Restrictions of higher derivatives of the Fourier transform

By Michael Goldberg and Dmitriy Stolyarov

Abstract

We consider several problems related to the restriction of to a surface with nonvanishing Gauss curvature. While such restrictions clearly exist if is a Schwartz function, there are few bounds available that enable one to take limits with respect to the norm of . We establish three scenarios where it is possible to do so:

When the restriction is measured according to a Sobolev space of negative index, we determine the complete range of indices for which such a bound exists.

Among functions where vanishes on to order , the restriction of defines a bounded operator from (this subspace of) to provided .

When there is a priori control of in a space , , this implies improved regularity for the restrictions of . If is large enough, then even can be controlled in terms of and alone.

The proofs are based on three main tools: the spectral synthesis work of Y. Domar, which provides a mechanism for approximation by “convolving along surfaces in spectrum”, a new bilinear oscillatory integral estimate valid for ordinary  functions, and a convexity-type property of the quantity  as a function of  and  that allows one to employ the control of .

1. Introduction

1.1. Overview of the derivative restriction problem

Questions regarding the fine properties of the Fourier transform of a function in have long played a central role in the development of classical harmonic analysis. While the Hausdorff–Young theorem guarantees that for , the Fourier transform of belongs to its dual space , it does not provide guidance on whether may be defined on a given measure-zero subset . The canonical question of this type, originating in the work of Stein circa 1967, is to find the complete range of pairs  for which the inequality

holds true. The problem was solved in the case  in Reference 8 and remains an active subject of research in higher dimensions (e.g., Reference 5Reference 13Reference 14).

In this paper we investigate the possibility of defining the surface trace of higher order gradients of the Fourier transform of an  function, with a focus on uniform estimates in the style of Equation 1. Let  be a closed smooth embedded -dimensional submanifold of . Assume that the principal curvatures of  are nonzero at any point. Let  be a compact subset of  and let  be a natural number. We consider as a model problem the inequality

Here and in what follows the Fourier transform has priority over differentiation: we first compute the Fourier transform and then differentiate it. We choose the standard Hausdorff measure on  to define the -space on the left hand side. The notation  signifies that the constant in the inequality may depend on the choice of , but should not depend on . We restrict our study to the case of  instead of  with arbitrary  on the left hand side, because the Hilbert space properties of make this case more tractable. In fact, the range of all possible  in Equation 1 when  is described by the classical Stein–Tomas theorem (established in Reference 25 and Reference 21).

Unfortunately, inequality Equation 2 cannot hold true unless . To see that, consider the shifts of a function , in other words , where  is a fixed point in . If we plug  into Equation 2 instead of , the norm on the left hand side will be of the order , whereas the quantity on the right will not depend on .

The next question along these lines is: what modifications can be made so that Equation 2 becomes a true statement for ? Since the original inequality Equation 1 is shift-invariant, we seek translation invariant conditions for . This rules out natural candidates such as requiring .

One possibility is to relax the desired local regularity from to a Sobolev space of negative order. Consider the inequality

Here  is an arbitrary compactly supported smooth function (the constant in the inequality may depend on it). The parameter  is a nonnegative real, and  is the -based Bessel potential space. Whenever Equation 3 holds, there is a trace value for in for all .

One might guess that the inequality Equation 3 gets weaker as we increase , opening the way to define the trace of  on  with an increasingly large range of . This is indeed the case. The case  in Equation 3 was considered by Cho, Guo, and Lee in Reference 6. They observed Sobolev space trace values of for with going up to the sharp exponent dictated by the Fourier transform of a surface measure.

There are two parameters that appear frequently as bounds in our arguments:

Where it occurs later on, we also use the standard notation for the dual exponent to .

Proposition 1.1 (Corollary of Theorem in Reference 6).

Let . The inequality Equation 3 is true if and only if

For fixed and with , that means . In the case , the case  is also permitted if .

The parameter  is related to the “surface measure extremizer”. When condition Equation 7 does not hold, Proposition 1.1 fails by testing its dual statement against a surface measure on . The parameter , and its role in condition Equation 8 are similarly associated with Knapp examples.

In odd dimensions there is an endpoint case , where inequality Equation 3 is true for . This is stated more precisely in Corollary 7.8 below. The proof of that bound is more direct than most of our other arguments (in fact it is nearly equivalent to the dispersive bound for the Schrödinger equation) and it is completely independent; see Proposition 7.7.

The paper contains two proofs of Proposition 1.1. First, it is a special case of the more extensive Theorem 1.16, whose proof is presented as Section 4. Then we also show in Subsection 7.3 how to derive Proposition 1.1 from the results of Reference 6. To be more specific, one can interpolate between the results of Reference 6 for and the Besov space bound in Proposition 7.7 for , to obtain the full range of Theorem 1.1.

If one is determined not to weaken the norm in Equation 2, it is necessary to consider belonging to an a priori narrower space than . We introduce the main character.

Definition 1.2.

Let  be a closed smooth embedded -dimensional submanifold of , , and . Define the space  by the formula

Define  to be simply . The first nontrivial space  will often be denoted by .

The symbol  denotes the Schwartz class of test functions. We note that in the definition above, we do not need any information about . In fact,  may be an arbitrary closed set. The restriction  is taken so that the Schwartz class is dense in , though one could replace closure with weak closure in the case  if needed. These generalities will not arise in the present paper. From now on we assume that  is a closed smooth embedded -dimensional submanifold of  with nonvanishing principal curvatures.

It will turn out (see Theorem 1.6 below) that for a certain range of and , the space contains precisely the functions whose Fourier transform vanishes on to order . We take advantage of the additional structure of the domain to formulate a second adaptation of inequality Equation 2, this time with the trace of still belonging to :

One might expect that a similar statement with the norm replaced by a weaker Sobolev norm will admit a larger range of , that is,

However at this point in the discussion it is not clear why Equation 10 should be true outside the range established in Proposition 1.1, or why Equation 9 should be true at all. Given a generic function , its Fourier transform is not differentiable even to fractional order. We have reduced the obstruction somewhat by seeking derivatives of only at the points , and by specifying a substantial number of its partial derivatives via the assumption . Nevertheless, values of alone do not uniquely determine , nor are they known to shed much light on the behavior of in a neighborhood of .

Theorem 1.4 below finds the complete range of for which an gradient restriction Equation 9 is true. In particular, the range is nonempty when . The range of permitted in Equation 10 is also sharp in the same way as Proposition 1.1 and the results in Reference 6. The range of we obtain here is much larger than what is true in the context of Proposition 1.1, but most likely not optimal due to some complications with linear programming over the integers.

The case of Theorem 1.4 shows that an a priori assumption leads to nontrivial bounds on . In fact there is a larger family of bounds for trace values of , and one can begin the bootstrapping process with a much milder assumption instead of requiring it to vanish. We explore these generalizations in Proposition 1.11, Theorem 1.12, and the related discussion. The inequality which takes the place of Equation 10 has the form

(the right hand side may be infinite). Remarkably, there are cases where this statement holds with only an norm on the left side. In Corollary 1.13 we find a sizable range of indices that admit a local bound on the gradient of ,

The spaces  that arise in Definition 1.2 are not a new construction. They appeared in Reference 12 (see Proposition in that paper) and Reference 11 where the authors investigated the action of Bochner–Riesz operators of negative order on these spaces. They arose in Reference 24 in connection with Sobolev-type embedding theorems. We describe this development in Section 2.

In fact, the spaces  played the central role in the study of the spectral synthesis problem in the 1960s and 1970s. We stress the work of Domar here (e.g., Reference 7) and will rely upon it in Section 3.

It is worth noting that the main inequality used to derive Equation 9 and Equation 10 is valid for all functions in , not just those whose Fourier restriction vanishes on . The formulation of this inequality, which may be of independent interest, is given in Equation 29 below and the sharp range of for which it holds is found in Theorem 1.16.

1.2. Statement of results

It follows from Definition 1.2 that the spaces  get more narrow as we increase :

The final space can be defined as the closure in  of the set of Schwartz functions whose Fourier transform vanishes in a neighborhood of . We claim that  when  is sufficiently large (i.e., the chain of spaces stabilizes). Here is the precise formulation.

Proposition 1.3.

We have  provided  and . If , this is true provided .

For the case , this proposition was proved in Reference 7, and the proof works for arbitrary  (except for, possibly, , which we do not consider here). The theorem is sharp in the sense that  provided  (see Theorem 1.6 below).

Theorem 1.4.

The inequality Equation 9 is true if and only if , or equivalently .

More generally, inequality Equation 10 is true for and

where the notation indicates the smallest integer greater than or equal to the enclosed value. This covers the entire range . When and the value is also permitted.

Remark 1.5.

The , endpoint case is handled in Corollary 7.8 below, with inequality Equation 10 holding for all .

The first claim in the theorem above is an “iff” statement. Usually, the “if” part is much more involved than the “only if” one. In fact, the “only if” part of Theorem 1.4 is proved with the standard Knapp example. Some of the other theorems in the paper will have a richer collection of “extremizers”. Moreover, one and the same “extremizer” may prove sharpness of several related estimates. We collect the descriptions of such type “extremizers” (and thus, the proofs of the “only if” parts) in Section 6.

Theorem 1.4 says that the operator

acts continuously from the space  to  when , or from to for some combinations of with . This allows us to define a new space

which consists of all  functions for which the ( or ) traces of all partial derivatives of order  vanish on . Note that  is formally defined on , and so on, thus we have vanishing of lower order derivatives as well. We also note that in the case when  is compact, one does not need to use the intersection in Equation 11 and may simply write . It follows from definitions that . In fact, the two spaces must coincide. This looks like a trivial approximation statement, however, we do not know a straightforward proof.

Theorem 1.6.

For any , the spaces  and  coincide with being regarded as a map from to . This occurs when .

For any , the spaces  and  coincide with being regarded as a map from to for the same range of as in Theorem 1.4. This occurs when , or when .

Remark 1.7.

Since acts nontrivially on the Schwartz functions contained in , it follows that in this range of . Thus, Proposition 1.3 and Theorem 1.6 completely classify the spaces , modulo some details about the optimal target space for .

Remark 1.8.

In the papers Reference 11 and Reference 12, the condition on the unit sphere” was understood in the sense of  traces.

Theorem 1.4 covers many combinations that are forbidden in Proposition 1.1 by demanding that vanishes to order on . We now introduce a family of statements which assume only smoothness of instead of vanishing. Bessel spaces already appear on the left hand side of inequality Equation 3, so it is reasonable to use the same scale to describe the smoothness of .

Definition 1.9.

Let  be a natural number, let  and  be nonnegative reals, and let . We say that the higher derivative restriction property holds true if for any smooth compactly supported function  in  variables, the estimate

holds true for any Schwartz function .

The HDR property resembles a Dirichlet-to-Neumann bound for Fourier transforms in the sense that regularity of along the surface implies a certain degree of improved regularity in the transverse direction. It is notable that the restriction does not uniquely determine or the values of anywhere else in , so the inequality Equation 12 must hold uniformly for all functions whose Fourier transforms coincide with on .

Remark 1.10.

A complete generalization of Theorem 1.4 would include a priori estimates on for . We consider only the case above for relative simplicity of notation.

Proposition 1.11.

If  holds true and , then

where the numbers  and  are defined by Equation 4 and Equation 5, respectively. In the case , equality in Equation 15 may also occur.

