Home On certain trilinear oscillatory integral inequalities
Article Open Access

On certain trilinear oscillatory integral inequalities

  • Michael Christ EMAIL logo
Published/Copyright: May 29, 2025

Abstract

Inequalities are established for certain trilinear scalar-valued functionals. These functionals act on measurable functions of one real variable, are defined by integration over two- or three-dimensional spaces, and are controlled in terms of Lebesgue space norms of the functions, and in terms of negative powers of large parameters describing a degree of oscillation. Related sublevel set inequalities are a central element of the analysis.

2010 Mathematics Subject Classification: 42B20; 26D15

1 Introduction

We investigate upper bounds for scalar-valued multilinear oscillatory integral forms

(1.1) T λ ϕ ( f ) = B e i λ ϕ ( x ) j J f j ( x j ) d x ,

along with related topics. Here B is a ball, or cube in ( R d ) J , J is a finite index set of cardinality |J| ≥ 2, f = (f j : jJ) is a tuple of rather arbitrary functions f j : R d C , x = ( x j : j J ) ( R d ) J , ϕ : ( R d ) J R is a C function, and λ R is a large parameter. More generally, one can form

(1.2) S λ ( f ) = B R D e i λ ϕ ( x ) j J ( f j φ j ) ( x ) d x ,

with B R D a ball, φ j : B R d smooth submersions, and with the cardinality of J finite but |J| ⋅ d possibly large relative to D. We seek upper bounds, for | T λ ϕ ( f ) | and for |S λ (f)|, that are small when |λ| is large, require no smoothness hypothesis on f, are uniform over a large class of f, reflect cancellation due to oscillation of eiλϕ when |λ| is large, and also reflect the influence of geometric and algebraic effects implicit in (ϕ, (φ j : jJ)). In this paper we establish such bounds, and deduce an application to related multilinear forms in which no oscillatory factors are overtly present.

1.1 Some background

Inequalities of the form

(1.3) | T λ ϕ ( f ) | C | λ | γ j J f j L p j ( R d ) ,

with γ > 0 and C < ∞ dependent on ϕ and on η, have been analyzed in various works. Hörmander [1] established the fundamental upper bound O | λ | d / 2 f 1 2 f 2 2 for the bilinear case |J| = 2, with the mixed Hessian 2 ϕ ( x , y ) x y everywhere nonsingular. The bilinear case, with x, y in spaces of unequal dimensions, has been intensively studied in connection with Fourier restriction inequalities. Likewise, in connection with Fourier restriction, multilinear forms have been investigated, in which each function f j is individually acted upon by a linear oscillatory integral operator, and the product of the resulting functions is integrated. Such multilinear forms are not studied here.

Forms | T λ ϕ ( f ) | and |S λ (f)|, with |J| ≥ 3, have been studied by Phong-Stein-Sturm [2], Gilula-Gressman-Xiao [3], and others. An introductory treatment can be found in Ref. [4]. The works [3], [2] deal with general phase functions ϕ, and seek optimal relationships between ϕ, decay exponents γ, and Lebesgue exponents p j [2]. Also emphasizes the issue of stability – whether the optimal exponent γ is a lower semicontinuous function of ϕ.

The regime |J| > D/d is singular, in the sense that the integral extends only over a positive codimension subvariety of the Cartesian product of the domains of the functions f j . Variants of the form (1.2), in the singular regime and with all mappings φ j linear, were investigated by Li, Tao, Thiele and the present author [5]. They established conditions under which there exists an exponent γ > 0 for which a corresponding inequality holds. One of the results of the present paper relaxes the assumption of linearity. Another treats certain cases with φ j linear that were not treated in Ref. [5], and provides an alternative proof of one of the main results of that work.

Results of this type under lower bounds for certain partial derivatives of the phase function, but with no upper bounds at all, have been investigated by Carbery and Wright [6] for |J| ≥ 3, building on earlier work [7] for the bilinear case |J| = 2. This thread is not developed further in the present paper, in which upper bounds are implicit through smoothness hypotheses on phase functions.

For certain ranges of exponent tuples p = (p j : jJ), the works [3], [2] establish upper bounds for (1.1) in terms of products of L p j norms of the factors f j , with optimal exponents γ as |λ| → ∞, up to powers of log(1 + |λ|), where “optimal” means largest possible under indicated hypotheses on ϕ for given p. We will not review the hypotheses of those works precisely, but their general form is significant for our discussion. For each α (with α j ≠ 0 for at least two distinct indices j) there are certain parameters p, γ for which nonvanishing of ∂ α ϕ/∂x α at x 0 implies validity of (1.3), with B = B(x 0, r) for sufficiently small r > 0. This conclusion is independent of other coefficients in the Taylor expansion of ϕ about x 0. Thus any other nonvanishing coefficients imply corresponding inequalities, and interpolation of the resulting inequalities yields further inequalities. All bounds obtained in these cited works are consequences of bounds obtained in this way, together with inclusions among L p spaces resulting from Hölder’s inequality.

1.2 Four questions

Question 1.1.

Let ϕ be a real analytic, real-valued phase function. Let x 0 ( R d ) J . For which γ > 0 do there exist C < ∞ and a neighborhood B of x 0 such that

(1.4) | T λ ϕ ( f ) | C | λ | γ j J f j L  for all functions  f j L ,

for every λ R ?

Thus we ask for the very best exponent γ, rather than the best for a bound in terms of j f j L p j .

For multilinear forms of the more general type (1.2), a less precise question is at present appropriate.

Question 1.2.

Let ϕ be a real analytic, real-valued phase function. Let φ j : R D R d be real analytic submersions. Let x 0 R D . Under what conditions on ϕ and on (φ j : jJ) does there exist γ > 0 such that

(1.5) | S λ ( f ) | C | λ | γ j J f j L  for all functions  f j L ,

for every λ R ?

Roughly speaking, the difficulty in establishing inequalities (1.5) increases as the ratio d/D increases.

In the formulation (1.4), the main structural hypothesis is that the nonoscillatory part of the integrand is a product of factors f j (x j ). No smoothness is required of these factors, and the strongest possible size restriction, L , is imposed. As a refinement, one could ask to what degree the L norms could be replaced by weaker L p j norms without reducing γ. One focus of the present paper is on obtaining a comparatively large exponent γ, rather than on weakening the hypotheses under which it is obtained. Although we are not able to determine optimal exponents γ in Question 1.1, we improve on the largest exponent previously known for generic real analytic phases in the trilinear case with d = 1. We find that interactions between monomial terms can give rise to upper bounds not obtainable from monomial-based inequalities.

A second focus, for related functionals, is on obtaining some decay inequality of power law type, for situations in which no decay bound was previously known, without attention to the value of the exponent γ. Theorem 4.3 is one result in that direction.

Oscillation can arise implicitly through the presence of high frequency Fourier components in the factors f j , instead of explicitly through the presence of overt oscillatory factors eiλϕ . This suggests a third question.

Question 1.3.

Let η be smooth and compactly supported. Let J be a finite index set. Let φ j : R D R d be real analytic submersions. Under what circumstances can the quantity ∫∏ jJ (f j φ j )(x) η(x) dx be majorized by a product of strictly negative Sobolev norms of the factors f j ?

In analyzing these three questions concerning oscillatory integral forms, we are led to questions concerning sublevel sets. Let φ j : [ 0,1 ] 2 R 1 and a j : [ 0,1 ] 2 R 1 be smooth. To an ordered triple f of Lebesgue measurable f j : R R , and to ɛ > 0, associate the sublevel set

(1.6) S ( f , ε ) = x [ 0,1 ] 2 : j = 1 3 a j ( x ) ( f j φ j ) ( x ) < ε .

Denote by |S(f, ɛ)| the Lebesgue measure of this set.

Question 1.4.

Under what hypotheses on (φ j , a j : j ∈ {1, 2, 3}) and what conditions on f do there exist γ > 0 and C < ∞ such that for every small ɛ > 0,

(1.7) | S ( f , ε ) | C ε γ ?

For such an inequality to hold, some condition on f is needed to exclude trivial solutions with f ≡ 0 or with every |f j | small. Likewise, the existence of solutions f of ∑ j a j ⋅ (f j φ j ) ≡ 0 must be excluded.

Situations in which there is a small family of such exact solutions, e.g. constant f or affine f, are also of interest. In such a situation, one asks instead whether the inequality (1.7) can fail to hold only for those f that are closely approximable by elements of the family of exact solutions.

One theme of this paper is the tapestry of interconnections between these questions.

1.3 Content of paper

We begin with remarks and examples placing Question 1.1 better in context. We then focus on the trilinear case, with d = 1. In all previous results for this trilinear one-dimensional case known to this author, the exponent γ in the upper bound (1.4) has been less than or equal to 1 2 . We introduce a property of the phase function ϕ, called rank one nondegeneracy. Its negation, rank one degeneracy, implies that (1.4) cannot hold for any γ > 1 2 . Our first main result is that under a mild auxiliary hypothesis, the inequality (1.4) does hold for some γ > 1 2 if ϕ is rank one nondegenerate. Building on this result, Zirui Zhou has relaxed, though not eliminated, the auxiliary hypothesis in her 2024 dissertation [8].

In a sense clarified in Section 15, generic C ω phase functions are rank one nondegenerate. We also explore the connection between multilinear oscillatory forms (1.2) and the multilinear oscillatory forms studied in Ref. [5].

Our main results concerning (1.4) and (1.5), respectively, are Theorems 4.1 and 4.2. Their proofs are based on decomposition in phase space, a dichotomy between structure and pseudo-randomness, a two scale analysis, and a connection with sublevel sets.

In Section 16 we use the same method to give an alternative proof of a theorem of Li, Tao, Thiele, and the author [5], and to establish an extension.

We utilize two term sublevel set inequalities (1.7) – meaning that only two general functions f j appear in their formulation – for both scalar- and vector-valued functions as a tool in proving trilinear oscillatory inequalities. In Section 17 we reverse this flow of ideas to study special types of three term sublevel set inequalities. We consider sublevel sets of the type (1.6), with constant coefficients a j . We use oscillatory inequalities established earlier in the paper to prove upper bounds for their Lebesgue measures under natural hypotheses on (φ j : j ∈ {1, 2, 3}) and appropriate nonconstancy hypotheses on f. In Section 18 we develop a rather different method to study sublevel set inequalities for nonconstant coefficients a j , in the special case in which the mappings φ j are all linear. In Section 20 we construct an example, based on multiprogressions of rank greater than 1, concerning one of the simplest possible vector-valued sublevel set inequalities.

In most of the paper, we assume phase functions and mappings φ j to be real analytic, rather than merely infinitely differentiable. This is done primarily because hypotheses can be formulated more simply in the C ω case, with its dichotomy between functions that vanish identically, and those that vanish to finite order.

There are connections between the results and methods in this paper, a much earlier work of Bourgain [9], and recent works of Peluse and Prendiville [10], [11], [12] and of Krause, Mirek, Peluse, and Wright [13] involving quantitative nonlinear analogues of Roth’s theorem, cut norms, and degree reduction. See also [14] for an exposition of some of these ideas.

The principal results of this paper were first announced in Ref. [15]. The techniques employed here have been further developed and applied in several sequels, jointly with Durcik and Roos [16] in work on multilinear singular integral operators, with Durcik, Kovač, and Roos [17] in work on averages associated to R -actions, and with Zhou [18] in work on certain bilinear maximal operators. An extension to data of limited regularity has been developed in Ref. [19]. More singular cases, such as |J| ≥ 4 in Theorem 4.2, will be treated in future work via a further extension of these techniques.

The author is indebted to Zirui Zhou for corrections and useful comments on the exposition, to Philip Gressman for a useful conversation, and to Terence Tao for pointing out the connection with the works of Bourgain, Peluse, and Prendiville. He thanks Craig Evans for serendipitously acquainting him with related work of Joly, Métivier, and Rauch [20], and for stimulating discussion.

1.4 Dedication

This paper is dedicated to Robert Fefferman. In the Fall of 1977, the author, as a beginning PhD student, attended Professor Fefferman’s course on harmonic analysis and singular integral operators at the University of Chicago. The brilliance of the lectures, the depth of the lecturer’s insight, his passion for ideas, and his kind encouragement inspired this pupil, laying a firm foundation within a few months for a career that has extended over nearly half a century, and continues. I was, am, and will remain deeply grateful.

2 Examples

In all of the examples of this section, and most of the main results of this paper, d = 1. B is often replaced by [0, 1] J , so T λ ϕ ( f ) = [ 0,1 ] J e i λ ϕ ( x ) j J f j ( x j ) d x . In these examples, excepting Example 2.3, T λ ϕ is trilinear.

Example 2.1

[21]. For |J| = 3 and ϕ(x 1, x 2, x 3) = x 2(x 1 + x 3), the inequality (1.4) holds with γ = 1. To justify this, write

T λ ϕ ( f ) = | λ | 1 / 2 [ 0,1 ] 2 f 1 ( x ) F ( y ) e i λ x y d x d y

where F ( y ) = c f 2 ( y ) | λ | 1 / 2 f 3 ̂ ( λ y ) for a certain harmless constant c ≠ 0. This function satisfies F 2 f 2 f 3 2 f 2 f 3 . One factor of |λ|−1/2 has already been gained. The remaining integral is O | λ | 1 / 2 f 1 2 F 2 by Plancherel’s theorem and a change of variables.

This gives | T λ ϕ ( f ) | C | λ | 1 f 1 2 f 2 f 3 2 . For p ∈ [2, ∞), the optimal bound in terms of f 1 f 2 f 3 p is O(|λ|−1|λ|1/p ). This can be seen by considering f j ( x j ) = e i λ x j 1 [ 0,1 ] ( x j ) for j = 1, 3. Integrating with respect to x 1, x 3 then leaves a function of x 2 whose real part is bounded below on [ 0 , π 4 | λ | 1 ] by a positive constant independent of λ. Choose f 2 to be the indicator function of [ 0 , π 4 | λ | 1 ] .

However, for ϕ = x 1 x 2 + x 2 x 3, the situation changes if f 2 is replaced by f 2 2 . The inequality | T λ ϕ ( f ) | C | λ | γ f 2 2 f 1 f 3 holds for γ = 1 2 , but not for any strictly larger exponent. This is seen by choosing f 2 to be the indicator function of [ 0 , π 4 | λ | 1 ] and f 1, f 3 each to be the indicator function of [0, 1].

This example will illustrate a subtle point regarding the necessity of hypotheses in some of our main results, below.

A more systematic analysis of examples related to this one has been carried out by Gressman and Urheim [22].

Example 2.2.

More generally, consider [ 0,1 ] n e i λ ϕ ( x ) j = 1 n f j ( x j ) d x . If (1.4) holds then γ 1 2 ( n 1 ) . Indeed, by a change of variables, one can replace [0, 1] by [−1,1]. Define a j = ∂ϕ/∂x j (0). For each jn − 1 define

f j ( x j ) = e i τ λ x j 2 e i λ a j x j η ( x j )

where η is a C function supported in a small neighborhood of 0, satisfying η(0) ≠ 0. If τ is a sufficiently large constant, depending on ϕ but not on λ, and if η is supported in a sufficiently small neighborhood of 0, then by the method of stationary phase, for a certain constant c ≠ 0, as λ → +∞,

[ 0,1 ] n 1 e i λ ϕ ( x ) j = 1 n 1 f j ( x j ) d x 1 d x 2 d x n 1 = c λ ( n 1 ) / 2 + O ( λ ( n 3 ) / 2 )

uniformly for all x n in some neighborhood V of 0 independent of λ. Choose f n (x n ) to vanish outside of V and to be equal to e i h ( x n ) with h real-valued so that

e i h ( x n ) [ 0,1 ] n 1 e i λ ϕ ( x ) j = 1 n 1 f j ( x j ) d x 1 d x 2 d x n 1 0

for each x n V. Thus | T λ ϕ ( f ) | = c λ ( n 1 ) / 2 + O ( λ ( n 3 ) / 2 ) with c′ ≠ 0.

Example 2.3.

For any n ≥ 3, the exponent γ = (n − 1)/2 is realized for

ϕ ( x 1 , , x n ) = x 1 x 2 + x 2 x 3 + + x n 1 x n .

This follows from the same reasoning as for n = 3.

Example 2.4.

