fresnel integral is bounded above.
In both cases, the value of the inner integral is contained in a cone in with angle and bounded away from . Using the upper bound , we can complete the proof of (5.6). ∎
Proof of Lemma 5.1.
Regarding the (elementary) lower bound (5.5), we first mention that this has also been observed by Ikromov in the case , see [Ikr97, page 183]. To show (5.5), note that the coefficient of in equals plus some function of and the components with .
Hence, for each , once have been chosen for all and all , we can choose each with in an interval of length . The measure of the set of all that we can choose is then proportional to to the power
In the end, we show that whenever . We argue by contradiction. Let and assume that there exists an . We now expand in the form of (5.3) with and separately. By only considering the first term in these two expansions, we obtain
(5.13) |
as functions depending on and . Recall the definition (5.4). We will arrive at a contradiction if we can show that the following polynomial in and
(5.14) |
has a coefficient of a non-constant term with absolute value .
The monomial has the coefficient , and the monomial has the coefficient . At least one of these coefficients has absolute value . ∎
6. Proof of Corollary 1.4
Fix , and with . We abbreviate to . Let be the measure supported on the surface whose projection to is given by , and let be that given by . We can therefore write
(6.1) |
Let be a non-negative smooth bump function supported on satisfying that is non-negative and for every . Define . Recall that is the smallest even number . As a consequence of Theorem 1.2 and Hölder’s inequality, we obtain that for every and , it holds that
(6.2) |
where is an arbitrary ball of radius . We will show that
(6.3) |
To prove this estimate, we first observe that, by modulating the function , which does not change its norm, it suffices to consider the case that is centered at the origin. In this case, we have
(6.4) |
We apply Plancherel’s theorem to the last term and obtain
(6.5) |
We apply Plancherel’s theorem back and obtain (6.3). In the end, to pass from the local estimate (6.3) to the desired global estimate in the corollary, one can follow Tao’s epsilon-removal lemma in [Tao99] and its variants in Bourgain and Guth [BG11] and Kim [Kim17].
References
- [ACK04] G. I. Arkhipov, V. N. Chubarikov, and A. A. Karatsuba. Trigonometric sums in number theory and analysis, volume 39 of De Gruyter Expositions in Mathematics. Walter de Gruyter GmbH & Co. KG, Berlin, 2004. Translated from the 1987 Russian original.
- [AKC79] G. I. Arhipov, A. A. Karacuba, and V. N. Chubarikov. Trigonometric integrals. Izv. Akad. Nauk SSSR Ser. Mat., 43(5):971–1003, 1197, 1979.
- [BBFL18] Jonathan Bennett, Neal Bez, Taryn C. Flock, and Sanghyuk Lee. Stability of the Brascamp-Lieb constant and applications. Amer. J. Math., 140(2):543–569, 2018.
- [BCT06] Jonathan Bennett, Anthony Carbery, and Terence Tao. On the multilinear restriction and kakeya conjectures. Acta Math., 196(2):261–302, 2006.
- [BG11] Jean Bourgain and Larry Guth. Bounds on oscillatory integral operators based on multilinear estimates. Geom. Funct. Anal., 21(6):1239–1295, 2011.
- [BNW88] Joaquim Bruna, Alexander Nagel, and Stephen Wainger. Convex hypersurfaces and Fourier transforms. Ann. of Math. (2), 127(2):333–365, 1988.
- [BPR06] Saugata Basu, Richard Pollack, and Marie-Françoise Roy. Algorithms in real algebraic geometry, volume 10 of Algorithms and Computation in Mathematics. Springer-Verlag, Berlin, second edition, 2006.
- [CCW99] Anthony Carbery, Michael Christ, and James Wright. Multidimensional van der Corput and sublevel set estimates. J. Amer. Math. Soc., 12(4):981–1015, 1999.
- [Cha03] M. A. Chakhkiev. On the exponent of convergence of a special integral of a multidimensional analogue of Tarry’s problem. Izv. Ross. Akad. Nauk Ser. Mat., 67(2):211–224, 2003.
- [CM11] Raf Cluckers and Daniel J. Miller. Stability under integration of sums of products of real globally subanalytic functions and their logarithms. Duke Math. J., 156(2):311–348, 2011.
- [Dru85] Stephen W. Drury. Restrictions of Fourier transforms to curves. Ann. Inst. Fourier (Grenoble), 35(1):117–123, 1985.
- [Dzh19] I. Sh. Dzhabbarov. Convergence exponent of a special integral in the two-dimensional Tarry problem with a homogeneous polynomial of degree 2. Mat. Zametki, 105(3):375–382, 2019.
- [Dzh20] I. Sh. Dzhabbarov. On the Multidimensional Tarry Problem for a Cubic Polynomial. Mat. Zametki, 107(5):657–673, 2020.
- [Gab68] A. M. Gabrièlov. Projections of semianalytic sets. Funkcional. Anal. i Priložen., 2(4):18–30, 1968.
- [GOZZK20] Shaoming Guo, Changkeun Oh, Ruixiang Zhang, and Pavel Zorin-Kranich. Decoupling inequalities for quadratic forms, 2020.
- [Gre10] Michael Greenblatt. Oscillatory integral decay, sublevel set growth, and the Newton polyhedron. Math. Ann., 346(4):857–895, 2010.
- [Gut10] Larry Guth. The endpoint case of the Bennett-Carbery-Tao multilinear Kakeya conjecture. Acta Math., 205(2):263–286, 2010.
- [Gut16] Larry Guth. A restriction estimate using polynomial partitioning. J. Amer. Math. Soc., 29(2):371–413, 2016.