The sufficient conditions we are able to provide for the  inequalities do not always match the necessary ones listed above. Roughly speaking, they get close to necessary conditions when  or  is relatively small and there is a gap if  and  are both large. By “getting close to necessary conditions” we mean that the nonsharpness comes only from our limitation of working with integer . The sufficient conditions we are able to obtain are rather bulky (this is again due to “integer arithmetic”). They are formulated in terms of certain convex hulls of finite collections of points in the plane. Since we need to introduce more notation before formulating the strongest available statement, we refer the reader to Theorem 5.21 in Section 5 for the details and state a representative subset of the results here.

Theorem 1.12.

Let  and . If

then  holds true provided Equation 13Equation 14Equation 15Equation 16, and Equation 17 are satisfied. If Equation 18 does not hold, then  holds true provided Equation 13Equation 17 are satisfied as well as the inequality

Here and in what follows,  is the upper integer part of a number, i.e., the smallest integer that is greater than or equal to the number; the notation  denotes the lower integer part of a number, i.e., the largest integer that does not exceed the number:

The , cases of Theorem 1.12 illustrate its ability to extract derivatives of in all directions when only regularity along is assumed.

Corollary 1.13.

Suppose for some integer , and let . Then

for any Schwartz function , compact subset , and smooth cutoff that is identically on .

In Section 6.4 we construct a translated Knapp example to show that the lower bound for is sharp.

The property has a dual formulation in terms of the Fourier extension operator. We denote the Lebesgue measure on  by .

Corollary 1.14.

Suppose holds true and (e.g., if the conditions of Theorems 1.12 or 5.21 are satisfied). Then for each , multi-index with , and smooth compactly supported , there exist and such that

and furthermore

Conversly, if for any compactly supported smooth function , for any , and for any  there exist  and  such that Equation 22 and Equation 23, then holds true (we still assume .

When is the paraboloid , the Fourier extension operator doubles as the linear propagator of the Schrödinger equation on with . In this context Corollary 1.13 implies a time-weighted scattering property for solutions of the Schrödinger equation.

Corollary 1.15.

Let be an integer . Given with Fourier support in the unit ball, there exists a function , also with Fourier support in the unit ball, such that

and

If is even, the result holds for provided the exponent of in Equation 24 is strictly less than .

Finally, we present the main analytic tool used in our proofs of inequalities. We will formulate it in local form: now  is a graph of a function on  rather than an arbitrary submanifold.

Let  be a neighborhood of the origin in . Let  be a -smooth function on  such that  and . We also assume that the Hessian of  at zero does not vanish,

Moreover, we assume that the gradient of  is sufficiently close to zero and the second differential is sufficiently close to :

The function  naturally defines the family of surfaces

We also take some small number  and consider the set .

We will be using Bessel potential spaces adjusted to these surfaces. Now we will need the precise quantity defining the Bessel norm. It is convenient to parametrize everything with . For  and a compactly supported function  on  (for some fixed ), define its -norm by the formula

The symbol  denotes the Fourier transform in  variables, and we have used the notation . We will also use the homogeneous norm

Since all our functions are supported on , this norm is equivalent to the inhomogeneous norm Equation 27 when . We will often use another formula for the homogeneous norm:

The constant  may be computed explicitly, however, we do not need the sharp expression for it.

Let  and  be integers between  and , let  be real, and let . Let also  be an arbitrary  function supported in . We are interested in the differentiated restriction inequality

So we compute the Fourier transform of an  function, calculate its derivative with respect to the th coordinate, compute the  norms of traces of this derivative on the surfaces , and then differentiate  times with respect to . We use the variable  for points in  on the spectral side and  for points in  decoding points on  (for example,  is quite often equal to ).

The crucial statement which unlocks most results in this paper is a sharp characterization of when Equation 29 is valid.

Theorem 1.16.

Let satisfy the assumptions above, and let . Inequality Equation 29 is true for the combination of if and only if

and the inequality Equation 31 is strict if .

In the case , the estimate Equation 29 reduces to the classical Stein–Tomas bound. Though our proof will follow the scheme of the fractional integration method (see, e.g., Reference 18, 11.2.2), both the  case and the interpolation of operators in the proof of Equation 29 require significant new efforts.

Note that Proposition 1.1, except for the endpoint case , , follows from choosing in Theorem 1.16 and applying the localization argument given in Subsection 7.1 below. It is not clear to the authors whether one can derive the full statement of Theorem 1.16 from the results of Reference 6 or from Proposition 1.1 (which correspond to the cases  and , respectively).

Our approach using Theorem 1.16 seems to be a different strategy from the one in Reference 6. It makes possible the extensive family of conditional restriction estimates proved in Theorem 1.12 and its corollary. The method also allows us to work with Strichartz estimates, i.e., consider the larger scale of mixed-norm Lebesgue spaces on on the right hand side of Equation 29.

The organization of the paper is as follows. Section 2 is a brief statement of some problems in the literature that provided motivation for the current work. Section 3 tackles the functional analysis of spaces . Proposition 1.3 is proved here, and Theorems 1.4 and 1.6 are reduced to corollaries of Theorem 1.16. The proof of Theorem 1.16 takes up the entirety of Section 4. In Section 5 we present the argument deriving an expanded version of Theorem 1.12 from Theorem 1.16. Section 6 contains examples demonstrating the necessity of conditions in Theorems 1.4 and 1.16 as well as Proposition 1.11. The final section contains miscellaneous technical results: an argument for working locally on , a Stein–Weiss inequality, and statements related to the , endpoint case of Proposition 1.1 in odd dimensions.

2. Precursors to the current work

Fredholm conditions.

Functions whose Fourier transform vanish on a compact surface in , and in particular on a sphere, arise in the study of spectral theory of Schrödinger operators . It is well known that the Laplacian has absolutely continuous spectrum on the positive halfline , and no eigenvalues or singular continuous spectrum. If can be approximated by bounded, compactly supported functions in a suitable norm (for example suffices when ), then is a relatively compact perturbation of the Laplacian and may have countably many eigenvalues with a possible accumulation point at zero. If is real-valued, then is a self-adjoint operator whose eigenvalues must all be real numbers as well.

It is not immediately obvious how the continuous spectrum of relates to that of the Laplacian, and whether any eigenvalues are embedded within it. An argument due to Agmon Reference 1 proceeds as follows. Suppose is a formal solution of the eigenvalue equation for some . Then

from which it follows that the imaginary parts of and must agree. The former is clearly zero since is real-valued. The latter turns out to be a multiple of , where is the sphere . Hence vanishes on the sphere of radius .

It is not surprising that the Fourier multiplication operator might have favorable mapping properties when applied specifically to , whose Fourier transform vanishes where is greatest. Bootstrapping arguments using Equation 33 show that , even if it was not assumed a priori to belong to that space.

Viewed another way, the eigenvalue problem is an inhomogeneous partial differential equation where the principal symbol is elliptic. The Fredholm condition for existence of solutions is that should be orthogonal to the nullspace of the adjoint operator . As we will argue later in Subsection 3.1, this nullspace consists of all distributions whose Fourier transform acts as a linear functional on . Thus satisfies the Fredholm condition precisely if .

The analysis in Reference 1 is carried out in polynomially weighted and applies to a wide family of elliptic differential operators . The main nondegeneracy condition is that the gradient of does not vanish on the level set . Similar arguments in Reference 12, Reference 15, and Reference 11 are carried out (for ) in and related Sobolev spaces for various ranges of . Curvature of the level sets of is a crucial feature in these works, as it is in the present paper.

Sobolev-type inequalities

We start with the classical Sobolev embedding theorem

For , it follows from the Hardy–Littlewood–Sobolev inequality. In the limiting case , the Hardy–Littlewood–Sobolev inequality fails, however, as it was proved by Gagliardo and Nirenberg, the Sobolev embedding holds. This happens because the space

is strictly narrower than  (they are even nonisomorphic as Banach spaces). Later, it was observed that there are many similar inequalities where  may be replaced with a more complicated differential vector-valued expression (see Reference 4Reference 19, and the survey Reference 20).

In Reference 24, the second-named author studied the anisotropic bilinear inequality

(such a type of inequalities were used in Reference 17 for purposes of Banach space theory). Here  is the anisotropic Bessel potential space equipped with the norm

the symbols  and  denote complex scalars, and  and  are partial derivatives with respect to the first and the second coordinates correspondingly. It appeared that Equation 34 holds even in the cases where the differential polynomials on the right hand side are not elliptic, however, this may happen only in the anisotropic case . This leads to the natural conjecture that the inequality

might hold true. We are especially interested in the case where the operator on the right hand side is nonelliptic, that is, . Assume this is so. Similar to the classical proof of the Sobolev embedding theorem, one may express  in terms of  using a certain integral operator. This will be a Bochner–Riesz-type operator of order  with the singularity on the curve

Note that this curve is convex outside the origin. Application of the Littlewood–Paley inequality and homogeneity considerations (see Reference 24) reduce Equation 35 to the case where the spectrum of  lies in a small neighborhood of a point on . So, by the results of Reference 2, the inequality Equation 35 is true if , , and . Moreover, one may construct examples to show that the conditions  and  are necessary.

Note that the Fourier transform of the function  vanishes on , which is a smooth convex curve in the plane (with, possibly, a singularity at zero). Thus, we need to analyze the action of a Bochner–Riesz-type operator on the space  with . It appears that passing to a narrower space allows one to get rid of the condition . This work was done a half year later in Reference 11.

Theorem 2.1 (Reference 24 and Reference 11).

The inequality Equation 35 holds true if , , and .

3. Study of the spaces

3.1. Description of the annihilator and Domar’s theory

Let  denote the operator of a normal derivative of order  with respect to , ,

The symbol  denotes the normal vector to  at the point . In particular,  is simply the restriction of  to .

There are conjugate operators . We can also form an operator  composed of pure normal derivatives:

This operator also has an adjoint, which maps a vector-valued distribution  with compact support on  to a Schwartz distribution on .

Lemma 3.1.

Let . The annihilator of  in  can be described as

If , the closure is with respect to the weak-* topology of .

Remark 3.2.

Since the distribution  has compact support,  has bounded spectrum. Clearly, if  is not compact, one may construct a function  in the annihilator of  whose spectrum is not bounded. That is why we need to add closure on the right hand side of Equation 36. In the case where  is compact, this is not needed:

The proof of Lemma 3.1 presented below also simplifies in the case where  is compact. The functions  and  may be omitted in this case.

We will need a technical fact to prove Lemma 3.1. It is standard, so we omit its proof.

Lemma 3.3.

For any bounded domain , consider the subspace  of vector-valued functions in  supported in . There exists a linear operator , which is inverse to  in the sense

Proof of Lemma 3.1.

First, we note that since  is a translation invariant space, the set of functions  with compact spectrum is dense in . Consider such a function . It suffices to construct

such that .

Let  be a bounded domain containing the spectrum of . Consider the operator  constructed in Lemma 3.3 and define  (as a functional on ) by the formula

where  is a smooth function on  supported in  that is equal to one in a neighborhood of . We are required to show that , which becomes

for every . Since  in a neighborhood of the support of , we may write

where  is smooth function supported in  that equals one in a neighborhood of . It remains to prove

Since  (recall that  annihilates ), it suffices to show that

which holds true since  by construction of .

Lemma 3.4.

The set

is dense in  if . In the case , this set is weakly dense.

Proof.