For ϕ(x 1, x 2, x 3) = x 1 x 2 + x 2 x 3 + x 3 x 1, the optimal exponent is γ = 1 2 . Choosing f j ( x ) = e i λ x j 2 / 2 , the integrand becomes eiλψ with ψ ( x ) = ( x 1 + x 2 + x 3 ) 2 . This net phase function ψ factors through a submersion from R 3 to R 1 , and has a critical point.

This example, contrasted with ϕ = x 1 x 2 + x 2 x 3, for which the optimal exponent is 1, demonstrates that enlarging the set of monomials that occur with nonzero coefficients can cause the optimal exponent γ to decrease, in contrast to the theory for a restricted range of parameters developed in Refs. [2], [3].

Example 2.5.

More generally, for any parameter r > 0, the optimal exponent is 1 2 for

ϕ ( x 1 , x 2 , x 3 ) = x 1 x 2 + x 2 x 3 + r x 3 x 1 .

Thus the optimal exponent is not lower semicontinuous with respect to ϕ. This also suggests that the exponent γ = (n − 1)/2 is rarely attained.

Example 2.6.

For ϕ(x) = x 1 x 2 x 3, the exponent γ = 1 2 is again optimal. Choosing f j ( x ) = e i λ ln ( x ) 1 1 2 , 1 ( x ) , the net oscillatory factor becomes eiλψ with ψ(x) = x 1 x 2 x 3 − ln(x 1 x 2 x 3). The gradient of ψ vanishes identically on the hypersurface x 1 x 2 x 3 = 1, and the integral is no better than O(|λ|−1/2).

Example 2.7.

ϕ(x) = (x 1 + x 2)x 3. This is merely Example 2.1, with the indices 1, 2, 3 permuted. Thus we have already observed that the inequality (1.4) holds with γ = 1. Here we reexamine that example from an alternative perspective. Integrating with respect to x 3 leads to

λ 1 / 2 [ 0,1 ] 2 f 1 ( x ) f 2 ( y ) F 3 ( x + y ) d x d y

where F 3 depends on λ and satisfies F 3 L 2 = O f 3 L 2 , but no stronger inequality for any L p norm of F 3 is available. No oscillatory factor remains, yet we have already shown in Example 2.1 that an upper bound with another factor of λ −1/2 does hold.

Example 2.8.

Consider ϕ(x) = x 3 φ(x 1, x 2) + ψ(x 1, x 2) where ψ is a polynomial and (x, y) ↦ φ(x, y) is a linear function that is a scalar multiple neither of x 1 nor of x 2. The same analysis as above leads to λ −1/2 multiplied by

(2.1) [ 0,1 ] 2 f 1 ( x ) f 2 ( y ) F ( φ ( x , y ) ) e i λ ψ ( x , y ) d x d y

with F 2 = O ( f 3 2 ) . Suppose that (φ, ψ) is nondegenerate in the sense that there do not exist polynomials h j satisfying

ψ ( x , y ) = h 1 ( x ) + h 2 ( y ) + h 3 ( φ ( x , y ) ) ( x , y ) .

Then according to an inequality of Li-Tao-Thiele and the present author [5], the integral (2.1) satisfies an upper bound of the form O λ δ f 1 2 f 2 2 F 2 for some[1] δ(φ, ψ) > 0. Thus | T λ ϕ ( f ) | C | λ | 1 2 δ j = 1 3 f j . An alternative proof of this inequality is developed in Section 16.

Moreover, the analysis of [5] implicitly proves that this bound holds uniformly for all sufficiently small perturbations of φ, ψ.

Example 2.9.

Consider ϕ ( x ) = x 1 x 2 + x 2 x 3 k , with N k 2 . For k = 2, (1.4) holds for every γ strictly less than 1. For k ≥ 3, it holds for every γ 1 2 + 1 k , and this exponent is optimal. This can be shown by substituting x 3 k = x ̃ 3 and using

[ 0,1 ] 3 e i λ x 1 x 2 + x 2 x 3 j = 1 3 g j ( x j ) d x C | λ | 1 / 2 g 1 g 2 g 3 2

with g 3 ( y ) = f 3 ( y 1 / k ) y 1 k 1 . That this exponent cannot be improved when k ≥ 3 can be shown by considering f 3 equal to the indicator function of [ 0 , π 4 λ 1 / k ] .

Example 2.10.

Let ϕ(x, y) = x 2 yxy 2, or more generally, any homogeneous cubic polynomial that is not a linear combination of x 3, y 3, (x + y)3. Then

[ 0,1 ] 2 e i λ ϕ ( x , y ) f 1 ( x ) f 2 ( y ) f 3 ( x + y ) d x d y C | λ | γ j f j

holds for γ = 1 4 [23]. Even for this simplest trilinear case of (1.2), the optimal exponent remains unknown.

3 Nondegeneracy and curvature

For convenience we sometimes integrate over [0, 1]3, rather than over a ball. The two formulations are often equivalent, by straightforward arguments involving partitions of unity and expansion of cutoff functions in Fourier series, resulting in unimodular factors that can be absorbed into the functions f j .

Thus we study functionals

(3.1) T λ ϕ ( f 1 , f 2 , f 3 ) = [ 0,1 ] 3 e i λ ϕ ( x ) j = 1 3 f j ( x j ) d x

with x = ( x 1 , x 2 , x 3 ) R 3 and f j : [ 0,1 ] C , and associated inequalities

(3.2) | T λ ϕ ( f ) | C | λ | γ j = 1 3 f j L .

We assume throughout the discussion that λ is positive (as may be achieved by complex conjugation if λ is initially negative) and that λ ≥ 1. For this situation, none of the results in Refs. [2], [3] yield any exponent γ strictly greater than 1 2 , and we focus on exceeding this benchmark exponent 1 2 .

In results of this type, ϕ may be regarded as an equivalence class of functions. If ϕ ̃ takes the form ϕ ̃ ( x ) = ϕ ( x ) j = 1 3 h j ( x j ) with all functions h j real-valued and Lebesgue measurable, then

sup f j 1 | T λ ϕ ( f ) | = sup f j 1 | T λ ϕ ̃ ( f ) |

since each function f j can be replaced by f j e i λ h j . Thus ϕ ̃ is equivalent to ϕ, so far as the inequality (3.2) is concerned. On the other hand, it is natural to require that the functions h j possess the same degree of regularity as is required of ϕ. The next definition is formulated in terms of maximally regular h j , but the minimally regular situation inevitably arises in the analysis.

The examples in Section 2 suggest a notion of degeneracy for phases ϕ. Write π j (x 1, x 2, x 3) = x j .

Definition 3.1.

Let U R 3 be open and nonempty. Let ϕ : U R be C ω . Let HU be a C ω hypersurface. ϕ is rank one degenerate on H if there exist C ω functions h j defined in neighborhoods of π j (U) such that the associated net phase function ϕ ̃ = ϕ j = 1 3 ( h j π j ) satisfies

(3.3) ( ϕ ̃ ) H 0 .

In this definition, H may be defined merely in some small nonempty open subset of U.

Definition 3.2.

ϕ : U R is said to be rank one degenerate on some hypersurface, or simply rank one degenerate, if there exist HU and functions h j such that (3.3) holds. ϕ : [ 0,1 ] 3 R is said to be rank one degenerate on some hypersurface if this holds for the restriction of ϕ to (0, 1)3.

If (3.3) holds, then the Hessian matrix of ϕ ̃ has rank less than or equal to 1 at each point of H, whence the term “rank one”. It is the restriction to H of the full gradient that is assumed to vanish in (3.3), rather than the gradient of the restriction.

Example 3.3.

ϕ(x) = x 1 x 2 + x 2 x 3 + x 3 x 1 is rank one degenerate on the hypersurface H defined by x 1 + x 2 + x 3 = 0. Indeed, choosing h j ( x j ) = x j 2 / 2 gives ϕ ̃ ( x ) = ( x 1 + x 2 + x 3 ) 2 / 2 , whose gradient vanishes on H.

More generally, for r ≠ 0, the rank one degenerate phases ϕ r (x) = x 1 x 2 + x 2 x 3 + rx 3 x 1 are equivalent to phases ϕ ̃ whose gradients vanish along hyperplanes H r defined by x 2 = −r(x 1 + x 3).

Example 3.4.

Let r R . ϕ(x) = x 3(x 1 + x 2) + rx 1 x 2 x 3 is not rank one degenerate. On the other hand, ϕ ( x ) = ( x 1 + x 2 + x 3 ) 2 + r x 1 x 2 x 3 is rank one degenerate if and only if r = 0.

Proposition 3.1.

If ϕC ω is rank one degenerate on a hypersurface H, then the inequality (1.4) cannot hold for any γ strictly greater than 1 2 on any ball B containing H.

Proof.

If the Hessian of ϕ does not vanish identically on H then there exists a relatively open subset H ̃ of H on which this Hessian has rank 1. Choose f j ( x j ) = e i λ h j ( x j ) , multiplied by cutoff functions that localize ∏ j f j (x j ) to a neighborhood of H ̃ , and invoke asymptotics provided by the method of stationary phase. The same reasoning applies so long as ϕ is not an affine function on [0, 1]3, by fibering a neighborhood of a point of H by line segments transverse to H, evaluating the asymptotic contribution of each line segment as |λ| → ∞, and integrating with respect to a transverse parameter. □

Example 3.5.

For any multi-index α N 3 , the phase function ϕ ( x ) = x α = j = 1 3 x j α j is rank one degenerate on every open subset of ( R \ { 0 } ) 3 . Therefore ϕ does not satisfy (1.4) with γ > 1 2 on any domain B.

We will also study integrals of the form

(3.4) R 2 e i λ ψ ( x ) j = 1 3 ( f j φ j ) ( x ) η ( x ) d x

with ψ , φ j : R 2 R 1 real analytic, λ R , and f j in Lebesgue spaces or Sobolev spaces of negative order. η C ( R 2 ) will be a compactly supported smooth cutoff function. Both the situations in which λ is a large parameter, and that in which λ = 0, are of interest.

The concepts of a 3-web, and its curvature, are relevant here. A 3-web in R 2 is by definition a 3-tuple of pairwise transverse smooth foliations of a connected open subset of R 2 with one-dimensional leaves [20], [24]. If φ i : R 2 R 1 are smooth functions, and if ∇φ j (x) and ∇φ k (x) are linearly independent at every point x for each pair of distinct indices j, k, then the datum (φ j : j ∈ {1, 2, 3}) defines a 3-web, whose leaves are level sets of these functions. Conversely, any 3-web is locally defined by such a tuple of functions. If φ and φ ̃ define the same foliation, then any fφ can be written as f ̃ φ ̃ , where f ̃ has L p and W s,p Sobolev norms comparable to those of f. Thus the inequalities that we will study will depend on the underlying web, rather than on the tuple (φ j ) used to describe it.

Associated to a 3-web on an open set U is its curvature, a real-valued function with domain U defined by Blaschke, and discussed in Ref. [20] and references cited there. This curvature vanishes at a point x 0 if and only if there exist smooth functions f j : R R satisfying f j ( φ j ( x 0 ) ) 0 for at least one index j, such that the associated function F = j = 1 3 f j φ j satisfies[2] F ( x ) F ( x 0 ) = O | x x 0 | 4 as xx 0. This condition depends only on the underlying 3-web, not otherwise on associated functions φ j . It is invariant under local diffeomorphism of the ambient space R 2 . The equivalence of this condition with vanishing curvature can be shown via a short calculation in local coordinates chosen so that φ j (x 1, x 2) ≡ x j for j = 1, 2.

If φ j (x) = x j for j = 1, 2, then a web defined by (φ j : j ∈ {1, 2, 3}) has curvature identically zero in an open set if and only if the ratio φ 3 / x 1 φ 3 / x 2 factors locally as the product of a function of x 1 alone with a function of x 2 alone [20].

If there exist f j such that F vanishes identically in a neighborhood of x 0 then necessarily f j ( φ j ( x 0 ) ) 0 and the change of variables x ↦ (f 1φ 1(x), f 2φ 2(x)) and the substitution φ 3f 3φ 3 transform all three functions φ i into affine functions. If φ j (x j ) ≡ x j for j = 1, 2, and if 2 φ 3 x 1 x 2 vanishes identically in a neighborhood of x 0, then φ 3 is a sum of functions of the individual coordinates. Therefore the curvature vanishes identically in a neighborhood of x 0.

We say that (φ j : j ∈ {1, 2, 3}) is equivalent to a linear system if there exist C ω real-valued functions H j , each with derivatives that do not vanish identically in any neighborhood of φ j ([0, 1]2), satisfying j = 1 3 H j φ j 0 . In this situation, ( φ ̃ j = H j φ j : j { 1,2,3 } ) defines the same 3-web as does (φ j ). Taking φ ̃ j as coordinates for j = 1, 2, all three functions φ ̃ j become linear.

If ∇φ j , ∇φ k are linearly independent at x 0 for each pair of distinct indices jk, then the curvature of the 3-web defined by (φ j ) vanishes identically in a neighborhood of x 0 if and only if (φ j : j ∈ {1, 2, 3}) is equivalent to a linear system in a neighborhood of x 0.

The following lemma connects two notions of curvature/nondegeneracy, and will be used in the proof of Theorem 4.2.

Lemma 3.2.

Suppose that in some nonempty open subset U R 2 , φC , ∂φ/∂x i vanishes nowhere for i = 1, 2, and the 3-web associated to (x 1, x 2, φ(x 1, x 2)) has nowhere vanishing curvature. Then the phase function ϕ(x 1, x 2, x 3) = x 3 φ(x 1, x 2) is not rank one degenerate in any open subset of U × ( R \ { 0 } ) .

Proof.

Write φ i = ∂φ/∂x i for i = 1, 2. Suppose that ϕ ̃ = x 3 φ ( x 1 , x 2 ) j = 1 3 h j ( x j ) has gradient identically vanishing on a smooth hypersurface H. If H can be expressed in some nonempty open set in the form x 3 = F(x 1, x 2), then x 3 φ i ( x 1 , x 2 ) h i ( x i ) for i = 1, 2. It is given that φ i does not vanish. Therefore we may form the ratio of partial derivatives φ 1/φ 2 and conclude that it can be expressed, in some nonempty open subset of U, as a product of a function of x 1 with a function of x 2. This contradicts the hypothesis of nonvanishing curvature, as shown in Ref. [20].

If ϕ ̃ has gradient identically vanishing on some smooth hypersurface H that cannot be expressed in the above form on any nonempty open set, then H must take the form Γ × I for some nonconstant curve Γ R and some interval I R of positive length. The equations x 3 φ i ( x 1 , x 2 ) h i ( x i ) force φ i (x 1, x 2) ≡ 0 on Γ, contradicting the assumption that φ i = ∂φ/∂x i vanishes nowhere. □

4 Formulations of some results

4.1 First main theorem

The first main result of this paper is concerned with multilinear expressions

(4.1) T λ ϕ ( f ) = [ 0,1 ] 3 e i λ ϕ ( x ) j = 1 3 f j ( x j ) d x .

We restrict attention here to three functions f j : [ 0,1 ] C , and integrate over [0, 1]3 rather than over a ball.

Theorem 4.1.

Let ϕ be a real analytic, real-valued function in a neighborhood U of [0, 1]3. Suppose that ϕ is not rank one degenerate on any hypersurface in U. Suppose that for each pair of distinct indices jk ∈ {1, 2, 3}, 2 ϕ x j x k vanishes nowhere on [0, 1]3. Then there exist γ > 1 2 and C < ∞ such that the operators defined in (4.1) satisfy

(4.2) T λ ϕ ( f ) C | λ | γ j f j 2

uniformly for all functions f j L 2 ( R 1 ) and all λ R .

The condition that a single partial derivative 2 ϕ x 1 x 2 vanishes nowhere suffices, for C phases ϕ without other hypotheses, to ensure that

[ 0,1 ] 2 e i λ ϕ ( x 1 , x 2 , x 3 ) j = 1 2 f j ( x j ) d x 1 d x 2 = O | λ | 1 / 2 f 1 2 f 2 2

uniformly in x 3 [1]. Consequently

T λ ϕ ( f ) = O | λ | 1 / 2 f 1 2 f 2 2 f 3 1 .

The content of Theorem 4.1 is the improvement, with appropriate norms on the right-hand side, of the exponent beyond this benchmark 1 2 .

The set of all ϕ that satisfy the hypotheses of Theorem 4.1 is nonempty, and is open with respect to the C 3 topology. The set of all 3-jets for ϕ at x 0 that guarantee validity of the hypotheses in some small neighborhood of x 0 is open and dense. Moreover, its complement is contained in a C ω variety of positive codimension in the space of jets. This is shown in Section 15.