- [GZ19] Shaoming Guo and Ruixiang Zhang. On integer solutions of Parsell-Vinogradov systems. Invent. Math., 218(1):1–81, 2019.
- [Hua47] Loo-Keng Hua. The additive prime number theory. Trav. Inst. Math. Stekloff, 22:179, 1947.
- [Hua52] Loo-Keng Hua. On the number of solutions of Tarry’s problem. Acta Sci. Sinica, 1:1–76, 1952.
- [Hua59] Loo-Keng Hua. On the tarry problem. Proc. Third All-Union Math. Congr. (Moscow, 1956). Vol. IV: Survey Reports, Izdat. Akad. Nauk SSSR, Moscow, pages 140–143, 1959.
- [Hua65] L. K. Hua. Additive theory of prime numbers. Translations of Mathematical Monographs, Vol. 13. American Mathematical Society, Providence, R.I., 1965.
- [IKM10] Isroil A. Ikromov, Michael Kempe, and Detlef Müller. Estimates for maximal functions associated with hypersurfaces in and related problems of harmonic analysis. Acta Math., 204(2):151–271, 2010.
- [Ikr97] I. A. Ikromov. On the convergence exponent of trigonometric integrals. Tr. Mat. Inst. Steklova, 218(Anal. Teor. Chisel i Prilozh.):179–189, 1997.
- [IM16] Isroil A. Ikromov and Detlef Müller. Fourier restriction for hypersurfaces in three dimensions and Newton polyhedra, volume 194 of Annals of Mathematics Studies. Princeton University Press, Princeton, NJ, 2016.
- [IS13] I. A. Ikromov and A. R. Safarov. On estimates for double trigonometric integrals. Uzbek. Mat. Zh., (4):7–15, 2013.
- [Kar80] V. N. Karpushkin. Uniform estimates of oscillating integrals in . Dokl. Akad. Nauk SSSR, 254(1):28–31, 1980.
- [Kar83] V. N. Karpushkin. Uniform estimates of oscillating integrals with a parabolic or a hyperbolic phase. Trudy Sem. Petrovsk., (9):3–39, 1983.
- [Kim17] Jongchon Kim. Some remarks on fourier restriction estimates, 2017.
- [Moc96] Gerd Mockenhaupt. Bounds in Lebesgue spaces of oscillatory integral operators, 1996.
- [NSW93] Alexander Nagel, Andreas Seeger, and Stephen Wainger. Averages over convex hypersurfaces. Amer. J. Math., 115(4):903–927, 1993.
- [PS97] D. H. Phong and E. M. Stein. The Newton polyhedron and oscillatory integral operators. Acta Math., 179(1):105–152, 1997.
- [PSS99] D. H. Phong, E. M. Stein, and J. A. Sturm. On the growth and stability of real-analytic functions. Amer. J. Math., 121(3):519–554, 1999.
- [Saf19] A. R. Safarov. On the -bound for trigonometric integrals. Anal. Math., 45(1):153–176, 2019.
- [SSS91] Andreas Seeger, Christopher D. Sogge, and Elias M. Stein. Regularity properties of Fourier integral operators. Ann. of Math. (2), 134(2):231–251, 1991.
- [Ste93] Elias M. Stein. Harmonic analysis: real-variable methods, orthogonality, and oscillatory integrals, volume 43 of Princeton Mathematical Series. Princeton University Press, Princeton, NJ, 1993. With the assistance of Timothy S. Murphy, Monographs in Harmonic Analysis, III.
- [Tao99] Terence Tao. The Bochner-Riesz conjecture implies the restriction conjecture. Duke Math. J., 96(2):363–375, 1999.
- [Tao03] Terence Tao. A sharp bilinear restrictions estimate for paraboloids. Geom. Funct. Anal., 13(6):1359–1384, 2003.
- [Tar51] Alfred Tarski. A decision method for elementary algebra and geometry. University of California Press, Berkeley and Los Angeles, Calif., 1951. 2nd ed.
- [TVV98] Terence Tao, Ana Vargas, and Luis Vega. A bilinear approach to the restriction and Kakeya conjectures. J. Amer. Math. Soc., 11(4):967–1000, 1998.
- [Var76] A. N. Varčenko. Newton polyhedra and estimates of oscillatory integrals. Funkcional. Anal. i Priložen., 10(3):13–38, 1976.
- [vdD98] Lou van den Dries. Tame topology and o-minimal structures, volume 248 of London Mathematical Society Lecture Note Series. Cambridge University Press, Cambridge, 1998.
- [vdDMM94] Lou van den Dries, Angus Macintyre, and David Marker. The elementary theory of restricted analytic fields with exponentiation. Ann. of Math. (2), 140(1):183–205, 1994.
- [Wol01] Thomas Wolff. A sharp bilinear cone restriction estimate. Ann. of Math. (2), 153(3):661–698, 2001.
- [Woo19] Trevor D. Wooley. Nested efficient congruencing and relatives of Vinogradov’s mean value theorem. Proc. Lond. Math. Soc. (3), 118(4):942–1016, 2019.
- [Wri11a] James Wright. From oscillatory integrals and sublevel sets to polynomial congruences and character sums. J. Geom. Anal., 21(1):224–240, 2011.
- [Wri11b] James Wright. From oscillatory integrals to complete exponential sums. Math. Res. Lett., 18(2):231–250, 2011.
- [Wri20] James Wright. The h and j functionals in the theory of complete exponential sums. arXiv preprint arXiv:2012.08211, 2020.