The case  had been considered in Reference 7. We repeat the argument for the general case here. Let  be a function in the said annihilator. After applying a partition of unity, we may assume that the corresponding vector-valued distribution  provided by Lemma 3.1 is supported in a chart neighborhood  of a point , as it will only be necessary to sum a finite number of such pieces. We may also suppose that  and

here  is a neighborhood of the origin in  and  is a smooth function such that , (see Subsection 7.1 for details). By our assumptions on the principal curvatures of , the second differential  is non-degenerate on . Consider the operator  that makes  flat:

We use the notation , so  is the last coordinate of  and  is the vector consisting of first  coordinates.

Let  be a compactly supported smooth function on  with unit integral, and let  be its dilations. The function  is also denoted by . Consider the family of operators , , is sufficiently large, given by the rule

Here  is a function that equals one on the support of . It is clear that the  are uniformly bounded as operators on . Lemma  in Reference 7 says that the  are also (uniformly in ) bounded as operators on , and since the dual operators have an identical structure they are also bounded on . By interpolation, the  are uniformly bounded on . Also, since  is an approximate identity,

Thus, for every , , we have . It remains to notice that  maps compactly supported distributions of the form

to the ones for which . Thus, if  is a function with bounded spectrum, then , is generated by smooth , and  in .

Proof of Proposition 1.3.

By Lemma 3.1, it suffices to show that any function  can be approximated by functions in  when , and  when . By Lemma 3.4, we may assume that

It suffices to prove that , where . We may suppose that  is supported in a neighborhood  of a point on . We may also assume Equation 37 and replace normal derivatives by derivatives with respect to :

where  are distributions generated by complex measures on  whose densities with respect to the Lebesgue measure on  are smooth functions. Note that  whenever . Since each function  has smooth density with respect to the Lebesgue measure on , one may use the stationary phase method to compute the asymptotics of  at infinity (see, e.g., Reference 22):

for all  such that  for some  with . Here  is a nonzero oscillating factor with constant amplitude that depends on  and the density of . This shows that

On the other hand, , which for requires , equivalently , contradicting our assumptions. Therefore,  and, thus, . If the contradiction is reached provided .

To show that , we note that the annihilator of the latter space consists of all  functions whose Fourier transform is supported on , recall the Schwartz theorem that any distribution supported on  may be represented in the form  for some , and use the reasoning above.

3.2. Coincidence of and the spaces defined as kernels of restriction operators

We relate the  spaces with restriction operators. Consider a neighborhood  such that Equation 37 holds true. We may redefine  in such a way that

This gives a natural parametrization of  by . We will need the translated copies of :

Note that this definition depends on the choice of .

Consider the restriction operators

Definition 3.5.

We say that the statement  holds true if the  admit continuous extensions as  operators for any choice of , and the norms of these extensions are uniform in  (however, we do not require any uniformity with respect to ).

We say that  is true if  is true and for any choice of  the operators  extend continuously from the domain

to a family of mappings  whose norms are bounded uniformly by .

Remark 3.6.

In the definitions above it is important to be consistent with regard to the construction of local Sobolev norms on . When we discuss , we will define the Sobolev norm by the rule Equation 28 for each particular choice of , , and .

Definition 3.7.

For a fixed , define the set  by the formula

Note that it is unclear whether  is closed in or not.

Lemma 3.8.

Suppose that  has nonvanishing curvature, , and holds. Then, .

Proof.

It is clear that , as contains all Schwartz functions whose Fourier transform vanishes to order on . In fact, thanks to the uniform convergence implied by condition it is even true that . So it suffices to show that . Assume the contrary. By the Hahn–Banach theorem,

By Lemma 3.4, we may assume that  is of the form

Applying the same reasoning as in the proof of Lemma 3.4, we may also assume that has compact support within a chart neighborhood where is the graph of a smooth function , , and .

Recall the “flattening” operator defined in Equation 38. There exists another set of functions such that . If one considers each component of as an element of via the parametrization of , then the components of are constructed from , its gradients (in ) up to order , and the partial derivatives of .

Let  be a compactly supported function on the unit interval whose integral equals one. Consider its dilations  and the formal convolution in the th variable

which is a function in . This function is bounded pointwise by the maximum size of , which is approximately . All of its partial derivatives in the directions are bounded by  as well, because those derivatives act on , not on .

Now define  by the formula

It should be clear that convolution in the direction commutes with operators and its inverse, thus the construction simplifies to

It follows that  in  (in the case  we have weak-* convergence relative to instead), which means that, . On the other hand,

That gives a bound

forcing . This contradicts the original assertion that .

Definition 3.9.

We say that the statement  holds true if the mapping

extends to a bounded linear operator between the spaces  and  for any compact set .

It is explained in Remark 7.3 below that  leads to .

We end this subsection with an analog of Lemma 3.8 for  inequalities. The proof is direct, i.e., does not use duality.

Lemma 3.10.

For any function  such that , where , there exists a sequence  of Schwartz functions such that

Proof.

As usual, we may assume that  and  are supported in a chart neighborhood  of a point . We may also suppose that  and Equation 37, where  is a neighborhood of the origin in  and  is a smooth function such that , (see Subsection 7.1 for details). By Proposition 1.3, in the regime , the set of Schwartz functions whose Fourier transform vanishes on , is dense in , and there is nothing to prove. Let us assume .

We construct the functions  by the rule

where  is a fixed Schwartz function with  and bounded spectrum, and  is a large number such that . The functions  approximate  in  norm, however, their Fourier transforms may have infinite  norms. There is a control on a weaker quantity, namely, Theorem 1.4 (in the case ) says that

for sufficiently large .

Let now , where  is the Domar operator Equation 39. We need to prove two limit identities

The first identity is simple since

by the properties of the operators  (see the proof of Lemma 3.4). For the second identity, we write

Note that  convolves the restriction to  of the Fourier transform of the function with  (see Equation 39). Thus, the first summand tends to zero by the approximation of identity properties (and since ), and the second summand is bounded by  by formula Equation 43.

3.3. Proofs of “if” part in Theorems 1.4 and 1.6

Since  leads to , Theorem 1.4 follows from the lemma below.

Lemma 3.11.

The statement  holds true if . For every there exists such that is true. When and , the value suffices. Finally, in odd dimensions holds for .

Proof.

Consider the case . It suffices to prove the bound

By definition of , we may assume that  is a Schwartz function. Then, the function  given by the rule

is smooth. We need to prove , and for that, it suffices to show an inequality and several equalities.

The inequality is

which follows from Theorem 1.16 (take , in the role of , and notice that Equation 30 is satisfied automatically, Equation 32 is equivalent to , and Equation 31 follows from Equation 32 in this case).

The equalities are

Indeed, we use the product rule:

and notice that in each scalar product on the right hand side, one of the functions is identically zero since either  or .

It remains to combine Equation 45Equation 46, and the Taylor integral remainder formula to complete the proof in the case .

When , the choice of , , and is no longer available in Theorem 1.16. Suppose . In order to use the product rule argument above, one must set , and it is desirable to keep as small as possible since is a prominent lower bound for . We can apply Theorem 1.16 with , (note that  and  here), and , then follow the above steps for

to conclude that for all , and every lower order derivative vanishes at because . Thus . Furthermore, is assumed to vanish at for each . The Taylor remainder formula and the Minkowski inequality show that , and we previously set .

When and , it is also permissible to apply Theorem 1.16 with , , and . In the endpoint case , Corollary 7.8 below directly states that for and .

Proof of Theorem 1.6.

Clearly,  whenever is suitably defined as a map from  into . To prove the reverse embedding, it suffices to show that , since by Lemmas 3.11 (with the value of specified there) and 3.8, we have  for these choices of  and .

We first consider the case , . By Definition 3.7, we need to prove

where the function  is defined by Equation 44. By the Taylor integral remainder formula and Equation 46, we simply need to show a slight refinement of Equation 45:

Note that Equation 45 holds for all . By approximating any such by Schwartz functions, we see that is also continuous in . If , the computation in Equation 47 with shows that , and for , the norm on the right hand side is zero.

The remaining case is essentially the same. This time is continuous for all , the lower order derivatives vanish at for , and finally if . Thus and the rest of the integrations are the same as in Lemma 3.11.

3.4. Proof of Corollary 1.14

Let be the vector space of functions equipped with the norm . This space contains all functions in the Schwartz class, and convergence with respect to the Schwartz class topology implies convergence in the norm of . Thus every bounded linear functional on belongs to the class of distributions .

Lemma 3.10 asserts that the Schwartz class is dense in . To show completeness of , observe that by the case of Proposition 1.1 (i.e., by Reference 6), if in , then there exists so that in . Every Cauchy sequence in has convergent to a limit in the stronger topology of , and the limit must be as well.

We may identify with the ordered pair . This gives an isometric embedding of into . Its image is closed, so the Hahn–Banach theorem implies that every linear functional extends to a functional on . Using Parseval’s identity there exists and with norms bounded by that of and which satisfy

The defining property of expressed in Equation 12 is that the linear map is continuous from to . The dual map, taking therefore is bounded from to , with elements of described as above.

Remark 3.12.

Due to the use of the Hahn–Banach theorem in this argument, we do not have a construction for and in Corollary 1.14. In fact, it is not proved here that these two functions can be chosen to depend linearly on .

4. Proof of Theorem 1.16

4.1. Pointwise estimates of the kernel

The quadratic inequality Equation 29 is equivalent to its bilinear version

We denote the bilinear form we estimate by  and its kernel by :

We also recall the notation

for .

Proposition 4.1.

The kernel  defined in Equation 50 satisfies the bound

Remark 4.2.

One can track the numerology of conditions Equation 30Equation 31, and Equation 32 from this proposition. The boundedness of  on  is equivalent to the uniform boundedness of . The right hand side of Equation 51 is uniformly bounded exactly when these three conditions hold for  (they reflect the behavior of the kernel along the directions , , and , respectively).

In the case , the inequality Equation 51 follows from the standard Van der Corput lemma, because in this case

the angular brackets denote the standard scalar product in .

So, we assume  in what follows. We start with explicit formulas for the kernel :

We want to pass to the dyadic version of the Bessel seminorm, namely,

based on the formula

The function  is supported outside zero and nonnegative.

We substitute formula Equation 53 into Equation 52 and split  into a dyadic sum:

where

Lemma 4.3.

For any , and any ,

Proof.

The integral in the formula for  may be thought of as the -dimensional Fourier integral:

It suffices to prove the inequality

We represent the function we apply to the Fourier transform as a product of two functions

By the Van der Corput lemma, the Fourier transform of the first function is uniformly (in ) bounded by the right hand side of Equation 54. It remains to notice that the Fourier transform of the second function is a complex measure whose total variation is bounded uniformly in . This is easiest to see by making a linear change of variables from  to .

Define the number  by the rule

Lemma 4.4.

For any , and any ,

Proof.

Let . It suffices to prove the estimate

We introduce new variables  and disregard oscillations in the  variable:

It remains to prove

The function  is uniformly (with respect to  and ) bounded in any Schwartz norm, so its Fourier transform is an -function whose norm is bounded independently of  and . Thus, it suffices to prove

uniformly in . This inequality is trivial if , so we assume the quantity  is sufficiently large. We represent the nonlinear part of the phase function as

where

since . Note that

when , is sufficiently large, and . In particular, the Hessians of the functions in the family  take the form  and are uniformly invertible. By similar reasons, the functions in the family  are uniformly bounded in any Schwartz norm. The version of Littman’s lemma from Reference 7 leads to Equation 56.

Proof of Proposition 4.1.