The theorem is not valid for C phases ϕ as stated. If ϕ were merely C , then ϕ could vanish to infinite order at a single point, without any equivalent phase ϕ ̃ satisfying ϕ ̃ | H 0 for any hypersurface H. Infinite order degeneracy at a point implies that (4.2) does not hold for any γ > 0, even with L norms on the right-hand side of the inequality. Corresponding remarks apply to other results formulated in this paper.

The norms appearing on the right-hand side of (4.2) are L 2 norms, rather than L . Thus phases that satisfy the hypotheses of the theorem enjoy stronger bounds on L 2 × L 2 × L 2 than does the example ϕ(x) = x 3(x 1 + x 2), which attains the largest possible exponent, γ = 1, on L  × L  × L , but only γ = 1 2 on L 2 × L 2 × L 2. This phase satisfies the main hypothesis of rank one nondegeneracy, but fails to satisfy the auxiliary hypothesis of three nonvanishing mixed second partial derivatives.

We believe that under the rank one nondegeneracy hypothesis, the conclusion holds if one of the three mixed second partial derivatives vanishes nowhere, but the other two are merely assumed not to vanish identically. Theorem 4.4, below, supports this belief. A partial result in this direction has been obtained by Zhou [8].

Functions associated to ϕ by solutions of certain implicit equations arise naturally in our analysis, so it is not natural to restrict attention to polynomial phases in the formulation of the theorems, as is done in some works on oscillatory integral inequalities. Example 2.6 also demonstrates that for polynomial phases, it is not always natural to restrict to polynomial functions h j in formulating the equivalence relation between phases or the notion of rank one degeneracy.

4.2 Second main theorem

Oscillatory factors do not appear explicitly in the formulation of our second main result, which is concerned with conditions under which the integral of ∏ jJ (f j φ j ) is well-defined. If ηC 0 has compact support in R 2 , and if ∇φ j and ∇φ k are linearly independent at each point in the support of η for every pair of distinct indices j, k ∈ {1, 2, 3}, and if each f j L 3 / 2 ( R 1 ) , then the product η ( x ) j = 1 3 f j φ j belongs to L 1 ( R 2 ) . This is a simple consequence of complex interpolation, since the product belongs to L 1 whenever two of the three functions belong to L 1 ( R 1 ) and the third belongs to L . The exponent 3 2 is optimal in this respect. This leaves open the possibility that the integral might be well-defined when the f j belong to certain Sobolev spaces of negative orders.

For p ∈ (1, ∞) and s R , denote by W s,p the Sobolev space of all distributions having s derivatives in L p .

Question 4.1.

Let J be a finite index set. Let U R 2 nonempty and open. For jJ, let φ j : U R be C ω with nowhere vanishing gradient. Suppose that for any jkJ, ∇φ j and ∇φ k are linearly independent at almost every point in U. Let η C 0 ( U ) . Do there exist s < 0, p < ∞, and C < ∞ such that

(4.3) R 2 j J ( f j φ j ) η C j J f j W s , p

for all functions f j C 1(φ j (U))?

The answer is negative without further hypotheses. In particular, it is negative whenever all φ j are linear. But inequalities (4.3) do hold under suitable conditions, as expressed by our second main theorem.

Theorem 4.2.

Let φ j C ω for each j ∈ {1, 2, 3}. Suppose that for every pair of distinct indices jk ∈ {1, 2, 3}, ∇φ j and ∇φ k are linearly independent at x 0. Suppose that the curvature of the web defined by (φ 1, φ 2, φ 3) does not vanish at x 0. Then there exist η C 0 satisfying η(x 0) ≠ 0 such that for any exponent p > 3 2 , there exist C < ∞ and s < 0 such that

(4.4) R 2 j = 1 3 ( f j φ j ) η C j f j W s , p  for all  f ( L p ( R 1 ) ) 3 .

The assumption that fL 3/2 guarantees absolute convergence of the integral. The particular instance of Theorem 4.2 with the ordered triple ( x 1 , x 2 ) x 1 , x 1 + x 2 , x 1 + x 2 2 of mappings was treated by Bourgain [9]. An alternative proof of Theorem 4.2 has subsequently been developed by Evans [25].

In this paper, we derive Theorem 4.2 as a straightforward consequence of a variant of Theorem 4.1, illustrating the relationship between the two theorems. The validity of the inequality (4.2) for some exponent strictly greater than 1 2 is the crucial ingredient in this derivation. It is convenient in this derivation to have the strong transversality hypothesis that det(∇φ j (x), ∇φ k (x)) vanishes nowhere whenever jk, but this hypothesis has been relaxed in subsequent work [18] to the condition that det(∇φ j (x), ∇φ k (x)) does not vanish identically.

It is significant that the deduction relies on the appearance of L 2 norms, rather than merely L norms, on the right-hand side of (4.2). The tuple Φ = (φ 1, φ 2, φ 3) = (x 1, x 2, x 1 + x 2) illustrates this relatively delicate distinction. Being linear, Φ does not satisfy the inequality (4.4). When the analysis used below to reduce Theorem 4.2 to (4.2) is applied to it, the phase that arises is ϕ(x 1, x 2, x 3) = x 3(x 1 + x 2). This is Example 2.1, for which the L inequality holds with γ = 1, but the L 2 inequality (4.2) holds only for γ = 1 2 , not for any larger exponent.

This paper is organized so that Theorem 4.2 is proved along with related results, including Theorems 4.1 and 4.4. A more direct and somewhat simpler proof of Theorem 4.2 can be extracted from the discussion; see [18], where the auxiliary hypotheses are relaxed; det(∇φ j , ∇φ k ) and the curvature of the web are allowed in that extension to vanish on analytic varieties of positive codimension.

4.3 Third main theorem

Our third main result is concerned with functionals of the form

(4.5) S λ ( f ) = [ 0,1 ] 2 e i λ ψ ( x ) j = 1 3 ( f j φ j ) ( x ) d x ,

with both explicit oscillatory factors eiλψ and mappings φ j : R 2 R 1 .

Theorem 4.3.

Let φ j : [ 0,1 ] 2 R and ψ : [ 0,1 ] 2 R be real analytic. Suppose that for any indices jk ∈ {1, 2, 3}, det(∇φ j , ∇φ k ) does not vanish identically on [0, 1]2. Suppose that there exist no nonempty open subset U ⊂ (0, 1)2 and C ω functions h j : φ j ( U ) R satisfying

(4.6) ψ ( x ) = j = 1 3 ( h j φ j ) ( x )  for all  x U .

Then there exist δ > 0 and C < ∞ satisfying

(4.7) | S λ ( f ) | C | λ | δ j = 1 3 f j L  for all  f  and all  λ R .

If for each jk, det(∇φ j , ∇φ k ) vanishes nowhere on [0, 1]2, then

(4.8) | S λ ( f ) | C | λ | δ j = 1 3 f j L 2  for all  f  and all  λ R .

If the Jacobian determinant of x ↦ (φ j (x), φ k (x)) vanishes nowhere for each pair of distinct indices j, k, then | S λ ( f ) | = O f i 1 f j 1 f k for any permutation (i, j, k) of (1, 2, 3). Then the stronger conclusion (4.8) follows from the conclusion (4.7) by interpolation.

For linear mappings φ j , two different generalizations of this inequality were proved by Li, Tao, Thiele, and this author [5]. For this linear case, and for any tuple (φ j : j ∈ {1, 2, 3}) reducible to a linear tuple by a change of variables, Theorem 4.3 is a special case of results obtained in that work. While the method of analysis in Ref. [5] exploited linearity of φ j , in Section 16 we sketch an alternative proof of one of the two main results of Ref. [5] by the method developed here, which allows an extension to the nonlinear case.

Example 4.2.

For (φ 1, φ 2, φ 3) = (x 1, x 2, x 1 + x 2) and ψ ( x ) = x 1 2 x 2 (or equivalently) ( x 1 x 2 ) 3 ), and with the L 2 norms on the right-hand side replaced by L norms, the inequality for S λ (f) holds with δ = 1 4 , and fails for δ > 1 3 [23]. The optimal exponent δ, for L norms, is unknown for even this (simplest) example.

4.4 Variants and extensions

We next formulate a result for the special case in which ϕ is an affine function of x 3; thus ϕ(x) = x 3 φ(x 1, x 2) + ψ(x 1, x 2).

Theorem 4.4.

Let J = {1, 2, 3} and d = 1. Let φ, ψ be real-valued real analytic functions defined in a neighborhood of [0, 1]2. Define

(4.9) ϕ ( x 1 , x 2 , x 3 ) = x 3 φ ( x 1 , x 2 ) + ψ ( x 1 , x 2 ) .

Suppose that ∂φ/∂x 1 and ∂φ/∂x 2 vanish nowhere on [0, 1]2. Suppose that there exists no nonempty open subset of [0, 1]2 in which ψ can be expressed in the form

(4.10) ψ ( x 1 , x 2 ) = h 1 ( x 1 ) + h 2 ( x 2 ) + ( h 3 φ ) ( x 1 , x 2 )

for C ω functions h j . Then there exist γ > 1 2 and C < ∞ satisfying

(4.11) T λ ϕ ( f ) C | λ | γ j = 1 3 f j 2

uniformly for all functions f j L 2 ( R 1 ) and all λ R .

An example for which the conclusion (4.11) fails to hold is (φ, ψ) = (x 1 + x 2, x 1 x 2). This datum violates the hypothesis (4.10), since x 1 x 2 = 1 2 x 1 2 1 2 x 1 2 + 1 2 ( x 1 + x 2 ) 2 .

Theorem 4.4 is not quite a special case of Theorem 4.1, because the hypotheses of the latter need not be satisfied; it is not assumed in Theorem 4.4 that 2 ψ x 1 x 2 is nonzero, and therefore 2 ϕ x 1 x 2 ( 0 ) could vanish.

The next result combines an explicit oscillatory factor with negative order Sobolev norms of the factors f j in the context of Theorem 4.2.

Theorem 4.5.

Consider S λ with J = {1, 2, 3}, d = 1, and D = 2. Let φ j C ω for each j ∈ {1, 2, 3}. Suppose that for any two indices jk, ∇φ j and ∇φ k are linearly independent at every point. Suppose that there exist no nonempty open subset U ⊂ (0, 1)2 and C ω functions h j : φ j ( U ) R satisfying

(4.12) ψ ( x ) = j = 1 3 ( h j φ j ) ( x )  for  x U .

Suppose also that (φ 1, φ 2, φ 3) is not equivalent to a linear system in any nonempty open set. Then for each p > 3 2 there exist C < ∞, δ > 0, and s < 0 such that

(4.13) | S λ ( f ) | C ( 1 + | λ | ) δ j = 1 3 f j W s , p  for all  f ( L p ) 3  and all  λ R .

The proofs of these theorems are organized as follows. We begin the proofs with Theorem 4.4, reducing it in Section 5 to a special bandlimited case of Theorem 4.3. We then treat that bandlimited case in Sections 68 and 10.

Theorem 4.1 is proved in Sections 11 and 12 by elaborating on that analysis. In Section 13 we complete the proof of Theorem 4.3, and derive Theorems 4.2 and 4.5 from the same methods and results. In Section 14 we enunciate and prove extensions to a more general framework considered by Joly-Métivier-Rauch [20], in which the condition that the factors f j be constant along leaves of foliations is relaxed to mild smoothness along those leaves. Section 15 contains remarks concerning the hypotheses, demonstrating that these are satisfied generically, in an appropriate sense.

5 First reductions

We begin by showing how Theorem 4.4 follows from a bandlimited case of Theorem 4.3. Let (φ, ψ) satisfy the hypotheses of Theorem 4.4. There are two cases, depending on whether or not φ can be expressed in the form

(5.1) φ ( x , y ) H ( h 1 ( x ) + h 2 ( y ) )  on  [ 0,1 ] 2

with H, h 1, h 2C ω . If φ does take the form (5.1) then a C ω change of variables with respect to x and to y, together with replacement of φ by H ̃ φ for appropriate H ̃ , reduces matters to the case in which h 1, h 2 are linear. In these new coordinates, ψ remains C ω , and (4.10) continues to hold. This places us in the setting of Example 2.8, which was treated above as a consequence of the results of [5]. We therefore restrict attention henceforth to the case in which φ cannot be expressed in the form (5.1).

Integrate with respect to x 3 to reexpress

[ 0,1 ] 3 e i λ x 3 φ ( x 1 , x 2 ) e i λ ψ ( x 1 , x 2 ) j = 1 3 f j ( x j ) d x 1 d x 2 d x 3 = | λ | 1 / 2 [ 0,1 ] 2 e i λ ψ ( x 1 , x 2 ) f 1 ( x 1 ) f 2 ( x 2 ) F 3 ( φ ( x 1 , x 2 ) ) d x 1 d x 2

with F 3 ( t ) = | λ | 1 / 2 f 3 ̂ ( λ t ) satisfying F 3 2 = c f 3 2 . Thus

T λ ϕ ( f ) = c | λ | 1 / 2 S λ ( f 1 , f 2 , F 3 )

with S λ defined in terms of the phase function ψ, and with the ordered triple of mappings

( φ 1 , φ 2 , φ 3 ) ( x 1 , x 2 ) = ( x 1 , x 2 , φ ( x 1 , x 2 ) ) .

The hypotheses of Theorem 4.4 ensure that the hypotheses of Theorem 4.3 are satisfied by (ψ, φ 1, φ 2, φ 3), including the strong transversality condition that ∇φ j , ∇φ k are linearly independent at every point for any two distinct indices j, k. Therefore the conclusion of Theorem 4.4 for T λ ϕ ( f ) is a consequence of the conclusion of Theorem 4.3, which yields a factor of |λ|δ with δ > 0, supplementing the factor |λ|−1/2 that is already present. □

The function F 3 is |λ|-bandlimited, that is, its Fourier transform was supported in [−|λ|, |λ|]. Thus this proof relies only on this bandlimited case of Theorem 4.3.

In the following sections we will establish the conclusion of Theorem 4.3 in the O(|λ|)–bandlimited case, thus completing the proof of Theorem 4.4. The general case of Theorem 4.3 will be treated later, in Section 13. Theorem 4.4 will be used in the proof for the general case, but our treatment of the bandlimited case of Theorem 4.3 will not rely on Theorem 4.4, so the reasoning will not be circular.

We begin the proof of Theorem 4.3, for general (ψ, φ 1, φ 2, φ 3) satisfying its hypotheses, without any bandlimitedness hypothesis for the present. Thus it is given that for each jk ∈ {1, 2, 3}, ∇φ j , ∇φ k are linearly independent on the complement of a analytic variety of positive codimension. If (φ j : j ∈ {1, 2, 3}) is equivalent to a linear system, then the conclusion (4.7) holds. Indeed, suppose that ∑ j H j φ j ≡ 0. Supposing initially that the derivatives of H j vanish nowhere, the change of variables x ↦ (H 1φ 1(x), H 2φ 2(x)) reduces matters to the case in which φ j (x) ≡ x j for j = 1, 2. In these new coordinates, φ 3(x) = −x 1x 2, so the three mappings φ j constitute a linear system. The nondegeneracy hypothesis for ψ is diffeomorphism-invariant, so continues to hold. For this situation, with linear mappings φ j , the conclusion (4.7) was established in Ref. [5].

In the more general case in which derivatives of the C ω mappings H j are permitted to vanish at isolated points, and gradients ∇φ j are permitted to be pairwise linearly dependent on analytic varieties of positive codimensions, the same conclusion is reached by partitioning [0, 1]2 into finitely many good rectangles, on each of which each derivative has absolute value bounded below by |λ|δ , together with a bad set of Lebesgue measure O(|λ|δ ′) for small exponents δ, δ′ > 0. The reasoning of the preceding paragraph gives the desired bound for the contribution of each good rectangle, while the contribution of the remaining bad set is majorized by a constant multiple of its Lebesgue measure.