We use Lemmas 4.3 and 4.4 (the case  has already been considered, so we assume  here):

4.2. Interpolation

To prove the “if” part of Theorem 1.16 for the case , we will have to work with “slices” of the kernel . For any  and , define the kernel  by the formula

Defining the bilinear forms  accordingly

here  and  are functions on . Proposition 4.1 now may be restated as

Lemma 4.5.

For any ,

Proof.

Let  and  be the Lebesgue measures on the hyperplanes

Then,

if we interpret  and  as functions of  variables that do not depend on the last coordinate. With this formula in hand, we may re-express :

Therefore, it suffices to prove the bound

We postulate the inequality

The space on the left is the Lorentz space; see Reference 3 for definitions. Inequality Equation 60 immediately leads to Equation 59:

We are required to prove Equation 60. Let us denote the operator we want to estimate by :

By the Plancherel theorem,

By the Van der Corput lemma,

The real interpolation formulas (see Reference 3, §5.3)

lead to the inequality

which is exactly Equation 60.

Interpolation between Equation 57 and Lemma 4.5 leads to the inequality

for . Let us restrict our attention to this case for awhile. To finish the proof of Theorem 1.16, we invoke a version of the Stein–Weiss inequality (Theorem 7.4 in Subsection 7.2 below):

provided  and  satisfy the requirements of Theorem 7.4. The inequality Equation 30 leads to , the requirement Equation 31 leads to  (with the same exclusion of the endpoint case if ), and Equation 32 gives . The case  and  is impossible ( is negative in this case). The “if” part of Theorem 1.16 is proved in the case .

To deal with the remaining case, we start from the estimate

which follows from the representation Equation 58; we use the trivial inequality

We interpolate this bound with Lemma 4.5:

We invoke Theorem 7.4 with  and . Since , the condition  is stronger than . The condition  is exactly Equation 31. The condition  also follows from it:

The “if” part of Theorem 1.16 is now proved.

Remark 4.6.

Note that we did not use that  or  are integers provided we define our bilinear form by Equation 52.

4.3. Strichartz estimates

With the same method as in the previous section, we can get a collection of sharp (up to the endpoint) Strichartz estimates. For that we need the mixed norm spaces :

We also use Theorem 7.6 here in order to work with the cases  as well. This provides some new information even in the case  considered in Reference 6. The cases  were excluded in that paper and it is not clear whether the methods of Reference 6 work in this situation.

Theorem 4.7.

The inequality

holds true if

(1)

and

;

;

with equality permitted if ;

with equality also permitted if or ;

(2)

and

;

;

;

.

The proof is a direct application of Theorems 7.4 and 7.6. Consider the case  Set , , and (that is, the value of in those theorems is replaced by ). We note that the conditions of Theorem 7.4 can be summarized as , , and . When , combinations with are also accepted, and when the case , is excluded.

The three conditions stated in the case are equivalent to , , and , respectively. The three conditions stated in the case are equivalent to the conditions in Theorem 7.6, namely , , and .

Consider the case  and set , , and  in the same sense as above. Since , the condition  is stronger than . The condition  is equivalent to  with equality permitted if . In the case , the requirement  is rewritten as . It also follows from . The condition  arising in the case  follows from the same inequality.

5. Robust estimates

5.1. Introduction to “numerology”

Remark 5.1.

We are mostly interested in the case  in Equation 12. We claim that in the subcritical case , the second term on the right hand side of this inequality is unnecessary. If  and  is true, then a simpler inequality

also holds true. Indeed, if  is true, then Equation 15 and Equation 17 are valid. However, in this case, these conditions are also sufficient for Equation 63 to be true (see Theorem 1.16 and Figure 1).

Remark 5.2.

In the supercritical case , the condition Equation 15 follows from Equation 17 since  in this case (see Figure 1 as well). Note also that Equation 16 is equivalent to

This inequality, in its turn, leads to Equation 13 provided  (which is true by Equation 17). Thus, in the case , the conditions in Proposition 1.11 are reduced to Equation 14, Equation 64, and Equation 17.

In our proof, the parameters  and  will be varied, however,  and  will be steady. It appears convenient to draw diagrams of admissible pairs . We have already drawn such a diagram for the case  (Figure 1). For our first attempt to the “numerology”, we neglect the integer nature of  and imagine this parameter is real positive. We have three inequalities in the subcritical case: , Equation 15, and Equation 17. The cases of equality correspond to lines on the diagram, and all three inequalities are satisfied inside the domain bounded by the bold broken line. We also note that the lines  and  intersect at the point , which we denote by .

Now we pass to the “supercritical” case . We need to draw two additional lines  and

which correspond to Equation 14 and Equation 16, respectively. The structure of the domain of admissible parameters will depend on the mutual disposition of these two lines and the line . Before we classify the cases of disposition, we note that the line Equation 65 passes through . There is one more nice point lying on it: the point . We will consider the cases  and  separately.

Case ().

In this case, the condition Equation 14 is unnecessary; it follows from Equation 17 and . This case, in its turn, is naturally split into subcases  (see Figure 2, note that the broken line has a nontrivial angle at ) and  (see Figure 3), note that Equation 16 follows from Equation 17 when .

Case ().

In this case, there will be three subcases:  (if this inequality turns into equality, then  passes through the point ), , and . In the first case, the condition Equation 14 is unnecessary (see Figure 4). In the second case, all the conditions are required (see Figure 5). In the third case, the condition Equation 16 is unnecessary (see Figure 6).

5.2. Convexity properties of the function

It is useful to consider the expression

as a function of the parameters  and . We always assume  is a nonnegative integer and  is a nonnegative real. Since we will be working with points in the -plane, we will give names to some regions there.

Definition 5.3.

Let , and  be fixed. The domain

is called the friendly region. The domain where  is called the subcritical region. The set of all points  such that  holds true is called the -domain.

If  is an arbitrary point in the  plane,  will usually denote its -coordinate, and  will denote its -coordinate.

Remark 5.4.

The -domain lies inside the friendly domain (by Proposition 1.11).

Lemma 5.5.

For any  and , there exists a constant  such that the inequality

is true provided  lies in the friendly region and  for .

We will need an “algebraic” lemma that will link the quantities , , and  together.

Lemma 5.6.

For any  such that , there exist coefficients  such that

for any function  and any .

Proof.

First, by the Newton–Leibniz formula,

Thus, it is clear that

is a linear combination of all the other terms in the identity Equation 66. The only nontrivial question is why does the term

have coefficient . For this we observe that the binomial coefficients that appear in Equation 66 once the Newton–Leibniz formula is applied are the same ones that arise in the trigonometric identity

The result follows by evaluating the trigonometric sum at .

Proof of Lemma 5.5.

Lemma 5.6 says that

By the Cauchy–Schwarz inequality, the first summand on the right can be estimated by . All the remaining terms are bounded by , provided Equation 29 holds true with for any . By Theorem 1.16, this holds exactly when

The first list of conditions turns into . So, the first and the third conditions are fulfilled inside the friendly region. The second list is reduced to .

Corollary 5.7.

For any  and , there exists a constant  such that the inequality

is true provided  lies in the friendly region and  for .

Lemma 5.8.

Let  be a finite sequence and let . Assume that

, and , . Then, .

Proof.

Consider the sequence , . Its terms satisfy the inequalities

and . In particular,  is convex on . We also subtract the linear function  from it:

The sequence  is convex on , equals zero at the endpoints  and , and also satisfies the inequality . Thus,  (otherwise, , which contradicts the convexity of  on the interval ). Therefore, , and finally, .

Remark 5.9.

In fact, we have proved that  for any .

Remark 5.10.

Using the homogeneity, one can replace the assumptions of Lemma 5.8 by

, and , for some positive constant . Then, .

Corollary 5.11.

The -domain is convex in the sense that if  is a convex combination of  and  (we assume ), and the latter two points belong to the -domain, then the former point lies in it as well.

Proof.

Consider the line passing through our three points. Let  be all the points with integer first coordinates lying on the segment connecting  and  (we enumerate the points in such a way that the -coordinate increases with the index). Consider also the sequence

By Corollary 5.7, this sequence satisfies the inequality

By the assumption, . Thus, by Lemma 5.8 with  (in the light of Remark 5.10),  is bounded by . In particular,  belongs to the -domain.

Corollary 5.12.

Let  be a point with natural -coordinate lying in the intersection of friendly and subcritical domains. Suppose that the point  lies on the segment , has natural first coordinate , and lies in the friendly domain. If , then  lies belongs to the -domain.

Proof.

The proof of this corollary is very much similar to the proof of the previous one. We consider all the points on the segment  that have integer first coordinates and lie inside the friendly domain. Suppose the leftmost of them has first coordinate ; let us call our points  (so, ). We also add the point  to our sequence and consider the numbers

These numbers satisfy the inequality Equation 67 for . Moreover, Corollary 5.7 provides the inequality

Note that  since  and  since  lies in the friendly domain and . At the endpoint , we have the inequality

since  lies in the subcritical part of the friendly domain (this inequality is the case  in Theorem 1.16). Thus, . Clearly, . So, Lemma 5.8 says all the points  belong to the -domain. In particular,  does.

Our general strategy will be to apply Corollary 5.12 to the points  close to the point . This will enable us to obtain “almost extremal points” of the region; after that, we will apply Corollary 5.11 to pass to convex hulls. Before we pass to the cases, we explain the obstructions that prevent us from proving the sufficiency of the conditions in Proposition 1.11. They are of two types. First, we are able to work with points whose first coordinates are integers only. However, in the general case, the extremal points of the domain of admissible parameters need not necessarily have integer first coordinates. So, we cannot prove (and even formulate)  for them. This makes the convex hull we obtain smaller (we are able to reach only some “integer” points close to the extremal points) than it should be. The second obstruction is more severe. The problem comes from the inequality  in Corollary 5.12. That restricts our “extremal points” from having too large -coordinate, roughly speaking, their -coordinates should satisfy , if we want to apply Corollary 5.12. This will result in a considerable gap between our results and the conditions listed in Proposition 1.11 in the case when .

Now we pass to the cases.

5.3. Statement of results by cases

Case ().

Our reasonings are illustrated by Figure 7. Clearly, here we are interested in the case  only (because if  and  lies in the  domain, then  automatically). We consider the point  and assume  lies in the friendly region, that is, . We draw a segment that connects  with  (it is the slant punctured segment on Figure 7). It crosses the line  at the point . We apply Corollary 5.12 to the points  as  and  as  and obtain the theorem below.

Theorem 5.13.

Let  and let . Then,  holds true provided

Case (, ).

Our reasonings are illustrated by Figure 8. This case is simpler than the previous one. We only need  here. In this case, if  lies on the vertical punctured segment, then it is an average of  and a point inside the intersection of the friendly domain with the subcritical domain. Thus, Corollary 5.12 leads to the theorem below.

Theorem 5.14.

Let , and let . Then,  holds true provided

Case ( and ).

Our reasonings are illustrated by Figure 9. We introduce two auxiliary points  and :

We have used two types of the notion “integer part of a number”; see formula Equation 20.

We connect the point  to  and . Since the point  lies in the intersection of friendly and subcritical regions, Corollary 5.12 applied to  in the role of  says that  is true for all pairs  such that ; in other words

Clearly, the same assertion is true for larger  when  is fixed. The situation with the point  is slightly more complicated: it may lie outside the friendly region if its -coordinate is too large. If it is not so (i.e., ), then we may apply Corollary 5.12 to the point  in the role of  and achieve  is true for all pairs  such that ; in other words

We summarize our results.

Theorem 5.15.