We claim further that in order to prove Theorem 4.3, it suffices to treat the special case in which φ j (x 1, x 2) ≡ x j for j = 1, 2, φ 3 x j vanishes nowhere on [0, 1]2 for j ∈ {1, 2}, and (φ j : j ∈ {1, 2, 3}) is not equivalent to a linear system. To justify this claim, let ɛ > 0 be a small auxiliary parameter, and partition [0, 1]2 into subcubes of sidelengths comparable to λ ɛ . Discard every subcube on which any one of the three Jacobian determinants fails to have magnitude greater than λ ɛ . The sum of the measures of these discarded subcubes is O(λ δ ) for some δ = δ(ɛ) > 0. Treat each of the remaining subcubes by reducing it to [0, 1]2 via an affine change of variables. This replaces λ by a positive power of λ, and likewise modifies φ j , ψ.

Next, make the change of variables x = (x 1, x 2) ↦ ϕ(x) = (φ 1(x), φ 2(x)), which is a local diffeomorphism because of the nonvanishing Jacobian condition. Replace φ 3 by φ 3ϕ −1, replace φ j (x) by x j for j = 1, 2, and replace ψ by ψϕ −1. The hypotheses of Theorem 4.3 continue to hold for this new system of data. ϕ([0, 1]2) is no longer equal to [0, 1]2, but is contained in a finite union of rectangles, in each of which the hypotheses of the theorem hold after affine changes of variables.

This change of variables introduces a Jacobian factor, which is a function of x rather than of individual coordinates. This Jacobian can be expanded into a Fourier series, expressing it as an absolutely convergent linear combination of products of unimodular functions of the individual coordinates. These factors can be absorbed into the functions f j . The case in which (φ j : j ∈ {1, 2, 3}) is equivalent to a linear system has already been treated.

Write D = d d x .

Definition 5.1.

Let λ ∈ (0, ∞) and N N . N , λ is the norm on the Banach space of N times continuously differentiable functions on [0, 1] given by

(5.2) f N , λ = k = 0 N λ k D k f L ( [ 0,1 ] ) .

Sections 68 and 10 are devoted to the proof of the following lemma.

Lemma 5.1.

Suppose that φ j (x j ) ≡ x j for j = 1, 2, that φ 3(x 1, x 2) is not expressible in the form h 1(x 1) + h 2(x 2), and that ψ is not expressible in the form (4.6). Then there exist N, C, δ such that for all f and every λ ≥ 1,

(5.3) S λ ( f ) C λ δ f 1 f 2 f 3 N , λ .

Conclusion of proof of Theorem 4.4.

Consider T λ ϕ ( g ) for arbitrary g with g 1, g 2L and g 3L 2. We have observed above that

(5.4) T λ ϕ ( g ) = λ 1 / 2 S λ ( f )

for a certain f which satisfies f j = g j for j = 1, 2, f 3 2 C g 3 2 , and f 3 is |λ|-bandlimited.

We may assume without loss of generality that λ > 0, by replacing ψ by −ψ if λ is initially negative. Since f 3 is λ-bandlimited, we may express f 3 = P λ (f 3), where P λ are linear smoothing operators that satisfy

(5.5) k P λ f q C q , k λ k f q  for all  f L q

uniformly for all q ∈ [1, ∞] and λ > 0, for each k ∈ {0, 1, 2, …}. Thus

(5.6) P λ f N , λ C N f

uniformly for all λ > 0 and fL .

The hypothesis that ∂φ/∂x 1 does not vanish leads immediately to an upper bound

| S λ ( f ) | C f 1 f 2 1 f 3 1 ,

and interchanging the roles of the coordinates gives a bound C f i f j 1 f k 1 for any permutation (i, j, k) of (1, 2, 3). Therefore by interpolation, since P λ is bounded on L q for all q uniformly in λ,

(5.7) S λ ( f 1 , f 2 , P λ ( f 3 ) ) C λ δ f 1 f 2 f 3 2

by Lemma 5.1. By (5.4), this completes the proof of Theorem 4.4, modulo proving Lemma 5.1. □

6 Microlocal decomposition

We decompose each f j in phase space into summands that are essentially supported in rectangles of dimensions (λ −1/2, λ 1/2) in [ 0,1 ] x × R ξ . To do this, partition [0, 1] into ≍ λ 1/2 intervals I m of lengths |I m | = λ −1/2. Let η m be C functions with each η m supported on the interval I m * of length 2λ −1/2 concentric with I m , with m η m 2 1 on [0, 1], and with d k η m /dx k = O(λ k/2) for each k ≥ 0.

For ν = (m 1, m 2) let Q ν = I m 1 × I m 2 [ 0,1 ] × [ 0,1 ] . Let z ν be the center of Q ν . To each ν are associated those intervals I m 3 for which there exists at least one point x = (x 1, x 2) ∈ Q ν such that φ ( x ) I m 3 * . Because each partial derivative ∂φ/∂x j vanishes nowhere, the number of such indices m 3 is majorized by a constant independent of λ, ν.

Let σ ∈ (0, 1] be a small quantity to be chosen at the very end of the analysis. For each interval I m , decompose f j η m 2 as

(6.1) f j η m 2 = g j , m + h j , m

with g j,m , h j,m identically zero outside of I m * ,

(6.2) g j , m ( x ) = η m ( x ) k = 1 N a j , m , k e i ξ j , m , k x | a j , m , k | = O ( f j ) , ξ j , m , k π λ 1 / 2 Z , N = λ 2 σ .

while

(6.3) h j , m ( x ) = η m ( x ) n Z b j , m , n e i π λ 1 / 2 n x n | b j , m , n | 2 1 / 2 = O ( f j ) | b j , m , n | = O λ σ f j .

Decompositions of this type were used by the author and J. Holmer, in unpublished work circa 2009, to prove upper bounds for certain generalizations of twisted convolution inequalities.

This is achieved by expanding f j η m into Fourier series

f j ( x ) η m ( x ) = 1 I m * ( x ) n Z c n e i π λ 1 / 2 n x ,

with coefficients c n that depend also on the indices j, m. Define g j,m to be the sum of all terms with | c n | > λ σ f j , multiplied by η m . Define h j,m = f j η m g j,m . By Parseval’s identity, there are at most ⌈λ 2σ ⌉ values of n for which |c n | > λ σ . Define the frequencies ξ j,m,k and associated coefficients a j,m,k to be those frequencies πλ 1/2 n and associated coefficients c n that satisfy | c n | > λ σ f 1 , with some arbitrary ordering. If there are fewer than N indices n for which | c n | > λ σ f 1 , then augment this list by introducing extra indices k so that there are exactly N terms, and set some a j,m,k = 0 for each of these extra indices. This is done purely for convenience of notation.

Define

g j = m g j , m  and  h j = m h j , m  for  j { 1,2 } .

For j = 3, this construction is modified in order to exploit the bandlimited character of f 3. Let ρ > 0 be another small parameter.[3] It follows from N-fold integration by parts that

| f 3 η m ̂ ( ξ ) | C N λ N | ξ | N f 3 N , λ ξ .

If N is chosen to satisfy Nρ −1, it follows that

| f 3 η m ̂ ( ξ ) | C N λ 1  whenever  | ξ | λ 1 + ρ .

Therefore the frequencies ξ 3,m,k defined above satisfy

(6.4) | ξ 3 , m , k | λ 1 + ρ .

Moreover, if N is chosen sufficiently large as a function of ρ, then the contribution made to h 3 by all terms b 3,m,k eikx with |k| ≥ λ 1+ρ has L 2 norm O(λ −1). Define F 3 to be the sum of all of these terms. Then f 3 is decomposed as

(6.5) f 3 = g 3 + h 3 + F 3 ,

with

(6.6) F 3 = O ( λ 1 ) ,

with g 3, h 3 enjoying all of the properties indicated above for j = 1, 2, and with the supplementary bandlimitedness property

(6.7) | n | λ 1 + ρ

for all frequencies n appearing in terms η m (x)b 3,m,n einx , as well as for all frequencies ξ 3,m,k .

7 Local bound

Recall that

S λ ( F 1 , F 2 , F 3 ) = R 2 F 1 ( x 1 ) F 2 ( x 2 ) F 3 ( φ ( x 1 , x 2 ) ) e i λ ψ ( x 1 , x 2 ) d x 1 d x 2 .

Let m = ( m 1 , m 2 , m 3 ) Z 3 . Write a r , p = n | a r , n | p 1 / p , with the usual limiting interpretation for p = ∞.

Lemma 7.1.

Let ρ > 0 be a small auxiliary parameter. Let f j be functions of the form

f j ( x ) = n Z a j , n e i π λ 1 / 2 n x

with a j , n C and |a 3,n | = 0 for all |n| > λ 1+ρ . Then for each m and any permutation (j, k, l) of (1, 2, 3),

(7.1) S λ ( f 1 η m 1 , f 2 η m 2 , f 3 η m 3 ) C λ 1 + 2 ρ a j , 2 a k , 2 a l ,

Proof.

For i = 1, 2 we write φ i , ψ i as shorthand for ∂φ/∂x i , ∂ψ/∂x i , respectively. Write ν = (m 1, m 2), and recall that z ν denotes the center of Q ν = I m 1 × I m 2 . Write x = (x 1, x 2).

Let ξ = ( ξ j : j { 1,2,3 } ) R 3 , and suppose that

(7.2) max j | ξ j | λ 1 + ρ .

Consider

(7.3) I ( ξ ) = R 2 e i ξ 1 x 1 e i ξ 2 x 2 e i ξ 3 φ ( x ) e i λ ψ ( x ) η m 1 ( x 1 ) η m 2 ( x 2 ) η m 3 ( φ ( x ) ) d x .

The net phase function in this integral is

(7.4) Φ ( x ) = ξ 1 x 1 + ξ 2 x 2 + ξ 3 φ ( x ) + λ ψ ( x ) ,

whose gradient is

Φ ( x ) = ξ 1 + ξ 3 φ 1 ( x ) + λ ψ 1 ( x ) ξ 2 + ξ 3 φ 2 ( x ) + λ ψ 2 ( x ) .

If

(7.5) Φ ( z ν ) λ 2 ρ λ 1 / 2

then

(7.6) | I | C ρ , K λ K for every  K < .

Indeed, if (7.5) holds then Φ x i ( z ν ) 1 2 λ 2 ρ λ 1 / 2 for at least one index i ∈ {1, 2}. Suppose without loss of generality that this holds for i = 1. Then

(7.7) Φ x 1 ( u 1 , u 2 ) λ ρ λ 1 / 2  for every point  ( u 1 , u 2 ) Q ν * = I m 1 * × I m 2 * .

This holds because the function λψ 1 varies by at most O(λλ −1/2) over Q ν * , while the assumption (7.2) guarantees that ξ 3 φ 1 varies by at most O ( λ 1 + ρ λ 1 / 2 ) = O ( λ 1 2 + ρ ) . Integrating by parts CKρ −1 times with respect to the x 1 coordinate and invoking (7.7) then yields (7.6).

Now writing n = ( n 1 , n 2 , n 3 ) Z 3 ,

S λ ( f 1 η m 1 , f 2 η m 2 , f 3 η m 3 ) = n 1 , n 2 , n 3 a 1 , n 1 a 2 , n 2 a 3 , n 3 I π λ 1 / 2 n 1 , π λ 1 / 2 n 2 , π λ 1 / 2 n 3 .

For any n = (n 1, n 2, n 3) there is the trivial bound | I ( n ) | = O | Q ν * | = O ( λ 1 ) . On the other hand, by (7.6), the n-th term is O(λ K ) if the associated phase function Φ defined by (7.4) with ξ = πλ 1/2 n satisfies | Φ ( z ν ) | λ 1 2 + 2 ρ .

For each n 1, there are O(λ 2ρ ) pairs (n 2, n 3) for which ξ = πλ 1/2 n fails to satisfy (7.5). This follows from the form of ∇Φ and the assumption that both partial derivatives φ 1, φ 2 are nowhere vanishing. The same holds with the roles of n 1, n 2, n 3 permuted in an arbitrary way. Since the total number of all tuples (m, n) is O(λ 3+2ρ ), the conclusion of the lemma follows directly from these facts by invoking (7.6) with K sufficiently large. □

8 Reduction to sublevel set bound

Let f 1, f 2 be decomposed as f j = g j + h j as in (6.1), (6.2), and let f 3 have the modified form f 3 = g 3 + h 3 + F 3 of (6.5), with the restriction (6.4). Then S λ ( f 1 , f 2 , F 3 ) = O λ 1 j = 1 3 f j , so the contribution of F 3 can be disregarded and f 3 may be replaced by f ̃ 3 = g 3 + h 3 . By summing over all cubes Q ν we conclude from Lemma 7.1 that

(8.1) | S λ ( h 1 , f 2 , f ̃ 3 ) | C λ σ λ 2 ρ j = 1 3 f j .

In the same way,

(8.2) | S λ ( g 1 , h 2 , f ̃ 3 ) | + | S λ ( g 1 , g 2 , h 3 ) | C λ σ λ 2 ρ j = 1 3 f j

so that

(8.3) | S λ ( f 1 , f 2 , f 3 ) | | S λ ( g 1 , g 2 , g 3 ) | + C λ σ λ 2 ρ j = 1 3 f j .

Thus matters are reduced to the analysis of S λ (g 1, g 2, g 3).

To complete the proof, we analyze functions of the special form

(8.4) G j ( x ) = m η m ( x ) a j , m e i x ξ j , m

with each a j , m C satisfying |a j,m | ≤ 1, and each ξ j , m R . For j = 3, we also assume

(8.5) | ξ 3 , m | λ 1 + ρ .

For each index j, g j is expressed as a sum over k j ∈ {1, 2, …, N} of functions G j of the form (8.4), multiplied by O f j . Moreover, each summand G 3 is bandlimited in the sense (8.5).

(8.6) S λ ( g 1 , g 2 , g 3 ) = O j = 1 3 f j ( k 1 , k 2 , k 3 ) | S λ G 1 , k 1 , G 2 , k 2 , G 3 , k 3 |

with the sum extending over (k 1, k 2, k 3) ∈ {1, 2, …, N}3, so that there are N 3 summands.

We will prove:

Lemma 8.1.

There exist τ 0 > 0 and C < ∞ such that for all functions of the form (8.4) satisfying also (8.5),

(8.7) | S λ ( G 1 , G 2 , G 3 ) | C λ τ 0

uniformly for all real λ ≥ 1.

Taking Lemma 8.1 for granted for the present, we can now complete the proof of Theorem 4.3 in the O(|λ|1+ρ )-bandlimited case, and hence the proof of Theorem 4.4. Applying Lemma 8.1 to each of the N 3 summands in (8.6) yields

(8.8) | S λ ( g 1 , g 2 , g 3 ) | C N 3 λ τ 0 j = 1 3 f j .

In all,

(8.9) | S λ ( f 1 , f 2 , f 3 ) | C N 3 λ τ 0 + C λ σ λ 2 ρ j = 1 3 f j C λ 6 σ λ τ 0 + C λ σ λ 2 ρ j = 1 3 f j ,

where C < ∞ depends only on φ, ψ and the auxiliary parameters σ, ρ > 0. The exponent σ remains at our disposal, while ρ may be taken to be arbitrarily small. Choosing σ = τ 0/7 gives

(8.10) | S λ ( f 1 , f 2 , f 3 ) | C λ τ j = 1 3 f j

for every τ < τ 0/7. □

We next reduce Lemma 8.1 to a sublevel set bound. Let G j have the above form for j ∈ {1, 2, 3}. By decomposing G 3 as a sum of O(1) subsums, we may assume that for each ν = (m 1, m 2) there exists at most one index m 3 = m 3(ν) for which the product η m 1 ( x 1 ) η m 2 ( x 2 ) η m 3 ( φ ( x 1 , x 2 ) ) does not vanish identically.

For ν = (m 1, m 2) and for m = (m 1, m 2, m 3(ν)), for each (x 1, x 2) ∈ Q ν define

Φ ν ( x 1 , x 2 ) = ξ 1 , m 1 x 1 + ξ 2 , m 2 x 2 + ξ 3 , m 3 φ ( x 1 , x 2 ) + λ ψ ( x 1 , x 2 ) .

Decompose S λ (G 1, G 2, G 3) as

(8.11) ν a 1 , m 1 a 2 , m 2 a 3 , m 3 e i Φ m ( x 1 , x 2 ) η m 1 ( x 1 ) η m 2 ( x 2 ) η m 3 ( φ ( x 1 , x 2 ) ) d x

with ν, m = (m 1, m 2, m 3) related as above. This sum is effectively taken over either a single index m 3 = m 3(ν), or over an empty set of indices m 3. Indices ν of the latter type may be dropped.

For each remaining ν, the integral in (8.11) is O(λ K ) for every K unless |∇Φ ν (z ν )| ≤ λ 2ρ λ 1/2.

Definition 8.1.

The sublevel set E is the union of all Q ν for which |∇Φ ν (z ν )| ≤ λ 2ρ λ 1/2.

The contribution of each such Q ν to S λ (G 1, G 2, G 3) is O ( | Q ν | j f j ) . Therefore

(8.12) | S λ ( G 1 , G 2 , G 3 ) | = O λ K + | E | j f j .

To complete the proof of Lemma 8.1 and hence the proofs of Theorems 4.3 and 4.4, it suffices to show that there exists τ 0 > 0 such that

(8.13) | E | = O λ τ 0

uniformly in all possible choices of functions m j ξ j , m j .

The main hypothesis of Theorem 4.3 – the nonexistence of C ω solutions (h j : 1 ≤ j ≤ 3) of the relation ψ = ∑ j (h j φ j ) – has not yet been invoked, except in the special case in which (φ j : 1 ≤ j ≤ 3) is equivalent to a linear system. This hypothesis is a necessary condition for validity of the conclusion. Thus the inequality (8.13) is a central element of the analysis. It is proved in Section 10.

9 Interlude

A connection between oscillatory integral bounds of the form

(9.1) B e i λ ψ j J ( f j φ j ) Θ ( λ ) j J f j ,

where Θ(λ) → 0 as |λ| → ∞, and bounds for Lebesgue measures of sublevel sets

(9.2) E = x B : ψ ( x ) j ( g j φ j ) ( x ) < ε ,

of the form

(9.3) | E | θ ( ε )

where θ(ɛ) → 0 as ɛ → 0+ with θ(ɛ) independent of (g j ), is well known. The former implies the latter: Fix an auxiliary compactly supported C function ζ : R [ 0 , ) satisfying ζ(t) = 1 for |t| ≤ 1. Then

| E | B ζ ε 1 ψ j ( g j φ j ) = R ζ ̂ ( t ) B e 2 π i ( t / ε ) ψ ( x ) j J ( f j , t φ j ) ( x ) d x d t

with f j , t = e 2 π i ( t / ε ) g j . Rewriting this as

R ε ζ ̂ ( ε λ ) B e 2 π i λ ψ ( x ) j J ( f j , ε λ φ j ) ( x ) d x d λ

and invoking (9.1) gives (9.3).

The analysis in this paper proceeds primarily in the opposite sense, using sublevel set bounds to deduce bounds for oscillatory integrals. However, the sublevel sets that arise here are variants of those defined by (9.2), in which ∇ψ appears, rather than ψ itself. The reasoning in the preceding paragraph is elaborated in 17 to establish an inverse theorem, roughly characterizing tuples g j : R 2 R for which associated sublevel sets E = x B : | j = 1 3 g j φ j ( x ) | < ε are relatively large.

Sublevel set bounds of the type (9.3), with E defined by (9.2) have been established in certain cases [26], with φ j : R D R 1 and |J| arbitrarily large relative to D, as consequences of an extension of Szemerédi’s theorem due to Furstenberg and Katznelson.

10 Proof of a sublevel set bound

Continue to denote by φ j , ψ j the partial derivatives of φ, ψ with respect to x j for j = 1, 2, respectively. The following lemma is essentially a restatement of the desired bound | E | = O λ τ 0 , with the substitutions

(10.1) h j = λ 1 m ξ j , m 1 I m

and ε = λ ρ δ 0 .

Lemma 10.1.

Let (φ, ψ) satisfy the hypotheses of Theorem 4.3. Suppose that the ordered triple of mappings (x 1, x 2) ↦ (x 1, x 2, φ(x 1, x 2)) is not equivalent to a linear system. Then there exist C < ∞ and ϱ > 0 with the following property. Let h j be real-valued Lebesgue measurable functions, and let ɛ ∈ (0, 1]. Let E be the set of all (x, y) ∈ [0, 1]2 that satisfy

(10.2) h 1 ( x ) + φ 1 ( x , y ) h 3 ( φ ( x , y ) ) + ψ 1 ( x , y ) ε h 2 ( y ) + φ 2 ( x , y ) h 3 ( φ ( x , y ) ) + ψ 2 ( x , y ) ε .

Then

(10.3) | E | C ε ϱ .

The upper bound (8.13) for the measure of the set E of Definition 8.1 is an immediate consequence of (10.3), so Lemma 10.1 suffices to complete the proof of Theorem 4.3 in the bandlimited case.

The proof of Lemma 10.1 relies on the next lemma, which should be regarded as being well known, though it is more often formulated only for the special case of families of polynomials of bounded degree, rather than for general finite dimensional real analytic families of real analytic functions. We write F ω (x) = F(x, ω).

Lemma 10.2.

Let Ω be a compact topological space, let K R D be a compact convex set with nonempty interior, and let V R D be an open set containing K. Assume that F ω C ω (V) for each ω ∈ Ω, and that the mappings ( x , ω ) x α F ( x , ω ) are continuous for every multi-index α. Suppose further that none of the functions F ω vanish identically on K. Then there exist τ > 0 and C < ∞ such that for every ɛ > 0 and every ω ∈ Ω,

(10.4) { x K : | F ω ( x ) | < ε } C ε τ .

Proof.

A simple compactness and slicing argument reduces matters to the case in which D = 1 and K has a single element. A proof for that case is implicit in proofs of van der Corput’s lemma concerning one-dimensional oscillatory integrals, for instance in Refs. [4], [27]. For a derivation as a corollary of bounds for oscillatory integrals, see [6], page 14. □

The following simple result will be used repeatedly.

Lemma 10.3.

Let (X, μ) and (Y, ν) be probability spaces. Let λ = μ × ν. Let EX × Y satisfy λ(E) > 0. Define

E ̃ = { x X : ν ( { y : ( x , y ) E } ) 1 2 λ ( E ) } .

There exists y 0Y such that

λ ( { ( x , y ) E : x E ̃  and  ( x , y 0 ) E } ) 1 8 λ ( E ) 2 .

Proof.

Since

λ E \ ( E ( E ̃ × Y ) ) Y 1 2 λ ( E ) d ν = 1 2 λ ( E ) ,

one has λ ( E ( E ̃ × Y ) ) 1 2 λ ( E ) and therefore μ ( E ̃ ) 1 2 λ ( E ) .

Consider

E * = { ( x , y , y ) : x E ̃ , ( x , y ) E ,  and  ( x , y ) E , }

which satisfies ( μ × ν × ν ) ( E * ) 1 4 λ ( E ) 2 . Indeed, by the Cauchy-Schwarz inequality,

1 4 λ ( E ) 2 λ ( E ( E ̃ × Y ) ) 2 = E ̃ Y 1 E ( x , y ) d ν ( y ) d μ ( x ) 2 E ̃ Y 1 E ( x , y ) d ν ( y ) Y 1 E ( x , y ) d ν ( y ) = ( μ × ν × ν ) ( E * ) .

The stated conclusion now follows from Fubini’s theorem. □

Proof of Lemma 10.1.

There is a C ω function κ 1(x, t) satisfying

φ ( x , κ 1 ( x , t ) ) t .

The hypothesis that ∂φ/∂x 2 vanishes nowhere implies that uniformly for all Lebesgue measurable sets A, |{(x, t): (x, κ 1(x, t)) ∈ A}| is comparable to |A|. Likewise, there exists κ 2 satisfying

φ ( κ 2 ( t , y ) , y ) t

with |{(y, t): (κ 2(t, y), y) ∈ A}| comparable to |A| for all measurable A.

Define

E 0 = { ( x , y ) E : | h 3 ( φ ( x , y ) ) | 1 } .

For N k > 0 let

E k = ( x , y ) E : 2 k 1 < | h 3 ( φ ( x , y ) ) | 2 k .

It follows immediately from (10.2) that |h 1(x)| and |h 2(y)| are O(2 k ) whenever (x, y) ∈ E k . We will show that |E k | = O(2 ɛ ϱ ). Summation with respect to k then yields (10.3).

Consider first E 0. Define

E 0 = x [ 0,1 ] : | { t : ( x , κ 1 ( x , t ) ) E 0 } | c 0 | E 0 | .

By Lemma 10.3, there exists t 0 such that the set

(10.5) E 0 = ( x , t ) : x E 0  and  ( x , κ 1 ( x , t ) ) E 0  and  ( x , κ 1 ( x , t 0 ) ) E 0

satisfies | E 0 | c | E 0 | 2 , where c > 0 is a constant that depends on the function κ 1, but not on |E 0|.

Define α = h 3(t 0). By definition of E 0, α ∈ [−1, 1]. For every (x, t) ∈ E 0,

(10.6) h 1 ( x ) + α φ 1 ( x , κ 1 ( x , t 0 ) ) + ψ 1 ( x , κ 1 ( x , t 0 ) ) ε .

Define

(10.7) h ̃ 1 ( x ) = α φ 1 ( x , κ 1 ( x , t 0 ) ) ψ 1 ( x , κ 1 ( x , t 0 ) ) .

For any ( x , t ) E 0 ,

(10.8) h ̃ 1 ( x ) + φ 1 ( x , κ 1 ( x , t ) ) h 3 ( t ) + ψ 1 ( x , κ 1 ( x , t ) ) 2 ε

by (10.6), the inequality

h 1 ( x ) + φ 1 ( x , κ 1 ( x , t ) ) h 3 ( t ) + ψ 1 ( x , κ 1 ( x , t ) ) ε  whenever  ( x , κ 1 ( x , t ) ) E 0 ,

and the triangle inequality.

The function h ̃ 1 belongs to a compact family of C ω functions of x ∈ [0, 1], parametrized by α, t 0. This family is defined solely in terms of φ, ψ. Defining

(10.9) E 0 ( 1 ) = { ( x , κ 1 ( x , t ) ) : ( x , t ) E 0 } ,

one has | E 0 ( 1 ) | c | E 0 | 2 and

(10.10) h ̃ 1 ( x ) + φ 1 ( x , y ) h 3 ( φ ( x , y ) ) + ψ 1 ( x , y ) 2 ε  for all  ( x , y ) E 0 ( 1 ) .

Repeating this reasoning with the roles of the two coordinates x, y interchanged and with E 0 replaced by E 0 ( 1 ) , we conclude that there exist a subset E 0 ( 2 ) E 0 ( 1 ) [ 0,1 ] 2 satisfying | E 0 ( 2 ) | | E 0 | 4 , and a function h ̃ 2 belonging to a compact family of C ω functions defined solely in terms of φ, ψ, that satisfy

h ̃ 2 ( y ) + φ 2 ( x , y ) h 3 ( φ ( x , y ) ) + ψ 2 ( x , y ) 2 ε  for all  ( x , y ) E 0 ( 2 ) .

The condition that (x, y) ∈ E 0 directly provides an upper bound |h 3(φ(x, y))| ≤ 1. It also implies upper bounds for |h j (x, y)| ≤ C < ∞ for j = 1, 2 via the inequalities (10.2) and the assumption that ɛ ≤ 1.

A third iteration of this reasoning yields a set E 0 ( 3 ) E 0 ( 2 ) and a function h ̃ 3 of the special form

h ̃ 3 ( x ) = [ φ 1 ( κ 2 ( x , s ) , s ) ] 1 α ψ 1 ( κ 2 ( x , s ) , s )

for some parameters s ∈ [0, 1] and α R , satisfying

(10.11) h ̃ 1 ( x ) + φ 1 ( x , y ) h ̃ 3 ( φ ( x , y ) ) + ψ 1 ( x , y ) C ε h ̃ 2 ( y ) + φ 2 ( x , y ) h ̃ 3 ( φ ( x , y ) ) + ψ 2 ( x , y ) C ε

for all ( x , y ) E 0 ( 3 ) , with | E 0 ( 3 ) | c | E 0 | 8 . Again, h ̃ 3 belongs to a compact family of C ω functions that is defined in terms of φ, ψ alone.

Define

(10.12) h 1 s , α ( x ) = α φ 1 ( x , κ 1 ( x , s ) ) h ̃ 3 ( s ) ψ 1 ( x , κ 1 ( x , s ) ) h 2 s , α ( y ) = α φ 2 ( κ 2 ( y , s ) , y ) h ̃ 3 ( s ) + ψ 2 ( κ 2 ( y , s ) , y ) h 3 s , α ( u ) = [ φ 1 ( κ 2 ( u , s ) , s ) ] 1 α ψ 1 ( κ 2 ( u , s ) , s ) .

Let F be the family of R 2 -valued C ω functions F (s,α) of (x, y) ∈ [0, 1]2, parametrized by (s, α) = (s 1, s 2, s 3, α 1, α 2, α 3) with each s j ∈ [0, 1], α 1, α 2 ∈ [−1, 1], and α 3 ∈ [−C, C] for some appropriate C < ∞, defined by

(10.13) F ( s , α ) ( x , y ) = h 1 s 1 , α 1 ( x ) + φ 1 ( x , y ) h 3 s 3 , α 2 ( φ ( x , y ) ) + ψ 1 ( x , y ) h 2 s 2 , α 2 ( y ) + φ 2 ( x , y ) h 3 s 3 , α 3 ( φ ( x , y ) ) + ψ 2 ( x , y ) .

There exist no real-valued functions h j in C 1 that satisfy

(10.14) h 1 ( x ) + φ 1 ( x , y ) h 3 ( φ ( x , y ) ) + ψ 1 ( x , y ) 0 h 2 ( y ) + φ 2 ( x , y ) h 3 ( φ ( x , y ) ) + ψ 2 ( x , y ) 0

on [0, 1]2. For if there were, then defining H j to be an antiderivative of h j , one would have

x , y ψ ( x , y ) H 1 ( x ) H 2 ( y ) H 3 ( φ ( x , y ) ) 0 ,

contradicting the nondegeneracy hypothesis on (φ, ψ). Therefore for any (s, α), the function F (s,α) does not vanish identically as a function of (x, y) ∈ [0, 1]2. Lemma 10.2 can now be applied to conclude that | E 0 ( 3 ) | C ε τ , with C < ∞ and τ > 0 depending only on φ, ψ. Threfore

(10.15) | E 0 | C ε τ / 8

for another constant C′ < ∞. This completes the analysis of E 0.

The same analysis yields an upper bound of the form |E k | ≤ C2 ɛ ϱ , uniformly for all k > 0. Indeed, define h ̃ j = 2 k h j for j ∈ {1, 2, 3}, and set ε ̃ = 2 k ε , to obtain

(10.16) h ̃ 1 ( x ) + φ 1 ( x , y ) h ̃ 3 ( φ ( x , y ) ) + 2 k ψ 1 ( x , y ) ε ̃ h ̃ 2 ( y ) + φ 2 ( x , y ) h ̃ 3 ( φ ( x , y ) ) + 2 k ψ 2 ( x , y ) ε ̃

for all (x, y) ∈ E k .

Compactify by considering the system of inequalities

(10.17) h 1 ( x ) + φ 1 ( x , y ) h 3 ( φ ( x , y ) ) + r ψ 1 ( x , y ) ε h 2 ( y ) + φ 2 ( x , y ) h 3 ( φ ( x , y ) ) + r ψ 2 ( x , y ) ε

for arbitrary r ∈ [0, 1] and ɛ′ ∈ [0, ɛ 0]. We may assume that ɛ 0 is as small as desired.

The situation differs from the analysis of E 0 in one respect: for (x, y) ∈ E k ,

(10.18) 1 2 | h 3 ( φ ( x , y ) ) | 1 .

The lower bound, of which we had no analogue in the analysis of E 0, will be crucial below.

By repeating the above reasoning, we find that if h j satisfy (10.17) and (10.18) on some set E then there exist functions h ̃ j drawn from a compact family of C ω functions associated to φ, ψ, that satisfy

(10.19) h ̃ 1 ( x ) + φ 1 ( x , y ) h ̃ 3 ( φ ( x , y ) ) + r ψ 1 ( x , y ) C ε h ̃ 2 ( y ) + φ 2 ( x , y ) h ̃ 3 ( φ ( x , y ) ) + r ψ 2 ( x , y ) C ε

for all ( x , y ) E ̃ , with | E ̃ | c | E | 8 . Moreover, the lower bound (10.18) implies that h 3 C 0 1 4 , provided that ɛ 0 is sufficiently small.