Let  and let . If , then  holds true if

If , then  holds true if

Remark 5.16.

If  and  holds true, then .

Case (, ).

This case will be split into many subcases. We will need to construct two sequences of points generated by  and .

The points , , are generated by . Namely,

The point  may be described as the lowest possible point on the line  that lies above the segment  and belongs to the friendly domain. See Figure 10.

Lemma 5.17.

For any , we have . For , all points  lie on the line .

Proof.

The equation of the line  is

To prove the first half of the lemma, it suffices to verify the inequality

when . This may be rewritten as

Clearly, it suffices to prove this inequality for the largest possible . In this case, we arrive at

We estimate the left hand side with , which, in its turn, does not exceed . The first assertion of the lemma is proved.

Similar to the previous reasoning, it suffices to verify the inequality

when , to prove the second assertion of the lemma. This may be rewritten as

which follows from , which is weaker than our assumption . So, we have proved the second half of the lemma.

The lemma says that, among all the points , only those with the indices , , and , may be the extremal points of the accessible domain.

The points , , are generated by  in a similar manner:

We also consider the point  separately:

Remark 5.18.

The point  lies on  if and only if .

Unfortunately, there is no analog of the first assertion of Lemma 5.17 for the points . Here we can only say that for small  the points  lie on the line  and then at some moment they jump to the line . However, this “moment” can happen much earlier than . We can only bound it from above.

Lemma 5.19.

For , all points  lie on the line .

Proof.

Consider the case  first. The equation of the line  is

So, we need to verify the inequality

This may be rewritten as

So, it suffices to prove

We will prove a stronger inequality

which is equivalent to

This may be restated as

which is true under our assumption .

In the other case , we have

so the statement of the lemma is empty in this case (we consider the points  with  only).

Lemma 5.20.

If  is a number between  and  and , then  belongs to the -domain. If  and  belongs to the friendly region, then  belongs to the -domain.

Note that  belongs to the friendly region if and only if .

Proof.

We prove the second assertion; the proof of the first one is completely similar. We consider two cases:  lies on  and above . In the first case, we may apply Corollary 5.12 with  in the role of  and  in the role of . In the second case, we may apply the same corollary with  in the role of  and the point of intersection of the lines  and  in the role of  (the latter point lies above  since  lies above the segment , and thus belongs to the friendly domain).

We finally summarize our results.

Theorem 5.21.

Assume , . The -domain contains the convex hull of points specified below. We always include the points , , and  in our list. The other points are specified in the following table:

This theorem is a straightforward consequence of Lemma 5.20.

We note that the cases  and  are the same for our result (our answer in these cases are given by the last row in the table above, at least when ). However, the forms of the -domain suggested by Proposition 1.11 differ in these cases (see Figures 5 and 6).

Proof of Theorem 1.12.

Since  is assumed to be a nonnegative integer, we have . Therefore, all the points  and  lie on the lines  and . Since , we have  as well.

When , Theorem 5.15 immediately implies that for points whose first coordinate is a positive integer, holds whenever conditions Equation 15 and Equation 16 are satisfied; Equation 68 turns into Equation 16.

When it remains to apply Theorem 5.21 and decode its results.

For the case , the points , , and straddle the intersection of the lines and . The convex hull of these three together with points and contains every point along with . Thus holds provided Equation 14Equation 15, and Equation 16 are all satisfied.

In the case  we have . Thus, this case is described in the intersection of the last row and first column in the table above. We observe that the points and coincide at the location . Then convex combinations of and form a segment of the line , and convex combinations of  and form a segment of the line . It follows that holds provided Equation 14Equation 15, and Equation 19 are satisfied.

6. Sharpness

In this section, we consider the case where  is the paraboloid  as a representative example. We also assume that  is nonnegative.

6.1. Surface measure conditions

Let  be a smooth function of one variable supported in  such that . Consider the functions  defined as

The function  can be written explicitly:

here  is the Lebesgue measure on the paraboloid . It is easy to observe two formulas:

We will also need the functions

Sharpness of Equation 7.

Assume  on the support of . We plug  into Equation 3. The left hand side is bounded away from zero by

As for the  norm, we note that the functions  have disjoint supports, so,

Since the left hand side of Equation 3 tends to infinity as , the right hand side cannot be uniformly bounded. This means Equation 7 holds true if . In the case , we get  instead.

Necessity of  in Theorem 1.4

Add the requirements  for all . Then,  and  belong to  and the same reasoning gives the necessity of  in Equation 9 and Equation 10, which is exactly . As usual, the cases  are permitted if .

Necessity of Equation 15

We plug exactly the same functions  into Equation 12. The -norm on the left hand side and the  norm on the right hand side behave in the same manner as previously. Since we have assumed , there is no summand  on the right hand side.

Necessity of Equation 31

As it was mentioned earlier, the quadratic inequality Equation 29 is equivalent to its bilinear version Equation 49. We work with the latter expression here. The functions  and  will be constructed from the functions  in a slightly different manner from before. To define , we take  that satisfies  and set . For the function , we require  for all  and , and set  (with ). We plug these functions  and  into Equation 49 and use the Newton–Leibniz formula (we assume  on the support of )

On the right hand side, we have

So, the necessity of Equation 31 is proved.

Necessity of  in Theorem 4.7

This is proved in the same manner as in the previous paragraph. One should only replace the formula for the  norm of  with

6.2. Knapp examples

We start with a Schwartz function  with compactly supported Fourier transform and define the functions  by the formula

By homogeneity,

with the caveat that the homogeneous Sobolev norm may already be infinite if .

Necessity of Equation 8

This can be obtained by simply plugging  into Equation 3 and assuming  in a neighborhood of the origin.

Necessity of the condition  in Theorem 1.4

We take  and note that  as well (recall that  is the paraboloid). It remains to plug  into Equation 10 with the same assumption about .

Necessity of Equation 17

Here we plug  generated by  into Equation 12 and note that .

Necessity of Equation 32

This follows from the formula

Necessity of  in Theorem 4.7

One can prove this in the same manner as in the previous paragraph. One should only replace the formula for the  norm of  with

6.3. Pure shifts

We start with a Schwartz function  and consider its shifts in the  direction:

We also assume

and  on the support of . Then  and

Necessity of Equation 6

This follows from the fact that  does not depend on .

Necessity of Equation 13

Note that  does not exceed  (this estimate reduces to the product rule in the case ; the general case follows from the case  by the Cauchy–Schwarz inequality). Comparing the left and right parts of Equation 12, we get Equation 13.

Necessity of Equation 14

We consider the functions  generated by the rule Equation 70 from a function  such that  and  on the support of , and all higher order (up to order ) derivatives of  vanish on . Then,  as well, so, there is no term  on the right hand side of Equation 12. However, on left hand side, we cannot have , but only have growth  since

Thus,  should be bounded if Equation 12 holds, which is exactly Equation 14.

Necessity of Equation 30

Consider a Schwartz function  of  variables such that for any  we have

Let  be generated by Equation 70 from . We plug  into Equation 29. We first compute the “interior” derivative:

Therefore,

Thus, the left hand side of Equation 29 grows at least as fast as , whereas the right hand side does not change. This proves the necessity of the condition Equation 30.

Necessity of condition  in Theorem 4.7

This is obtained by completely the same method in the case . For the case , we can only prove the necessity of the nonstrict inequality . For that we slightly modify the construction above. We consider the function

where the functions  are generated by Equation 70 is a sufficiently large number, and  are randomly chosen signs. Then,

On the other hand, disregarding the choice of the signs ,

provided  is sufficiently large (this number is needed to diminish the influence of Schwartz tails on this almost orthogonality). It remains to choose  with the largest possible quantity on the left hand side and compare the two sides.

Necessity of condition  in Theorem 4.7

This can be obtained by a construction similar to the one described in the previous paragraph, except with functions shifted in the  direction instead of the  direction.

6.4. Shifted Knapp example

We need to modify the classical Knapp construction to get the necessity of Equation 16. We take some sequence  and modify the functions  generated by Equation 69. Now we also shift them:

We require  and do not require the vanishing . The  norms are influenced by scaling but do not depend on the size of the shifts:

Let  be , here  is a smooth function, let us assume it is compactly supported and has nonzero integral. Then,

The latter estimate can be proved via the product rule for the case  and reduced to this case with the help of the Cauchy–Schwarz inequality. Similarly,

So, if Equation 12 is true, then

whenever . We recall  by Equation 13 (the necessity of which is already proved), so, the first term on the right dominates the left hand side when  is sufficiently large. We want to make  as small as possible in such a way that the left hand side is still greater than the second summand on the right. Let

Note that such a choice of  guarantees  by Equation 17 and the assumption . Plugging it back to Equation 73, we get

which, after a tiny portion of algebra and Equation 13, leads to

which is Equation 16.

7. Additional lemmas and supplementary material

7.1. Localization argument

We need to localize the  inequalities and also replace the gradient with a single directional derivative. Namely, we want to reduce  to a collection of statements  defined below. A similar principle works for inequalities of the type Equation 3Equation 9Equation 10 and the proof is completely identical.

Definition 7.1.

Let the numbers  be of the same nature as in Definition 1.9. Let  be a neighborhood of the origin in , let  be a smooth function such that , , and the determinant of the Hessian of  at the origin does not vanish. Further, we assume Equation 26. We say that the statement  holds true if the inequality

holds true for any smooth function  supported in .

Lemma 7.2.

The statement  is true provided the statement is true for any  satisfying the conditions of Definition 7.1.

Proof.

We need to prove Equation 12 with a fixed compactly supported smooth function . We find a smooth partition of unity  on , each function  supported in a small ball  and each  lies in a chart neighborhood of a certain point . For each  fixed, we identify  with the origin of , the tangent plane  with , and get a graph representation for :

where  is a neighborhood of the origin in . If the partition of unity is sufficiently fine, then the function  satisfies Equation 26. We estimate the left hand side of Equation 12 by the triangle inequality

Note that the sum on the right is, in fact, finite. We fix . We are going to use the following algebraic fact: there exists a finite collection of vectors  in  such that any homogeneous polynomial of degree  is a linear combination of the monomials ; moreover, such vectors  may be chosen arbitrarily close to any fixed vector. Since the determinant of the Hessian of  is nonzero, the normals  to  at the points  cover a neighborhood of the vector  in  (the unit sphere in ). Thus, we may choose finitely many points  in a sufficiently small neighborhood of the origin such that

This allows us to write the estimate

Now we restrict our attention to each point  individually. We adjust our coordinates to this point: now  is the origin, we also identify  with . The summand corresponding to  on the right hand side of the previous inequality transforms into

where  is a certain smooth function supported in . By the assumption Equation 75,

where  is a neighborhood of the origin in , and  satisfies Equation 26 (with the constant  instead of  possibly). Take a smooth nonnegative function  that is supported in  and is bounded away from zero on the projection of the support of  to . Then, clearly,

We also note that the norms

are comparable for functions  supported on . Thus, by , we may bound each summand in Equation 76 by

It remains to note that we have a finite number of summands both over  and .

Remark 7.3.

Consider Banach spaces  of functions on  such that multiplication operators

are bounded on  whenever . The inequality

may be reduced to local form

and  satisfies the usual assumptions, with the same argument as in the proof of Lemma 7.2. In particular, the case  allows one to reduce  to  (see Definitions 3.5 and 3.9).

7.2. A version of the Stein–Weiss inequality

7.2.1. Case 

Let  be the weighted Lebesgue space:

Let also  be the operator of convolution with the function . In this section, we work with functions on .