There exists no solution ( h ̃ j : j { 1,2,3 } ) of the system of equations

h ̃ 1 ( x ) + φ 1 ( x , y ) h ̃ 3 ( φ ( x , y ) ) + r ψ 1 ( x , y ) = 0 h ̃ 2 ( y ) + φ 2 ( x , y ) h ̃ 3 ( φ ( x , y ) ) + r ψ 2 ( x , y ) = 0  on  [ 0,1 ] 2 .

For r ≠ 0, this follows from the same reasoning as given above for r = 1 in the analysis of E 0. For r = 0, the simplified system

(10.20) h 1 ( x ) + φ 1 ( x , y ) h 3 ( φ ( x , y ) ) 0 h 2 ( y ) + φ 2 ( x , y ) h 3 ( φ ( x , y ) ) 0

admits no solutions with h 3 vanishing nowhere. For if there were such a solution, defining H j to be an antiderivative of h ̃ j and adjusting H 1 by an appropriate additive constant,

(10.21) H 3 ( φ ( x , y ) ) + H 1 ( x ) + H 2 ( y ) 0 .

If H 3 = h 3 vanishes nowhere, this contradicts the hypothesis that (x 1, x 2, φ(x 1, x 2)) is not equivalent to a linear system. Thus h 3 must vanish, contradicting the lower bound (10.18).

By the same reasoning as in the case k = 0, it follows that | E | C ( ε ) ϱ for a certain exponent ϱ > 0. Applying this with E = E k and ɛ′ = 2k ɛ gives |E k | ≤ C2 ɛ ϱ . Summing over all k ≥ 0 completes the proof of the lemma. □

This completes the proof of Theorem 4.3 in the |λ|1+ρ -bandlimited case, and hence also completes the proof of Theorem 4.4.

11 Proof of Theorem 4.1

In the deduction of Theorems 4.4 from 4.3, we were able to immediately gain a factor of |λ|−1/2 upon integration with respect to x 3, reducing matters to a self-contained situation in which a supplementary factor of |λ|δ was to be gained. In the framework of Theorem 4.1, the analysis does not split cleanly into two separate steps.

Let η ̃ be a C 0 cutoff function supported in a small neighborhood of [0, 1]3 and identically equal to 1 on [0, 1]3, such that ϕ is real analytic and continues to satisfy the linear independence hypotheses of the theorem in a neighborhood of the support of η ̃ . Modify the definition of T λ ϕ to

T λ ϕ ( f ) = R 3 e i λ ϕ ( x ) j = 1 3 f j ( x j ) η ̃ ( x ) d x .

We will show that this modified form satisfies the indicated upper bound as λ → +∞.

It suffices to prove the conclusion (4.2) with j f j 2 replaced by j f j on the right-hand side. Indeed, the assumption that 2 ϕ x 1 x 2 vanishes nowhere implies that

[ 0,1 ] 2 e i λ ϕ ( x 1 , x 2 , x 3 ) f 1 ( x 1 ) f 2 ( x 2 ) d x 1 d x 2 C | λ | 1 / 2 f 1 2 f 2 2

uniformly for all x 3. Therefore

| T λ ϕ ( f ) | C | λ | 1 / 2 f 1 2 f 2 2 f 3 1 .

Therefore by interpolation, it suffices to establish the conclusion with f 1 2 f 2 2 f 3 on the right-hand side and some exponent γ > 1 2 . By repeating this reduction with the roles of f 2, f 3 interchanged, interpolating between bounds in terms of f 1 2 f 2 1 f 3 and f 1 2 f 2 f 3 to conclude a bound in terms of f 1 2 f 2 2 f 3 , we infer that it suffices to establish the conclusion in terms of f 1 2 f 2 f 3 . Repeating this step once more reduces matters to a bound in terms of the product of L norms. Note that this reasoning requires nonvanishing of all three mixed second partial derivatives 2 ϕ x j x k , hence does not apply to ϕ = x 1 x 2 + x 2 x 3.

Write e ξ (x) = eiξx . There exists a constant A depending only on ϕ and on the choice of η ̃ such that

(11.1) | T λ ϕ ( e ξ , f 2 , f 3 ) | C N | ξ | N f 2 1 f 3 1  for every  | ξ | A λ

for every N < ∞ and every λ ≥ 1. This is proved by writing

e i ξ x 1 + i λ ϕ ( x ) = i ξ + i λ ϕ x 1 ( x ) 1 x 1 N e i ξ x 1 + i λ ϕ ( x )

and integrating by parts N times with respect to x 1 while holding x 2, x 3 fixed. The same holds with the role of x 1 taken by x 2 or x 3. As a consequence, it suffices to analyze T λ ϕ ( f ) under the bandlimitedness assumption that for each j ∈ {1, 2, 3}, f j ̂ ( ξ ) = 0 whenever |ξ| ≥ . We assume this for the remainder of the proof of Theorem 4.1.

Suppose that each function f j satisfies f j 1 . Expand each f j in the form

(11.2) f j ( x ) = m η m ( x ) k Z a j , m , k e i π λ 1 / 2 k x

with

(11.3) k | a j , m , k | 2 C <  uniformly in  j , m , λ .

Decompose f j = g j + h j + F j where F j is the sum of those terms with |k| > λ 1/2 λ ρ , h j is the sum of those terms with |k| ≤ λ 1/2 λ ρ and |a j,m,k | ≤ λ σ , and g j is the sum of all remaining terms. From the O(λ)-bandlimitedness condition of the preceding paragraph, it follows that

(11.4) F j 2 = O ( λ N )  for every  N < .

T λ ϕ ( f ) equals T λ ϕ ( g 1 + h 1 , g 2 + h 2 , g 3 + h 3 ) plus terms involving one or more of the functions F j . Each of the latter terms is O(λ N ) for every N < ∞, and may consequently be disregarded henceforth. Thus henceforth, f j = g j + h j and |k| ≤ λ 1/2 λ ρ in (11.2).

Expand

(11.5) T λ ϕ ( f ) = m k j = 1 3 a j , m j , k j e i Φ k ( x ) η m ( x ) d x

with η m ( x ) = l = 1 3 η l ( x l ) and with the net phase function

(11.6) Φ k ( x ) = π λ 1 / 2 k x + λ ϕ ( x ) ,

whose partial derivatives satisfy

(11.7) λ 1 / 2 Φ k x j = π k j + λ 1 / 2 ϕ x j  for each  j { 1,2,3 } .

Φ k x j ( x ) depends only on the single component k j of k = (k 1, k 2, k 3); this will be exploited. We will establish an upper bound for the sum of absolute values

(11.8) m k j = 1 3 | a j , m j , k j | e i Φ k ( x ) η m ( x ) d x .

For any (m, k),

(11.9) e i Φ k ( x ) η m ( x ) d x = O ( λ 3 / 2 ) .

A tuple of indices (m, k) is said to be nonstationary if

(11.10) | Φ k ( z m ) | λ ρ λ 1 / 2 ,

and otherwise is said to be stationary. For any nonstationary (m, k), repeated integration by parts gives

(11.11) e i Φ k ( x ) η m ( x ) d x = O ( λ N )  for every  N < .

The total number of ordered pairs (m, k) is O ( ( λ 1 / 2 ) 6 ) = O ( λ 3 ) . Therefore the total contribution made to (11.8) by all nonstationary (m, k) is O(λ M ) for all M < ∞.

For each (m 1, m 2, k 1) there are at most O(λ ρ ) values of m 3 that satisfy

(11.12) Φ k x 1 ( z m ) λ 1 / 2 λ ρ ,

with the standing notation m = (m 1, m 2, m 3). The condition (11.12) is independent of k 2, k 3, since Φ k x 1 ( z m ) does not depend on these quantities. The derivative x 3 Φ k x 1 vanishes nowhere and has absolute value c λ . Therefore for each (x 1, x 2, k 1), Φ k x 1 ( x 1 , x 2 , x 3 ) λ 1 / 2 λ ρ only on a single interval whose length is O(λ −1/2 λ ρ ). Such an interval intersects the support of η m 3 for at most O(λ ρ ) values of m 3. Thus for each (m 1, m 2, k 1), for every m 3 with at most O(λ ρ ) exceptions, (m, k) is nonstationary for every choice of k 2, k 3.

Likewise, for any (m 1, m 2, k 1, m 3), there are most O(λ ρ ) values of k 2 for which Φ k x 2 ( z m ) λ 1 / 2 λ ρ , and most O(λ ρ ) values of k 3 for which Φ k x 3 ( z m ) λ 1 / 2 λ ρ . Thus for each (m 1, m 2, k 1, m 3) there are at most O(λ ρ ) values of k 2 for which there exists k 3 such that (m, k) is stationary; and for any such k 2, there are at most O(λ ρ ) such k 3. Therefore for each (m 1, m 2, k 1), there are at most O(λ ) values of k 2 for which there exists (m 3, k 3) such that (m, k) is stationary; and for any such k 2, there are at most O(λ ) such pairs (m 3, k 3).

Decompose T λ ϕ ( f ) = T λ ϕ ( f 1 , f 2 , g 3 ) + T λ ϕ ( f 1 , f 2 , h 3 ) and consider the second summand. All coefficients arising in the expansion of h 3 satisfy | a 3 , m 3 , k 3 | λ σ . Therefore

| T λ ϕ ( f 1 , f 2 , h 3 ) | O ( λ N ) + C λ 3 / 2 m 1 , m 2 k 1 m 3 , k 2 , k 3 | a 1 , m 1 , k 1 a 2 , m 2 , k 2 a 3 , m 3 , k 3 | O ( λ N ) + C λ 3 / 2 λ σ m 1 , m 2 k 1 m 3 , k 2 , k 3 | a 1 , m 1 , k 1 a 2 , m 2 , k 2 |

for every N < ∞, with the inner sums over m 3, k 2, k 3 extending only over those indices such that (m, k) is stationary. Thus

(11.13) | T λ ϕ ( f 1 , f 2 , h 3 ) | O λ N + O λ 3 / 2 λ C ρ σ m 1 , m 2 k 1 , k 2 a 1 , m 1 , k 1 a 2 , m 2 , k 2 ,

with the inner sum taken only over those (k 1, k 2) for which there exist m 3, k 3 such that (m, k) is stationary.

For each (m 1, m 2, k 1), at most O(λ 2ρ ) indices k 2 appear in this sum. Likewise, for each (m 1, m 2, k 2), at most O(λ 2ρ ) indices k 1 appear. For each j, m j , the sequence a j , m j , k j belongs to 2 with respect to k j , with norm O(1). Therefore an application of Cauchy-Schwarz to the inner sum gives an upper bound

O ( λ 3 / 2 λ C ρ σ ) m 1 , m 2 O ( 1 ) + O ( λ N ) ,

which is O(λ −1/2 λ σ ) + O(λ N ) since there are O(λ 2/2) ordered pairs (m 1, m 2). The conclusion is that

(11.14) | T λ ϕ ( f 1 , f 2 , f 3 ) | | T λ ϕ ( f 1 , f 2 , g 3 ) | + O λ 1 / 2 λ C ρ σ .

Repeating this reasoning with indices permuted gives

| T λ ϕ ( f 1 , f 2 , g 3 ) | | T λ ϕ ( f 1 , g 2 , g 3 ) | + O λ 1 / 2 λ C ρ σ ,

and after one more repetition,

(11.15) | T λ ϕ ( f ) | | T λ ϕ ( g ) | + O λ 1 / 2 λ C ρ σ ,

where each component of g = (g 1, g 2, g 3) satisfies (11.2) with at most O(λ 2σ ) nonzero coefficients a j , m j , k j for each m j .

It remains to treat T λ ϕ ( g ) . By the same reasoning as in the proof of the bandlimited case of Theorem 4.3, in order to complete the proof of Theorem 4.1 it now suffices to prove an appropriate upper bound for measures of associated sublevel sets, formulated below as Lemma 12.1.

12 Sublevel set analysis for Theorem 4.1

Write j = x j and j , k 2 = 2 x j x k .

Lemma 12.1.

Suppose that for every distinct pair of indices jk ∈ {1, 2, 3}, the mixed partial derivative 2 ϕ x j x k vanishes nowhere on the support of η ̃ . Then there exist δ > 0 and C < ∞ such that for any ɛ ∈ (0, 1] and any Lebesgue measurable real-valued functions h 1, h 2, h 3, the sublevel set

(12.1) E = x : | j ϕ ( x ) h j ( x j ) | ε  for each  j { 1,2,3 }

satisfies

(12.2) | E | C ε 1 + δ .

Here we seek a bound with an exponent strictly greater than 1, whereas in Lemma 10.1 above, we merely sought an exponent greater than 0. Invoking Lemma 12.1 with ɛ = λ −1/2 λ , for ρ sufficiently small relative to δ, completes the proof of Theorem 4.1.

We may assume that h j (x j ) belongs to the range of ∇ j ϕ for each index j. By the implicit function theorem together with the hypothesis 2 ϕ x 1 x 3 0 , there exists a C ω function κ 0 satisfying

(12.3) 1 ϕ ( x 1 , x 2 , κ 0 ( x 1 , x 2 , x 3 ) ) = x 3 .

Differentiating this equation with respect to x 2 gives

κ 0 ( x ) x 2 = 1,2 2 ϕ 1,3 2 ϕ ( x 1 , x 2 , κ 0 ( x ) ) .

Therefore since 1,2 2 ϕ never vanishes, the mapping x ↦ (x 1, κ 0(x), x 3) is locally invertible.

Define

(12.4) κ ( x 1 , x 2 ) = κ 0 ( x 1 , x 2 , t ) with  t = h 1 ( x 1 ) .

Thus for each x 1, x 2κ(x 1, x 2) is a C ω function that satisfies

(12.5) 1 ϕ ( x 1 , x 2 , κ ( x 1 , x 2 ) ) = h 1 ( x 1 ) .

This function of x 2 is drawn from a compact family of C ω functions that is specified in terms of ϕ and is parametrized by (x 1, t) with x 1 ∈ [0, 1] and | t | 1 ϕ C 0 ( [ 0,1 ] ) + 1 .

Write y = ( y 1 , y 2 ) R 2 . By the nonvanishing of x 3 1 ϕ = 1,3 2 ϕ , the relation ∇1 ϕ(y 1, y 2, x 3) = h 1(y 1) + O(ɛ) implies that |x 3κ(y)| = O(ɛ). Thus | E | = O ( ε ) .

Define

(12.6) E 1 = y [ 0,1 ] 2 : | ( j ϕ ) ( y , κ ( y ) ) h j ( y j ) | C 0 ε  for each  j { 2,3 }

with the convention y 3 = κ(y) and with C 0 a sufficiently large constant. Then

E { ( x 1 , x 2 , x 3 ) : ( x 1 , x 2 ) E 1  and  | x 3 κ ( x 1 , x 2 ) | C 0 ε } .

Define E 2 in the same way that E 1 was defined, but with the roles of the coordinates x 1 and x 2 interchanged, relying on the assumption that 2,3 2 ϕ never vanishes and replacing κ in the construction by the corresponding function κ 2 defined by

(12.7) κ 2 ( x 1 , x 2 ) = κ 0 ( x 1 , x 2 , t ) with  t = h 2 ( x 2 ) .

Then

E { ( x 1 , x 2 , x 3 ) : ( x 1 , x 2 ) E 2  and  | x 3 κ 2 ( x 1 , x 2 ) | C 0 ε } .

Define

(12.8) E = E 1 E 2 .

Then

(12.9) E { ( x 1 , x 2 , x 3 ) : ( x 1 , x 2 ) E  and  | x 3 κ ( x 1 , x 2 ) | + | x 3 κ 2 ( x 1 , x 2 ) | 2 C 0 ε } ,

whence

(12.10) | E | C ε | E | .

In order to complete the proof of Lemma 12.1, it remains only to show that | E | is suitably small, as asserted in the next lemma.

Lemma 12.2.

Suppose that ϕC ω is not rank one degenerate. Suppose that for every pair of distinct indices jk, j , k 2 ϕ vanishes nowhere in a neighborhood of [0, 1]3. Then there exists δ such that for any measurable functions h j and any ɛ > 0, the set E introduced in (12.6) satisfies | E | = O ( ε δ ) .

Proof.

The first step is to replace h 2 by a C ω function, drawn from a compact family specified in terms of ϕ alone. There exists a set E R 1 satisfying | E | | E | such that for each y 2E,

| y 1 : ( y 1 , y 2 ) E | | E | .