Theorem 7.4.

Let  and let . The operator  maps the space to its dual space  if

(1)

and

and ;

and ;

(2)

and ;

(3)

and

;

and ;

(4)

.

Theorem 7.4 is a variation on the classical Stein–Weiss inequality from Reference 23. In the classical setting, the convolutional kernel and weights are homogeneous.

Remark 7.5.

The conditions listed in Theorem 7.4 are also necessary.

7.2.2. Case 

Theorem 7.6.

Let , and let

(1)

;

(2)

;

(3)

.

Then the operator  maps  to .

Theorems 7.4 and 7.6 are proved by directly examining the boundedness of integral operators with kernel

in the cases , , and , then performing complex interpolation of operators. These are elementary bounds based on when , the Schur test when , and on when .

7.3. Some endpoint estimates

To formulate the endpoint version of inequality Equation 3, we need some Besov spaces (see Reference 3). Given a function , we define the Besov  norm by the formula

where the , , are the Littlewood–Paley projectors on the annuli  and  is the spectral projector on the unit ball  (the symbol  denotes the -dimensional Euclidean ball of radius  centered at ). Using the standard properties of Besov spaces, one may then define Besov spaces on smooth submanifolds of  as well as on their reasonable subdomains.

Proposition 7.7.

The inequality

is true for any  and  satisfying the standard requirements.

The norm in the weighted space on the right hand side is given by the formula

Similarly,  whenever .

Proof.

Since the delta measures are the extremal points of the unit ball in the space of measures, it suffices to prove the proposition for the case where  is a delta measure:

By the Van der Corput lemma (for , this is also the Schrödinger dispersive bound),

Thus, we need to prove the inequality

which is obvious.

Corollary 7.8.

Let  be odd. If we apply Proposition 7.7 with the function , we get the local form of the endpoint case in Equation 3:

Similar to Remark 7.3, we may pass to the global form:

Since  for , we also have

for .

Now we will show how to derive Proposition 1.1 from the case  considered in Reference 6 and Proposition 7.7. We consider the inequality

which, as we have seen, is stronger than Equation 3. Note that in such a formulation,  might be real. We know the inequality holds true in the case  (from Reference 6) and is almost true when , (from Proposition 7.7). We claim that any triple  that satisfies the necessary conditions of Proposition 1.1 might be represented as a convex combination of the said cases:

Solving several elementary equations, we see

We leave to the reader the verification of the conditions

(the easiest way to do this is to sketch the -domain of admissible ) and explain how we interpolate the inequality. First, we note that a linear operator that maps  to  does not depend on the varying parameters, so we may use the classical interpolation theory, specifically, the real interpolation method (see Reference 3). For the image of our operator, we use the formula

see Reference 3. For the domain, we need to show that

In fact,

where the latter space is the space of all functions  such that  (see Reference 10). It is clear that  since .

Acknowledgment

The authors would like to thank Tony Carbery for introducing them to Y. Domar’s papers.

Table of Contents

  1. Abstract
  2. 1. Introduction
    1. 1.1. Overview of the derivative restriction problem
    2. Proposition 1.1 (Corollary of Theorem in 6).
    3. Definition 1.2.
    4. 1.2. Statement of results
    5. Proposition 1.3.
    6. Theorem 1.4.
    7. Theorem 1.6.
    8. Definition 1.9.
    9. Proposition 1.11.
    10. Theorem 1.12.
    11. Corollary 1.13.
    12. Corollary 1.14.
    13. Corollary 1.15.
    14. Theorem 1.16.
  3. 2. Precursors to the current work
    1. Sobolev-type inequalities
    2. Theorem 2.1 (24 and 11).
  4. 3. Study of the spaces
    1. 3.1. Description of the annihilator and Domar’s theory
    2. Lemma 3.1.
    3. Lemma 3.3.
    4. Lemma 3.4.
    5. 3.2. Coincidence of and the spaces defined as kernels of restriction operators
    6. Definition 3.5.
    7. Definition 3.7.
    8. Lemma 3.8.
    9. Definition 3.9.
    10. Lemma 3.10.
    11. 3.3. Proofs of “if” part in Theorems 1.4 and 1.6
    12. Lemma 3.11.
    13. 3.4. Proof of Corollary 1.14
  5. 4. Proof of Theorem 1.16
    1. 4.1. Pointwise estimates of the kernel
    2. Proposition 4.1.
    3. Lemma 4.3.
    4. Lemma 4.4.
    5. 4.2. Interpolation
    6. Lemma 4.5.
    7. 4.3. Strichartz estimates
    8. Theorem 4.7.
  6. 5. Robust estimates
    1. 5.1. Introduction to “numerology”
    2. 5.2. Convexity properties of the function
    3. Definition 5.3.
    4. Lemma 5.5.
    5. Lemma 5.6.
    6. Corollary 5.7.
    7. Lemma 5.8.
    8. Corollary 5.11.
    9. Corollary 5.12.
    10. 5.3. Statement of results by cases
    11. Theorem 5.13.
    12. Theorem 5.14.
    13. Theorem 5.15.
    14. Lemma 5.17.
    15. Lemma 5.19.
    16. Lemma 5.20.
    17. Theorem 5.21.
  7. 6. Sharpness
    1. 6.1. Surface measure conditions
    2. Necessity of  in Theorem 1.4
    3. Necessity of 15
    4. Necessity of 31
    5. Necessity of  in Theorem 4.7
    6. 6.2. Knapp examples
    7. Necessity of 8
    8. Necessity of the condition  in Theorem 1.4
    9. Necessity of 17
    10. Necessity of 32
    11. Necessity of  in Theorem 4.7
    12. 6.3. Pure shifts
    13. Necessity of 6
    14. Necessity of 13
    15. Necessity of 14
    16. Necessity of 30
    17. Necessity of condition  in Theorem 4.7
    18. Necessity of condition  in Theorem 4.7
    19. 6.4. Shifted Knapp example
  8. 7. Additional lemmas and supplementary material
    1. 7.1. Localization argument
    2. Definition 7.1.
    3. Lemma 7.2.
    4. 7.2. A version of the Stein–Weiss inequality
    5. Theorem 7.4.
    6. Theorem 7.6.
    7. 7.3. Some endpoint estimates
    8. Proposition 7.7.
    9. Corollary 7.8.
  9. Acknowledgment

Figures

Figure 1.

Diagram for the case .

Graphic without alt text
Figure 2.

Diagram for the case .

Graphic without alt text
Figure 3.

Diagram for the case  and .

Graphic without alt text
Figure 4.

Diagram for the case  and .

Graphic without alt text
Figure 5.

Diagram for the case  and .

Graphic without alt text
Figure 6.

Diagram for the case .

Graphic without alt text
Figure 7.

What we can reach in the case .

Graphic without alt text
Figure 8.

What we can reach in the case  and .

Graphic without alt text
Figure 9.

What we can reach in the case  and .

Graphic without alt text
Figure 10.

Construction of the points .

Graphic without alt text

Mathematical Fragments

Equation (1)
Equation (2)
Equation (3)
Equations (4), (5)
Proposition 1.1 (Corollary of Theorem in Reference 6).

Let . The inequality Equation 3 is true if and only if

For fixed and with , that means . In the case , the case  is also permitted if .

Definition 1.2.

Let  be a closed smooth embedded -dimensional submanifold of , , and . Define the space  by the formula

Define  to be simply . The first nontrivial space  will often be denoted by .

Equation (9)
Equation (10)
Proposition 1.3.

We have  provided  and . If , this is true provided .

Theorem 1.4.

The inequality Equation 9 is true if and only if , or equivalently .

More generally, inequality Equation 10 is true for and

where the notation indicates the smallest integer greater than or equal to the enclosed value. This covers the entire range . When and the value is also permitted.

Equation (11)
Theorem 1.6.

For any , the spaces  and  coincide with being regarded as a map from to . This occurs when .

For any , the spaces  and  coincide with being regarded as a map from to for the same range of as in Theorem 1.4. This occurs when , or when .

Definition 1.9.

Let  be a natural number, let  and  be nonnegative reals, and let . We say that the higher derivative restriction property holds true if for any smooth compactly supported function  in  variables, the estimate

holds true for any Schwartz function .

Proposition 1.11.

If  holds true and , then

where the numbers  and  are defined by Equation 4 and Equation 5, respectively. In the case , equality in 15 may also occur.

Theorem 1.12.

Let  and . If

then  holds true provided Equation 13Equation 14Equation 15Equation 16, and Equation 17 are satisfied. If 18 does not hold, then  holds true provided Equation 13Equation 17 are satisfied as well as the inequality

Equation (20)
Corollary 1.13.

Suppose for some integer , and let . Then

for any Schwartz function , compact subset , and smooth cutoff that is identically on .

Corollary 1.14.

Suppose holds true and (e.g., if the conditions of Theorems 1.12 or 5.21 are satisfied). Then for each , multi-index with , and smooth compactly supported , there exist and such that

and furthermore

Conversly, if for any compactly supported smooth function , for any , and for any  there exist  and  such that 22 and 23, then holds true (we still assume .

Corollary 1.15.

Let be an integer . Given with Fourier support in the unit ball, there exists a function , also with Fourier support in the unit ball, such that

and

If is even, the result holds for provided the exponent of in 24 is strictly less than .

Equation (26)
Equation (27)
Equation (28)
Equation (29)
Theorem 1.16.

Let satisfy the assumptions above, and let . Inequality Equation 29 is true for the combination of if and only if

and the inequality 31 is strict if .

Equation (33)
Equation (34)
Equation (35)
Lemma 3.1.

Let . The annihilator of  in  can be described as

If , the closure is with respect to the weak-* topology of .

Lemma 3.3.

For any bounded domain , consider the subspace  of vector-valued functions in  supported in . There exists a linear operator , which is inverse to  in the sense

Lemma 3.4.

The set

is dense in  if . In the case , this set is weakly dense.

Equation (37)
Equation (38)
Equation (39)
Definition 3.5.

We say that the statement  holds true if the  admit continuous extensions as  operators for any choice of , and the norms of these extensions are uniform in  (however, we do not require any uniformity with respect to ).

We say that  is true if  is true and for any choice of  the operators  extend continuously from the domain

to a family of mappings  whose norms are bounded uniformly by .

Definition 3.7.

For a fixed , define the set  by the formula

Lemma 3.8.

Suppose that  has nonvanishing curvature, , and holds. Then, .

Definition 3.9.

We say that the statement  holds true if the mapping

extends to a bounded linear operator between the spaces  and  for any compact set .

Lemma 3.10.

For any function  such that , where , there exists a sequence  of Schwartz functions such that

Equation (43)
Lemma 3.11.

The statement  holds true if . For every there exists such that is true. When and , the value suffices. Finally, in odd dimensions holds for .

Equation (44)
Equation (45)
Equation (46)
Equation (47)
Equation (49)
Equation (50)
Proposition 4.1.

The kernel  defined in Equation 50 satisfies the bound

Equation (52)
Equation (53)
Lemma 4.3.

For any , and any ,

Equation (54)
Equation (55)
Lemma 4.4.

For any , and any ,

Equation (56)
Equation (57)
Lemma 4.5.

For any ,

Equation (58)
Equation (59)
Equation (60)
Equation (61)
Theorem 4.7.

The inequality

holds true if

(1)

and

;

;

with equality permitted if ;

with equality also permitted if or ;

(2)

and

;

;

;

.

Remark 5.1.

We are mostly interested in the case  in Equation 12. We claim that in the subcritical case , the second term on the right hand side of this inequality is unnecessary. If  and  is true, then a simpler inequality

also holds true. Indeed, if  is true, then Equation 15 and Equation 17 are valid. However, in this case, these conditions are also sufficient for 63 to be true (see Theorem 1.16 and Figure 1).

Remark 5.2.

In the supercritical case , the condition Equation 15 follows from Equation 17 since  in this case (see Figure 1 as well). Note also that Equation 16 is equivalent to

This inequality, in its turn, leads to Equation 13 provided  (which is true by Equation 17). Thus, in the case , the conditions in Proposition 1.11 are reduced to Equation 14, 64, and Equation 17.

Equation (65)
Lemma 5.5.

For any  and , there exists a constant  such that the inequality

is true provided  lies in the friendly region and  for .

Lemma 5.6.

For any  such that , there exist coefficients  such that

for any function  and any .

Corollary 5.7.

For any  and , there exists a constant  such that the inequality

is true provided  lies in the friendly region and  for .

Lemma 5.8.

Let  be a finite sequence and let . Assume that

, and , . Then, .

Remark 5.10.

Using the homogeneity, one can replace the assumptions of Lemma 5.8 by

, and , for some positive constant . Then, .

Corollary 5.11.

The -domain is convex in the sense that if  is a convex combination of  and  (we assume ), and the latter two points belong to the -domain, then the former point lies in it as well.

Equation (67)
Corollary 5.12.

Let  be a point with natural -coordinate lying in the intersection of friendly and subcritical domains. Suppose that the point  lies on the segment , has natural first coordinate , and lies in the friendly domain. If , then  lies belongs to the -domain.

Case ( and ).

Our reasonings are illustrated by Figure 9. We introduce two auxiliary points  and :

We have used two types of the notion “integer part of a number”; see formula Equation 20.

We connect the point  to  and . Since the point  lies in the intersection of friendly and subcritical regions, Corollary 5.12 applied to  in the role of  says that  is true for all pairs  such that ; in other words

Clearly, the same assertion is true for larger  when  is fixed. The situation with the point  is slightly more complicated: it may lie outside the friendly region if its -coordinate is too large. If it is not so (i.e., ), then we may apply Corollary 5.12 to the point  in the role of  and achieve  is true for all pairs  such that ; in other words

We summarize our results.

Theorem 5.15.

Let  and let . If , then  holds true if

If , then  holds true if

Remark 5.16.

If  and  holds true, then .

Lemma 5.17.

For any , we have . For , all points  lie on the line .

Lemma 5.20.

If  is a number between  and  and , then  belongs to the -domain. If  and  belongs to the friendly region, then  belongs to the -domain.

Theorem 5.21.

Assume , . The -domain contains the convex hull of points specified below. We always include the points , , and  in our list. The other points are specified in the following table:

Equation (69)
Equation (70)
Equation (71)
Equation (72)
Equation (73)
Definition 7.1.

Let the numbers  be of the same nature as in Definition 1.9. Let  be a neighborhood of the origin in , let  be a smooth function such that , , and the determinant of the Hessian of  at the origin does not vanish. Further, we assume Equation 26. We say that the statement  holds true if the inequality

holds true for any smooth function  supported in .

Lemma 7.2.

The statement  is true provided the statement is true for any  satisfying the conditions of Definition 7.1.

Equations (74), (75)
Equation (76)
Remark 7.3.

Consider Banach spaces  of functions on  such that multiplication operators

are bounded on  whenever . The inequality

may be reduced to local form

and  satisfies the usual assumptions, with the same argument as in the proof of Lemma 7.2. In particular, the case  allows one to reduce  to  (see Definitions 3.5 and 3.9).

Theorem 7.4.

Let  and let . The operator  maps the space to its dual space  if

(1)

and

and ;

and ;

(2)

and ;

(3)

and

;

and ;

(4)

.

Theorem 7.6.

Let , and let

(1)

;

(2)

;

(3)

.

Then the operator  maps  to .

Proposition 7.7.

The inequality

is true for any  and  satisfying the standard requirements.

Corollary 7.8.

Let  be odd. If we apply Proposition 7.7 with the function , we get the local form of the endpoint case in Equation 3:

Similar to Remark 7.3, we may pass to the global form:

Since  for , we also have

for .