Therefore the subset E E defined by E = ( y 1 , y 2 ) E : y 2 E satisfies | E | | E | 2 .

By Fubini’s theorem, there exists y ̄ 1 such that

| y 2 E : ( y ̄ 1 , y 2 ) E | | E | | E | .

Consider the relation 2 ϕ ( y ̄ 1 , y 2 , κ ( y ̄ 1 , y 2 ) ) = h 2 ( y 2 ) + O ( ε ) for those y 2E satisfying ( y ̄ 1 , y 2 ) E . Since ϕC ω and κ ( y ̄ 1 , y 2 ) is a C ω function of y 2, drawn from a compact family specified in terms of ϕ alone, this relation expresses h 2(y 2) as h ̃ 2 ( y 2 ) + O ( ε ) for these values of y 2, with h ̃ 2 drawn from another compact family of C ω functions. Therefore h 2 can be replaced by h ̃ 2 in the definition of E , at the cost of replacing E by its subset E and modifying the constant C 0 in that definition.

In the preceding two paragraphs, the roles of the variables y 1 and y 2 can be interchanged, since the definition of E = E 1 E 2 is invariant under this interchange. Therefore by replacing E by an appropriate subset E , satisfying | E | | E | 2 | E | 4 , we can reduce matters to the case in which h 1 is also drawn from a compact set of C ω functions specified solely in terms of ϕ.

Return to the equation ∇1 ϕ(x 1, x 2, κ(x 1, x 2)) = h 1(x 1), restricted now to ( x 1 , x 2 ) E . Since the right-hand side differs from a C ω function by O(ɛ) on E , and since ∇3(∇1 ϕ) never vanishes, the implicit function theorem can now be applied to conclude that κ differs on E by O(ɛ) from a C ω function, drawn from an appropriate compact family. Therefore by (12.5), κ can in turn be replaced by a C ω function of y ∈ [0, 1]2, at the price of replacing C 0 by a yet larger constant.

κ was defined by the relation ∇1 ϕ(x 1, x 2, κ(x 1, x 2)) − h 1(x 1) = 0. Differentiating this equation with respect to x 2 gives

1,2 2 ϕ ( x 1 , x 2 , κ ( x 1 , x 2 ) ) + 1,3 2 ϕ ( x 1 , x 2 , κ ( x 1 , x 2 ) ) κ ( x 1 , x 2 ) x 2 = 0 .

Since 1,2 2 ϕ vanishes nowhere by hypothesis, it follows that κ ( x 1 , x 2 ) x 2 vanishes nowhere. Therefore the relation

x 3 = κ ( x 1 , x 2 ) x 2 = κ ̃ ( x 1 , x 3 )

defines a C ω function κ ̃ .

The relation

3 ϕ ( x 1 , x 2 , κ ( x 1 , x 2 ) ) = h 3 ( κ ( x 1 , x 2 ) ) + O ( ε )  for  ( x 1 , x 2 ) E

can be rewritten with the aid of κ ̃ as

(12.11) 3 ϕ ( x 1 , κ ̃ ( x 1 , x 3 ) , x 3 ) = h 3 ( x 3 ) + O ( ε )  when  ( x 1 , κ ̃ ( x 3 ) ) E .

Therefore h 3 can likewise be replaced by a C ω function drawn from an appropriate compact set.

We have thus shown that under the hypotheses of Lemma 12.1, there exist E R 2 satisfying | E | C ε | E | 1 / 4 and C ω functions h ̃ j , κ belonging to appropriate compact families such that with x 3 = κ(x 1, x 2), | j ϕ ( x ) h ̃ j ( x j ) | = O ( ε ) for all x E for each j ∈ {1, 2, 3}.

With this analyticity in hand, Lemma 10.2 gives | E | ε δ unless there exists a choice of C ω functions h ̃ j , κ in the indicated families satisfying the exact equations

(12.12) j ϕ ( x ) h j ( x j )  for  j { 1,2,3 } ,

with x 3 = κ(x 1, x 2), identically in [0, 1]2. If such h j , κ do exist, then for each index j, define H j to be an antiderivative of h j . Define ϕ ̃ : [ 0,1 ] 3 R by

ϕ ̃ ( x ) = ϕ ( x ) H 1 ( x 1 ) H 2 ( x 2 ) H 3 ( x 3 ) .

The equations (12.12) imply that ϕ ̃ 0 on the graph x 3 = κ(x 1, x 2). Thus ϕ is rank one degenerate, contradicting a hypothesis of Theorem 4.1.

Therefore | E | ε δ and consequently | E | ε δ / 4 , completing the proof of Lemma 12.2. Therefore Lemma 12.1 is proved, as well. □

13 Completion of proofs of Theorems 4.3, 4.5, and 4.2

Conclusion of proof of Theorem 4.3 in the general case.

This theorem has been reduced to the situation in which (φ 1, φ 2, φ 3)(x 1, x 2) = (x 1, x 2, φ(x 1, x 2)) and in which (φ 1, φ 2, φ 3) is not equivalent to a linear system. In that situation, we denote the multilinear form under investigation by S λ ( φ , ψ ) .

Moreover, this subcase has been treated above, under a supplementary bandlimitedness hypothesis on f 3. Therefore by choosing τ to be a positive integral power of 2 and summing, it suffices to analyze S λ ( φ , ψ ) ( f 1 , f 2 , g ) with g ̂ supported in [τ, 2τ], and with τλ 1+ρ/2.

Assume that 2 φ x 1 x 2 does not vanish identically. We will prove that

(13.1) | S λ ( φ , ψ ) ( f 1 , f 2 , g ) | C τ δ g 2 j = 1 2 f j 2

under this hypothesis, finally completing the proof of Theorem 4.3 in its general case.

By Plancherel’s theorem and an affine change of variables, we may express

g ( t ) = τ 1 / 2 e i τ [ 0,1 ] f 3 ( x 3 ) e i t τ x 3 d x 3

with f 3 2 = c g 2 . Thus

S λ ( φ , ψ ) ( f 1 , f 2 , g ) = c e i τ τ 1 / 2 [ 0,1 ] 3 e i λ ψ ( x 1 , x 2 ) e i τ x 3 φ ( x 1 , x 2 ) j = 1 3 f j ( x j ) d x .

Thus

| S λ ( φ , ψ ) ( f 1 , f 2 , g ) | = c τ 1 / 2 T τ Ψ τ , λ ( f )

with

Ψ τ , λ ( x ) = x 3 φ ( x 1 , x 2 ) + τ 1 λ ψ ( x 1 , x 2 ) .

The factor τ −1 λ is λ ρ / 2 1 for large λ. Provided that λ is large, Ψ τ,λ is well approximated by x 3 φ(x 1, x 2).

If ψ is any C ω function and the partial derivatives φ x j for j = 1, 2 and 2 φ x 1 x 2 vanish nowhere on [0, 1]2, then Ψ τ,λ satisfies all hypotheses of Theorem 4.4, uniformly for all sufficiently large λ and all τλ 1+ρ/2. The proof of Theorem 4.4 relied only on a special bandlimited case of Theorem 4.3 that has already proved in full, so we may invoke Theorem 4.4 here without circularity in the reasoning. We conclude that for |λ| sufficiently large,

| T τ Ψ τ , λ ( f ) | C τ γ j f j 2

with C < ∞ and γ > 1 2 independent of λ, τ. This establishes (13.1) with δ = γ 1 2 > 0 , completing the proof of Theorem 4.3 under the supplemental hypothesis that the mixed second derivative 2 φ x 1 x 2 vanishes nowhere on [0, 1]2.

This nonvanishing hypothesis can be weakened; it suffices to assume that the partial derivative does not vanish identically on any open set. Ineed, we have already implicitly proved a more quantitative result, namely an upper bound of the form

C ( 1 + | λ | ) δ min j k 2 φ x j x k N

for some N, C < ∞ provided that φ, ψ lie in some compact (with respect to the C 3 norm) family of C ω functions.

Let ɛ 0 be a sufficiently small positive number, depending only on φ, ψ. Partition a neighborhood of the support of the cutoff function η into squares of sidelengths | λ | ε 0 . The union of those squares on which some mixed second partial derivative of φ has magnitude < | λ | ε 0 has Lebesgue measure O(|λ|ɛ ) for some ɛ > 0 that depends only on φ, ψ and the choice of ɛ 0. The number of remaining squares is O | λ | 2 ε 0 . The contribution of each such square can be analyzed by making an affine change of variables that converts it to [0, 1]2. Invoking the more quantitative result produces a bound of the form C | λ | δ C ε 0 for each. If ɛ 0 is sufficiently small, the result follows. □

Conclusion of proof of Theorem 4.5.

The roles of the indices 1, 2, 3 in Theorem 4.3 can be freely permuted by making changes of coordinates (x 1, x 2) ↦ (x 1, φ(x 1, x 2)) and ↦(x 2, φ(x 1, x 2)). Therefore the roles of the three functions can be freely interchanged in (13.1). Theorem 4.5 is an immediate consequence for p = 2. For p ( 3 2 , 2 ) it is obtained by interpolating between this result for p = 2 and the elementary result for ( p , s ) = ( 3 2 , 0 ) . □

Proof of Theorem 4.2.

It suffices to analyze the case in which two functions are in L 2 and one is in a negative order Sobolev space, that is, to prove that

(13.2) η j = 1 3 ( f j φ j ) = O f 1 2 f 2 2 f 3 H s .

A simple interpolation then completes the proof.

By introducing a partition of unity and making local changes of coordinates, we may reduce matters to the case in which φ(x) = x i for i = 1, 2, and φ = φ 3 has a mixed second partial derivative 2 φ x 1 x 2 that vanishes nowhere on the support of η.

Express

f 3 ( φ ( x 1 , x 2 ) ) = c 0 R e i τ φ ( x 1 , x 2 ) f 3 ̂ ( τ ) d τ .

It suffices to show that for large positive λ, the contribution of the interval τ ∈ [λ, 2λ] is O(λ δ ) for some δ > 0.

Substituting τ = λx 3, with x 3 ∈ [1, 2], expresses this contribution as a constant multiple of

λ 1 / 2 R 2 × [ 1,2 ] e i λ ψ ( x ) j = 1 3 g j ( x j ) η ( x 1 , x 2 ) d x

with

ψ ( x ) = x 3 φ ( x 1 , x 2 ) ,

g i = f i for i = 1, 2, and g 3 ( t ) = λ 1 / 2 f 3 ̂ ( λ 1 t ) . The function g 3 satisfies

g 3 2 C λ s f 3 H s .

According to Lemma 3.2, ψ is not rank one degenerate on the product of the support of η with [1, 2]. Moreover, for any pair of distinct indices jk ∈ {1, 2, 3}, 2 ψ x j x k vanishes nowhere on the domain of integration. For 2 ψ x j x 3 , this is equivalent to nonvanishing of φ x j , which is a hypothesis. For 2 ψ x 1 x 2 , it follows from the nonvanishing of 2 φ x 1 x 2 and of x 3. Thus ψ satisfies all hypotheses of Theorem 4.1. Therefore

R 2 × [ 1,2 ] e i λ ψ ( x ) j = 1 3 g j ( x j ) η ( x 1 , x 2 ) d x C λ γ j = 1 3 g j 2 C λ γ + 1 2 s f 1 2 f 2 2 f 3 H s

for some γ > 1 2 . If s < 0 is sufficiently close to 0, then γ > 1 2 s , and the proof is complete. □

14 Yet another variant

Let U R 2 be a nonempty open set. For j ∈ {1, 2, 3}, let X j be a C ω nowhere vanishing vector field in U. Suppose that for any distinct indices jk ∈ {1, 2, 3}, all integral curves of X j , X k intersect transversely at every point of U.

The weak convergence theorem of Joly, Métivier, and Rauch [20] is concerned with functions that satisfy g j L 2(U) and X j g j L 2(U), whereas the results stated above in Section 4 are concerned with the special case in which X j g j ≡ 0. Here we extend the results of Section 4 to this more general situation.

Theorem 14.1.

Let (X j : j ∈ {1, 2, 3}) be as above. Suppose that the curvature of the 3-web associated to (X j : j ∈ {1, 2, 3}) does not vanish at any point of U. Then for any exponent p > 3 2 and any auxiliary function η C 0 ( U ) , there exist C < ∞ and s < 0 such that

(14.1) R 2 η j = 1 3 g j C j g j W s , p + X j g j W s , p  for all  g j C 1 ( R 2 ) .

Corollary 14.2.

Let (X j : j ∈ {1, 2, 3}) be as above. Let p > 3 2 . Let g j ν , X j g j ν L p ( R 2 ) be uniformly bounded, and suppose that g j ν g j weakly as ν → ∞ for j = 1, 2, 3. Then

(14.2) j = 1 3 g j ν j = 1 3 g j  weakly as  ν

in every relatively compact open subset of U.

To deduce Theorem 14.1 from the results proved above, introduce C ω diffeomorphisms ϕ j = φ j 1 , φ j 2 from U to open subsets of R 2 , such that the curves x : φ j 1 ( x ) = t are the integral curves of X j . Write g j = F j ϕ j . Then the W s,p norms of g j and of X j g j together control the W s,p norms of F j and of F j y 2 . By simple decomposition and interpolation, it suffices to bound the integral under the assumption that for each j,

F j W s , p + M F j y 2 M W s , p 1

for some M < ∞; we may choose M as large as may be desired.

Expand F j in Fourier series with respect to the second variable:

F j ( y , t ) = n Z f j , n ( y ) e i nt .

Then

f j , n W s , p = O ( 1 + | n | ) N

with N as large as may be desired. Thus we are led to

n Z 3 η ( x ) e n ( x ) j f j , n φ j 1

with

e n ( x ) = k = 1 3 e i n k φ k 2 ( x ) .

Set η n = ηe n . These functions satisfy

η n C K = O ( 1 + | n | ) K N

for any K < ∞.

Thus it suffices to invoke a small improvement on Theorem 4.2: under the hypotheses of that theorem, there exists K < ∞ such that

(14.3) η j ( f j φ j ) C η C K j f j W s , p ,

uniformly for all C K functions η supported in a fixed compact region in which the hypotheses hold. This can be deduced from the formally more restrictive result already proved, by introducing a C partition of unity ζ α 2 to reduce to the case in which φ j (x) ≡ x j for j = 1, 2 for each α, then expanding ζ α η in Fourier series and incorporating factors e i n j x j into f j . □

15 Remarks on the nondegeneracy hypotheses

(1) Phases ϕ that satisfy the hypotheses of Theorem 4.1 exist in profusion. Given a point x ̄ [ 0,1 ] 3 , for generic tuples (a j,k , b i,j,k ) of real numbers satisfying the natural symmetry conditions, any phase satisfying

(15.1) 2 ϕ x j x k ( x ̄ ) = a j , k  and  3 ϕ x i x j x k ( x ̄ ) = b i , j , k

is rank one nondegenerate in some neighborhood of x ̄ . In other words, we claim that if ϕ is rank one degenerate in every neighborhood of x ̄ , then its second and third order partial derivatives at x ̄ must satisfy certain algebraic relations (15.3).

Restrict attention to phases whose mixed second order partial derivatives 2 ϕ x j x k , jk, are all nonzero at x ̄ . Suppose that H is a small C ω hypersurface containing x ̄ , on which ϕ ̃ vanishes identically, with ϕ ̃ ( x ) = ϕ ( x ) j h j ( x j ) . Suppose that H is represented by an equation x 3 = κ(x 1, x 2) in a neighborhood of ( x ̄ 1 , x ̄ 2 ) , with κ smooth. Thus κ ( x ̄ 1 , x ̄ 2 ) = x ̄ 3 .

Write ϕ j for ϕ x j , ϕ j,k for the corresponding second partial derivatives, and ϕ i,j,k for third order derivatives. Denote partial derivatives of κ by κ j , for j = 1, 2.