References

Reference [1]
Shmuel Agmon, Spectral properties of Schrödinger operators and scattering theory, Ann. Scuola Norm. Sup. Pisa Cl. Sci. (4) 2 (1975), no. 2, 151–218. MR397194,
Show rawAMSref \bib{Agmon}{article}{ author={Agmon, Shmuel}, title={Spectral properties of Schr\"{o}dinger operators and scattering theory}, journal={Ann. Scuola Norm. Sup. Pisa Cl. Sci. (4)}, volume={2}, date={1975}, number={2}, pages={151--218}, issn={0391-173X}, review={\MR {397194}}, }
Reference [2]
Jong-Guk Bak, Sharp estimates for the Bochner-Riesz operator of negative order in , Proc. Amer. Math. Soc. 125 (1997), no. 7, 1977–1986, DOI 10.1090/S0002-9939-97-03723-4. MR1371114,
Show rawAMSref \bib{Bak}{article}{ author={Bak, Jong-Guk}, title={Sharp estimates for the Bochner-Riesz operator of negative order in $\mathbf {R}^2$}, journal={Proc. Amer. Math. Soc.}, volume={125}, date={1997}, number={7}, pages={1977--1986}, issn={0002-9939}, review={\MR {1371114}}, doi={10.1090/S0002-9939-97-03723-4}, }
Reference [3]
Jöran Bergh and Jörgen Löfström, Interpolation spaces. An introduction, Springer-Verlag, Berlin-New York, 1976. Grundlehren der Mathematischen Wissenschaften, No. 223. MR0482275,
Show rawAMSref \bib{BerghLofstrom}{book}{ author={Bergh, J\"{o}ran}, author={L\"{o}fstr\"{o}m, J\"{o}rgen}, title={Interpolation spaces. An introduction}, note={Grundlehren der Mathematischen Wissenschaften, No. 223}, publisher={Springer-Verlag, Berlin-New York}, date={1976}, pages={x+207}, review={\MR {0482275}}, }
Reference [4]
Jean Bourgain and Haïm Brezis, On the equation and application to control of phases, J. Amer. Math. Soc. 16 (2003), no. 2, 393–426, DOI 10.1090/S0894-0347-02-00411-3. MR1949165,
Show rawAMSref \bib{BourgainBrezis}{article}{ author={Bourgain, Jean}, author={Brezis, Ha\"{\i }m}, title={On the equation $\mathrm {div}\, Y=f$ and application to control of phases}, journal={J. Amer. Math. Soc.}, volume={16}, date={2003}, number={2}, pages={393--426}, issn={0894-0347}, review={\MR {1949165}}, doi={10.1090/S0894-0347-02-00411-3}, }
Reference [5]
Jean Bourgain and Larry Guth, Bounds on oscillatory integral operators based on multilinear estimates, Geom. Funct. Anal. 21 (2011), no. 6, 1239–1295, DOI 10.1007/s00039-011-0140-9. MR2860188,
Show rawAMSref \bib{BourgainGuth}{article}{ author={Bourgain, Jean}, author={Guth, Larry}, title={Bounds on oscillatory integral operators based on multilinear estimates}, journal={Geom. Funct. Anal.}, volume={21}, date={2011}, number={6}, pages={1239--1295}, issn={1016-443X}, review={\MR {2860188}}, doi={10.1007/s00039-011-0140-9}, }
Reference [6]
Yonggeun Cho, Zihua Guo, and Sanghyuk Lee, A Sobolev estimate for the adjoint restriction operator, Math. Ann. 362 (2015), no. 3-4, 799–815, DOI 10.1007/s00208-014-1130-7. MR3368083,
Show rawAMSref \bib{ChoGuoLee}{article}{ author={Cho, Yonggeun}, author={Guo, Zihua}, author={Lee, Sanghyuk}, title={A Sobolev estimate for the adjoint restriction operator}, journal={Math. Ann.}, volume={362}, date={2015}, number={3-4}, pages={799--815}, issn={0025-5831}, review={\MR {3368083}}, doi={10.1007/s00208-014-1130-7}, }
Reference [7]
Yngve Domar, On the spectral synthesis problem for -dimensional subsets of , Ark. Mat. 9 (1971), 23–37, DOI 10.1007/BF02383635. MR324319,
Show rawAMSref \bib{Domar}{article}{ author={Domar, Yngve}, title={On the spectral synthesis problem for $(n-1)$-dimensional subsets of $\mathbf {R}^{n},\,n\geq 2$}, journal={Ark. Mat.}, volume={9}, date={1971}, pages={23--37}, issn={0004-2080}, review={\MR {324319}}, doi={10.1007/BF02383635}, }
Reference [8]
Charles Fefferman, Inequalities for strongly singular convolution operators, Acta Math. 124 (1970), 9–36, DOI 10.1007/BF02394567. MR257819,
Show rawAMSref \bib{Fefferman}{article}{ author={Fefferman, Charles}, title={Inequalities for strongly singular convolution operators}, journal={Acta Math.}, volume={124}, date={1970}, pages={9--36}, issn={0001-5962}, review={\MR {257819}}, doi={10.1007/BF02394567}, }
[9]
Elida V. Ferreyra, On a negative result concerning interpolation with change of measures for Lorentz spaces, Proc. Amer. Math. Soc. 125 (1997), no. 5, 1413–1417, DOI 10.1090/S0002-9939-97-03699-X. MR1363418,
Show rawAMSref \bib{Ferreyra}{article}{ author={Ferreyra, Elida V.}, title={On a negative result concerning interpolation with change of measures for Lorentz spaces}, journal={Proc. Amer. Math. Soc.}, volume={125}, date={1997}, number={5}, pages={1413--1417}, issn={0002-9939}, review={\MR {1363418}}, doi={10.1090/S0002-9939-97-03699-X}, }
Reference [10]
Dietrich Freitag, Real interpolation of weighted -spaces, Math. Nachr. 86 (1978), 15–18, DOI 10.1002/mana.19780860103. MR532769,
Show rawAMSref \bib{Freitag}{article}{ author={Freitag, Dietrich}, title={Real interpolation of weighted $L_{p}$-spaces}, journal={Math. Nachr.}, volume={86}, date={1978}, pages={15--18}, issn={0025-584X}, review={\MR {532769}}, doi={10.1002/mana.19780860103}, }
Reference [11]
Michael Goldberg, The Helmholtz equation with data and Bochner-Riesz multipliers, Math. Res. Lett. 23 (2016), no. 6, 1665–1679, DOI 10.4310/MRL.2016.v23.n6.a5. MR3621102,
Show rawAMSref \bib{Goldberg}{article}{ author={Goldberg, Michael}, title={The Helmholtz equation with $L^p$ data and Bochner-Riesz multipliers}, journal={Math. Res. Lett.}, volume={23}, date={2016}, number={6}, pages={1665--1679}, issn={1073-2780}, review={\MR {3621102}}, doi={10.4310/MRL.2016.v23.n6.a5}, }
Reference [12]
M. Goldberg and W. Schlag, A limiting absorption principle for the three-dimensional Schrödinger equation with potentials, Int. Math. Res. Not. 75 (2004), 4049–4071, DOI 10.1155/S1073792804140324. MR2112327,
Show rawAMSref \bib{Goldberg_Schlag}{article}{ author={Goldberg, M.}, author={Schlag, W.}, title={A limiting absorption principle for the three-dimensional Schr\"{o}dinger equation with $L^p$ potentials}, journal={Int. Math. Res. Not.}, date={2004}, number={75}, pages={4049--4071}, issn={1073-7928}, review={\MR {2112327}}, doi={10.1155/S1073792804140324}, }
Reference [13]
Larry Guth, A restriction estimate using polynomial partitioning, J. Amer. Math. Soc. 29 (2016), no. 2, 371–413, DOI 10.1090/jams827. MR3454378,
Show rawAMSref \bib{Guth1}{article}{ author={Guth, Larry}, title={A restriction estimate using polynomial partitioning}, journal={J. Amer. Math. Soc.}, volume={29}, date={2016}, number={2}, pages={371--413}, issn={0894-0347}, review={\MR {3454378}}, doi={10.1090/jams827}, }
Reference [14]
Larry Guth, Restriction estimates using polynomial partitioning II, Acta Math. 221 (2018), no. 1, 81–142, DOI 10.4310/ACTA.2018.v221.n1.a3. MR3877019,
Show rawAMSref \bib{Guth2}{article}{ author={Guth, Larry}, title={Restriction estimates using polynomial partitioning II}, journal={Acta Math.}, volume={221}, date={2018}, number={1}, pages={81--142}, issn={0001-5962}, review={\MR {3877019}}, doi={10.4310/ACTA.2018.v221.n1.a3}, }
Reference [15]
Alexandru D. Ionescu and Wilhelm Schlag, Agmon-Kato-Kuroda theorems for a large class of perturbations, Duke Math. J. 131 (2006), no. 3, 397–440, DOI 10.1215/S0012-7094-06-13131-9. MR2219246,
Show rawAMSref \bib{Ionescu_Schlag}{article}{ author={Ionescu, Alexandru D.}, author={Schlag, Wilhelm}, title={Agmon-Kato-Kuroda theorems for a large class of perturbations}, journal={Duke Math. J.}, volume={131}, date={2006}, number={3}, pages={397--440}, issn={0012-7094}, review={\MR {2219246}}, doi={10.1215/S0012-7094-06-13131-9}, }
[16]
R. A. Kerman, Convolution theorems with weights, Trans. Amer. Math. Soc. 280 (1983), no. 1, 207–219, DOI 10.2307/1999609. MR712256,
Show rawAMSref \bib{Kerman}{article}{ author={Kerman, R. A.}, title={Convolution theorems with weights}, journal={Trans. Amer. Math. Soc.}, volume={280}, date={1983}, number={1}, pages={207--219}, issn={0002-9947}, review={\MR {712256}}, doi={10.2307/1999609}, }
Reference [17]
S. V. Kislyakov, D. V. Maksimov, and D. M. Stolyarov, Differential expressions with mixed homogeneity and spaces of smooth functions they generate in arbitrary dimension, J. Funct. Anal. 269 (2015), no. 10, 3220–3263, DOI 10.1016/j.jfa.2015.09.001. MR3401616,
Show rawAMSref \bib{KislyakovMaximovStolyarov}{article}{ author={Kislyakov, S. V.}, author={Maksimov, D. V.}, author={Stolyarov, D. M.}, title={Differential expressions with mixed homogeneity and spaces of smooth functions they generate in arbitrary dimension}, journal={J. Funct. Anal.}, volume={269}, date={2015}, number={10}, pages={3220--3263}, issn={0022-1236}, review={\MR {3401616}}, doi={10.1016/j.jfa.2015.09.001}, }
Reference [18]
Camil Muscalu and Wilhelm Schlag, Classical and multilinear harmonic analysis. Vol. I, Cambridge Studies in Advanced Mathematics, vol. 137, Cambridge University Press, Cambridge, 2013. MR3052498,
Show rawAMSref \bib{MuscaluSchlag}{book}{ author={Muscalu, Camil}, author={Schlag, Wilhelm}, title={Classical and multilinear harmonic analysis. Vol. I}, series={Cambridge Studies in Advanced Mathematics}, volume={137}, publisher={Cambridge University Press, Cambridge}, date={2013}, pages={xviii+370}, isbn={978-0-521-88245-3}, review={\MR {3052498}}, }
Reference [19]
Jean Van Schaftingen, Limiting Sobolev inequalities for vector fields and canceling linear differential operators, J. Eur. Math. Soc. (JEMS) 15 (2013), no. 3, 877–921, DOI 10.4171/JEMS/380. MR3085095,
Show rawAMSref \bib{vanSchaftingen}{article}{ author={Van Schaftingen, Jean}, title={Limiting Sobolev inequalities for vector fields and canceling linear differential operators}, journal={J. Eur. Math. Soc. (JEMS)}, volume={15}, date={2013}, number={3}, pages={877--921}, issn={1435-9855}, review={\MR {3085095}}, doi={10.4171/JEMS/380}, }
Reference [20]
Jean Van Schaftingen, Limiting Bourgain-Brezis estimates for systems of linear differential equations: theme and variations, J. Fixed Point Theory Appl. 15 (2014), no. 2, 273–297, DOI 10.1007/s11784-014-0177-0. MR3298002,
Show rawAMSref \bib{vanSchaftingen2}{article}{ author={Van Schaftingen, Jean}, title={Limiting Bourgain-Brezis estimates for systems of linear differential equations: theme and variations}, journal={J. Fixed Point Theory Appl.}, volume={15}, date={2014}, number={2}, pages={273--297}, issn={1661-7738}, review={\MR {3298002}}, doi={10.1007/s11784-014-0177-0}, }
Reference [21]
E. M. Stein, Oscillatory integrals in Fourier analysis, Beijing lectures in harmonic analysis (Beijing, 1984), Ann. of Math. Stud., vol. 112, Princeton Univ. Press, Princeton, NJ, 1986, pp. 307–355. MR864375,
Show rawAMSref \bib{Stein}{article}{ author={Stein, E. M.}, title={Oscillatory integrals in Fourier analysis}, conference={ title={Beijing lectures in harmonic analysis}, address={Beijing}, date={1984}, }, book={ series={Ann. of Math. Stud.}, volume={112}, publisher={Princeton Univ. Press, Princeton, NJ}, }, date={1986}, pages={307--355}, review={\MR {864375}}, }
Reference [22]
Elias M. Stein, Harmonic analysis: real-variable methods, orthogonality, and oscillatory integrals, Princeton Mathematical Series, vol. 43, Princeton University Press, Princeton, NJ, 1993. With the assistance of Timothy S. Murphy; Monographs in Harmonic Analysis, III. MR1232192,
Show rawAMSref \bib{SteinBook}{book}{ author={Stein, Elias M.}, title={Harmonic analysis: real-variable methods, orthogonality, and oscillatory integrals}, series={Princeton Mathematical Series}, volume={43}, note={With the assistance of Timothy S. Murphy; Monographs in Harmonic Analysis, III}, publisher={Princeton University Press, Princeton, NJ}, date={1993}, pages={xiv+695}, isbn={0-691-03216-5}, review={\MR {1232192}}, }
Reference [23]
E. M. Stein and Guido Weiss, Fractional integrals on -dimensional Euclidean space, J. Math. Mech. 7 (1958), 503–514, DOI 10.1512/iumj.1958.7.57030. MR0098285,
Show rawAMSref \bib{SteinWeiss}{article}{ author={Stein, E. M.}, author={Weiss, Guido}, title={Fractional integrals on $n$-dimensional Euclidean space}, journal={J. Math. Mech.}, volume={7}, date={1958}, pages={503--514}, review={\MR {0098285}}, doi={10.1512/iumj.1958.7.57030}, }
Reference [24]
D. M. Stolyarov, Bilinear embedding in theorems for differential operators in (Russian, with English summary), Zap. Nauchn. Sem. S.-Peterburg. Otdel. Mat. Inst. Steklov. (POMI) 424 (2014), no. Issledovaniya po Lineĭnym Operatorami Teorii Funktsiĭ. 42, 210–234, DOI 10.1007/s10958-015-2527-x; English transl., J. Math. Sci. (N.Y.) 209 (2015), no. 5, 792–807. MR3481451,
Show rawAMSref \bib{Stolyarov}{article}{ author={Stolyarov, D. M.}, title={Bilinear embedding in theorems for differential operators in $\mathbb {R}^2$}, language={Russian, with English summary}, journal={Zap. Nauchn. Sem. S.-Peterburg. Otdel. Mat. Inst. Steklov. (POMI)}, volume={424}, date={2014}, number={Issledovaniya po Line\u {\i }nym Operatorami Teorii Funktsi\u {\i }. 42}, pages={210--234}, issn={0373-2703}, translation={ journal={J. Math. Sci. (N.Y.)}, volume={209}, date={2015}, number={5}, pages={792--807}, issn={1072-3374}, }, review={\MR {3481451}}, doi={10.1007/s10958-015-2527-x}, }
Reference [25]
Peter A. Tomas, A restriction theorem for the Fourier transform, Bull. Amer. Math. Soc. 81 (1975), 477–478, DOI 10.1090/S0002-9904-1975-13790-6. MR358216,
Show rawAMSref \bib{Tomas}{article}{ author={Tomas, Peter A.}, title={A restriction theorem for the Fourier transform}, journal={Bull. Amer. Math. Soc.}, volume={81}, date={1975}, pages={477--478}, issn={0002-9904}, review={\MR {358216}}, doi={10.1090/S0002-9904-1975-13790-6}, }

Article Information

MSC 2010
Primary: 42B10 (Fourier and Fourier-Stieltjes transforms and other transforms of Fourier type)
Secondary: 42B20 (Singular and oscillatory integrals)
Author Information
Michael Goldberg
Department of Mathematical Sciences, University of Cincinnati, Cincinnati, Ohio 45221-0025
goldbeml@ucmail.uc.edu
ORCID
MathSciNet
Dmitriy Stolyarov
Chebyshev Lab, St. Petersburg State Univeristy, 14th line 29b, Vasilyevsky Island, St. Petersburg 199178, Russia; and St. Petersburg Department of Steklov Mathematical Institute, Fontanka 27, St. Petersburg 191023, Russia
d.m.stolyarov@spbu.ru
MathSciNet
Additional Notes

The first author received support from Simons Foundation grant #281057.

The second author received support from Russian Foundation for Basic Research grant #17-01-00607.

Journal Information
Transactions of the American Mathematical Society, Series B, Volume 7, Issue 3, ISSN 2330-0000, published by the American Mathematical Society, Providence, Rhode Island.
Publication History
This article was received on and published on .
Copyright Information
Copyright 2020 by the authors under Creative Commons Attribution-Noncommercial 3.0 License (CC BY NC 3.0)
Article References
  • Permalink
  • Permalink (PDF)
  • DOI 10.1090/btran/45
  • MathSciNet Review: 4147581
  • Show rawAMSref \bib{4147581}{article}{ author={Goldberg, Michael}, author={Stolyarov, Dmitriy}, title={Restrictions of higher derivatives of the Fourier transform}, journal={Trans. Amer. Math. Soc. Ser. B}, volume={7}, number={3}, date={2020}, pages={46-96}, issn={2330-0000}, review={4147581}, doi={10.1090/btran/45}, }

Settings

Change font size
Resize article panel
Enable equation enrichment

Note. To explore an equation, focus it (e.g., by clicking on it) and use the arrow keys to navigate its structure. Screenreader users should be advised that enabling speech synthesis will lead to duplicate aural rendering.

For more information please visit the AMS MathViewer documentation.