The vanishing of j ϕ ̃ at (x 1, x 2, κ(x 1, x 2)) for j = 1, 2 implies that ϕ 1(x 1, x 2, κ(x 1, x 2)) is independent of x 2 in a neighborhood of ( x ̄ 1 , x ̄ 2 ) . Therefore

ϕ 1,2 ( x 1 , x 2 , κ ( x 1 , x 2 ) ) + κ 2 ( x 1 , x 2 ) ϕ 1,3 ( x 1 , x 2 , κ ( x 1 , x 2 ) ) = 0

in a neighborhood of ( x ̄ 1 , x ̄ 2 ) . We write this relation as ϕ 1,2 + κ 2 ϕ 1,3 = 0, leaving it understood that ϕ and its partial derivatives are evaluated at (x 1, x 2, κ(x 1, x 2)) while κ is evaluated at (x 1, x 2), and that (x 1, x 2) varies within a small neighborhood of ( x ̄ 1 , x ̄ 2 ) . Likewise, ϕ 2,1 + κ 1 ϕ 2,3 = 0. Thus

(15.2) κ 2 = ϕ 1,2 ϕ 1,3 1  and  κ 1 = ϕ 2,1 ϕ 2,3 1 .

Differentiating the first of these relations with respect to x 1 and the second with respect to x 2, and invoking the relation κ 2,1 = κ 1,2, we find that

(15.3) ϕ 2,3 2 ϕ 1,2,1 ϕ 1,3 ϕ 1,2 ϕ 1,3,1 ϕ 1,3 2 ϕ 1,2,2 ϕ 2,3 ϕ 1,2 ϕ 2,3,2

at (x 1, x 2, κ(x 1, x 2)), for all (x 1, x 2) in a neighborhood of ( x ̄ 1 , x ̄ 2 ) . In particular, (15.3) holds at x ̄ .

Equation (15.3) was derived under the assumption that the third coordinate vector does not belong to the tangent space to H at x ̄ . Thus without that assumption, we conclude that if ϕ is rank one degenerate in every neighborhood of x ̄ , then at least one of three variants of (15.3), obtained from (15.3) by permuting the three coordinate variables, must hold for the partial derivatives of ϕ at x ̄ . Rank one nondegeneracy therefore holds in all sufficiently small neighborhoods of x ̄ , for generic values of second and third partial derivatives of ϕ at x ̄ .

(2) The hypotheses of Theorem 4.1, taken as a whole rather than individually, are stable with respect to small perturbations of ϕ. Indeed, the hypothesis that all three mixed second partial derivatives are nowhere vanishing is manifestly stable. A phase ϕ satisfying this auxiliary hypothesis is rank one degenerate if and only if there exist C ω functions h j (x j ) such that ∇ j ϕ(x) = h j (x j ) for every xH, for some piece of C ω hypersurface H ⊂ (0, 1)3.

An exhaustive class of candidate hypersurfaces H can be constructed, in terms of ϕ, as follows. Fix a base point x ̄ and consider hypersurfaces H x ̄ such that the third coordinate vector does not lie in the tangent space to H at x ̄ . Express H locally as a graph x 3 = κ(x 1, x 2). Determine κ ( x 1 , x ̄ 2 ) by solving the differential equation

κ x 1 ( x 1 , x ̄ 2 ) = ϕ 2,1 ϕ 2,3 1 ( x 1 , x ̄ 2 )

derived above, with initial condition κ ( x ̄ 1 , x ̄ 2 ) = x ̄ 3 . Recall that the mixed second partial derivative ϕ 2,3 vanishes nowhere, by hypothesis.

For each x 1 in a small neighborhood of x ̄ 1 , determine κ(x 1, x 2) by solving

κ x 2 ( x 1 , x 2 ) = ϕ 1,2 ϕ 1,3 1 ( x 1 , x 2 )

with the initial condition κ ( x 1 , x ̄ 2 ) determined in the preceding step. This defines a C ω hypersurface H containing x ̄ , and this is locally the only such hypersurface passing through x ̄ whose tangent space does not contain the third coordinate vector and that could potentially satisfy the condition in the definition of rank one degeneracy of ϕ. Repeating this construction twice more with suitable permutations of the coordinate indices yields three (or fewer) candidate hypersurfaces for each point x ̄ . Plainly this construction is continuous with respect to ϕ , x ̄ .

Once a hypersurface H is specified, the vanishing of the gradient of ϕ ̃ ( x ) = ϕ ( x ) j = 1 3 h j ( x j ) at each point of H determines the derivative h j at each point of R sufficiently close to x j . Thus the functions h j are completely determined in a neighborhood of x ̄ , up to additive constants. Again, these depend continuously on ϕ , x ̄ .

If ϕ satisfies the hypotheses of Theorem 4.1, if x ̄ [ 0,1 ] 3 , and if H and associated functions h j are as above, then ϕ ̃ fails to vanish identically on H, so by real analyticity, some partial derivative along H of ∇ϕ fails to vanish at x ̄ . This nonvanishing is stable under small perturbations of ϕ , x ̄ .

(3) The Example 2.5 also demonstrate that the optimal exponent 1 + δ in Lemma 12.1 is not stable with respect to perturbations of ϕ, if the auxiliary hypothesis on the nonvanishing of all three mixed partial derivatives is relaxed.

16 More on integrals with oscillatory factors

Li, Tao, Thiele, and the present author [5] investigated multilinear functionals

S λ ( f ) = R D e i λ ψ η j J ( f j φ j )

with φ j : R D R d j linear, and established bounds of the type O | λ | γ j f j , for certain tuples (ψ, (φ j : jJ)), under two different sets of hypotheses. Both sets of hypotheses were rather restrictive. In one set, it was required that d j = D − 1 for every jJ. In the other, d j = 1 for every j, and |J| < 2D. The latter result was invoked in the discussion above.

The method developed above yields an alternative proof of these results, and thus our discussion can be modified to be self-contained, with no invocation of results from [5]. The same method also makes it possible to remove the assumption that d j = 1, as we next show.

Let D > d N . Let {φ j : jJ} be a family of surjective linear mappings from R D to R d . Such a family is said to be in general position if for any subset J ̃ J satisfying 0 < | J ̃ | D / d , the linear mapping

(16.1) R D x ( φ j ( x ) : j J ̃ ) ( R d ) J ̃

is injective.

Theorem 16.1.

Let d , D N with D / d N . Let η C 0 ( R D ) . Let ψ be a real-valued C ω function defined in a neighborhood U of the support of η. Let J be a finite index set of cardinality |J| satisfying 1 | J | < 2 D / d .

Let {φ j : jJ} be a family of surjective linear mappings φ j : R D R d in general position. Suppose that ψ cannot be expressed in the form ψ = ∑ jJ h j φ j in any nonempty open set, with h j C ω .

Then there exist δ > 0 and C < ∞ such that for all λ R and all functions f j L ( R d ) , the form

(16.2) S λ ( f ) = R D e i λ ψ j J ( f j φ j ) η

satisfies

(16.3) | S λ ( f ) | C | λ | δ j J f j L .

This extends Theorem 2.1 of [5], in which it is assumed that d = 1, and that ψ is a polynomial. The polynomial hypothesis is not essential to the proof given in Ref. [5], but the restriction d = 1 is.

The simplest instance of Theorem 16.1 with d > 1 is as follows. Let B R d be a ball centered at 0. Let Q : R d × R d R be a homogeneous quadratic real-valued polynomial. To Q, associate its antisymmetric part Q * ( x , y ) = 1 2 ( Q ( x , y ) Q ( y , x ) ) . Denote by any norm on the vector space of all antisymmetric quadratic real-valued polymomials.

Corollary 16.2.

Let d ≥ 2. There exist C < ∞ and γ > 0 such that for all functions f j L 2,

(16.4) B × B e i Q ( x , y ) f 1 ( x ) f 2 ( y ) f 3 ( x + y ) d x d y C Q * γ j f j L 2 .

A variant in which Q was a more general phase function whose mixed Hessian matrix has full rank at every point was treated by the author and J. Holmer in unpublished work, with an optimal exponent γ.

Example 16.1.

Let d = 2 and Q((x 1, x 2), (y 1, y 2)) = x 1 y 2. Then

(16.5) [ 0,1 ] 2 × [ 0,1 ] 2 e i λ x 1 y 2 f 1 ( x ) f 2 ( y ) f 3 ( x + y ) d x d y C | λ | γ j f j L 2 .

Proof of Theorem 16.1.

The proof of Theorem 16.1 has the same overarching structure as the analysis developed above for Theorem 4.3. However, one step of the proof of Theorem 4.3 broke down when the mappings φ j were linear, and Theorem 2.1 of [5] was invoked, in a black box spirit, to treat the linear case. Much of the proof of Theorem 16.1 closely follows arguments above and hence will be merely sketched, but we will show in more detail how the problematic step, which arises near the end of the analysis, can be modified to handle linear mappings.

Let ρ > 0 be a small exponent, which will ultimately depend on another exponent σ introduced below, which in turn will depend on an exponent τ in a sublevel set bound (16.17). Assume without loss of generality that λ ≥ 1 and that f j 1 . Decompose

(16.6) f j ( y ) = m η m ( y ) k Z d a j , m , k e i π λ 1 / 2 k y

with each η m supported on the double of a cube of sidelength λ −1/2 and |η m | + |λ −1/2η m | = O(1), and

k | a j , m , k | 2 = O ( 1 )

uniformly in j, m, λ. Decompose f j = g j + h j + F j where F j = O ( λ N ) for all N < ∞, h j is the sum over m, k of those terms satisfying |a j,m,k | ≤ λ σ , and g j has an expansion of the same type with a j,m,k = 0 for all but at most O(λ 2σ ) indices k for each j, m. The contributions of all F j are negligible, and we may therefore henceforth replace f j by g j + h j for each index j.

Write k = ( k j : j J ) ( Z d ) J . Define the linear mapping L : ( Z d ) J R D to be the transpose of x ↦ (φ j (x): jJ); thus

(16.7) L ( k ) = j J φ j * ( k j )

where φ j * denotes the transpose of the linear mapping φ j . Writing m = (m j : jJ) and x R D , our functional can be expanded as

S λ ( f ) = m k j J a j , m j , k j I ( m , k )

with

I ( m , k ) = e i λ Φ k ( x ) ζ m ( x ) d x Φ k ( x ) = ψ ( x ) + π λ 1 / 2 L ( k ) x ζ m ( x ) = j J η m j ( φ j ( x ) ) .

While the number of indices m in play is comparable to ( λ d / 2 ) | J | , there are only O ( λ D / 2 ) indices m for which ζ m does not vanish identically. We claim that there exists θ ∈ (0, 1), which depends only on the ratio D / ( d | J | ) , such that for any m and any sequences of scalars b j (⋅),

(16.8) k j J | b j ( k j ) | | I ( m , k ) | C λ C ρ λ D / 2 j J b j l 2 1 θ b j l θ

uniformly in m, λ. Indeed,

(16.9) | I ( m , k ) | = O ( λ D / 2 )

uniformly in m, λ. For each m for which ζ m does not vanish identically, choose z m in the support of ζ m . Integrating by parts sufficiently many times gives

(16.10) | I ( m , k ) | C N λ N ( 1 + | ψ ( z m ) + L ( k ) | ) N  for every  N <

unless

(16.11) | Φ k ( z m ) | λ ρ .

Recalling that | J | D / d , consider any subset SJ of cardinality equal to D / d . If j S φ j * ( k j ) = 0 then k j = 0 for every jS, since the mapping x ↦ (φ j (x): jS) is bijective by the general position hypothesis. Therefore if N is chosen to be sufficiently large then the summation over all vectors (k j : jS) of min λ D / 2 , λ N 1 + | ψ ( z m ) + L ( k ) | N is O ( λ D / 2 λ C ρ ) , uniformly for all vectors (k j : jJ\S). It follows that

k j J | b j ( k j ) | | I ( m , k ) | C λ C ρ λ D / 2 j J \ S b j l 1 j S b j l

by (16.9), (16.10), and the general position assumption (16.1). Since | J | < 2 d 1 D , it follows by interpolation that

k j J | b j ( k j ) | | I ( m , k ) | C λ C ρ λ D / 2 j J b j l q

for some exponent q > 2. Then b j l q b j l 2 1 θ b j l θ , for some θ = θ(q) > 0, yielding (16.8).

From (16.9) and (16.8), for f j = g j + h j with the properties indicated above, there follows

(16.12) | S λ ( f ) | | S λ ( g ) | + O ( λ σ + C ρ ) .

Therefore, choosing ρ to be sufficiently small relative to σ, it suffices to analyze S λ (g).

The quantity S λ (g) can in turn be expressed as a sum of O(λ ) terms, in each of which each function g j takes the simple form

(16.13) g j ( x ) = m j a j , m j e i π λ 1 / 2 k j , m j ζ m j ( x )

with | a j , m j | = O ( 1 ) . We assume this form henceforth, at the expense of a factor O(λ ). This factor can be absorbed at the end of the proof, by choosing σ sufficiently small relative to the exponent τ that appears below, just as was done in other proofs earlier in the paper. Thus

(16.14) | S λ ( g ) | C m | I ( m , k m ) |

with k m = ( k j , m j : j J ) .

Define

(16.15) Φ ( x ) = ψ ( x ) + π λ 1 / 2 L ( k ) x .

Consider those m that are stationary in the sense that |∇Φ(z m )| ≥ λ ρ . By (16.10), the sum of the contributions of all such m is O(λ N ) for every N < ∞. Therefore in order to complete the analysis, it suffices to show that the number of m for which Φ is nonstationary, is O ( λ τ λ D / 2 ) for some exponent τ > 0.

Let B R D be any ball of finite radius. Let h j : R d R d be arbitrary Lebesgue measurable functions. Define

(16.16) E = x B : | ψ ( x ) j J ( h j φ j ) ( x ) D φ j | < ε .

Here, h j takes values in R d , and (h j φ j ) ⋅ j takes values in R D . To complete the proof of Theorem 16.1, it now suffices to show that there exist C < ∞ and τ > 0 such that

(16.17) | E | C ε τ

uniformly for all ɛ ∈ (0, 1] and all functions h j .

Assume temporarily that | J | D / d . Let J ̃ J be any subset of cardinality | J ̃ | = D / d . The general position hypothesis ensures that there exists a linear subspace V R D of dimension d such that kernel(φ j ) ⊂ V for each j J \ J ̃ , but for each j J ̃ , φ j | V is an invertible linear mapping from V to R d .

If the system of equations ψ ( x ) j J ( h j φ j ) ( x ) D φ j | < ε = 0 is restricted to any translate V + y of V, those terms h j φ j with j J \ J ̃ become constant functions of xV. For any xV, what results is an invertible linear system of d equations for d unknowns h j φ j ( x ) D φ j , with the index j running over J ̃ .

By the same reasoning as developed in the analyses of upper bounds for measures of sublevel sets above, we may conclude that there exist functions of the form H j + r j , where H j : R d R d are drawn from a compact family of C ω functions specified in terms of ψ, {φ j : jJ} alone and r j R d are constant vectors, and a set E ̃ satisfying | E ̃ | c | E | C , such that

(16.18) | ψ ( x ) j J ( H j φ j ) ( x ) + r j D φ j | < C ε  for all  x E ̃ .

We have reached the point at which the proof of Theorem 4.3 must be augmented in order to treat Theorem 16.1. Let C 0 be some finite constant. If

(16.19) | r j | C 0  for all  j J

then the functions H ̃ j = H j + r j are drawn from a compact family of C ω functions, and the same reasoning as in the proof of Theorem 4.3 can be applied to conclude that | E ̃ | C ε τ , and hence that the same holds for | E | with modified constants C, τ.

However, it is not true that there exists C 0 such that (16.19) holds. Indeed, if G j : R d R are linear functions satisfying ∑ jJ G j φ j ≡ 0, then for any t ∈ (0, ∞), replacement of h j by h j + tG j does not change the quantity ∇ψ − ∑ j (h j φ j ) ⋅ j , and consequently does not change E . If | J | > D / d then there exists a linear solution of ∑ jJ G j φ j ≡ 0, with at least one G j not identically zero. By taking t arbitrarily large, one finds that no uniform a priori bound (16.19) is available for the functions h j in terms of ψ, (φ j : jJ), and ɛ alone.

If ɛ ≤ 1, as we may assume, and if E is nonempty, then while the individual quantities r j may be large,

(16.20) j J r j D φ j = O ( 1 ) .

This follows from the condition

ψ ( x ) j ( H j φ j ) D φ j j r j D φ j < ε 1

by the triangle inequality, since ∇ψ and H j are uniformly bounded. There exist r ̃ j R D satisfying

j J r ̃ j D φ j = j J r j D φ j

and | r ̃ j | = O ( 1 ) for every jJ. Define h ̃ j = H j + r ̃ j . These modified functions define the same sublevel set as do the original h j , since

j J ( h ̃ j φ j ) D φ j j J ( h j φ j ) D φ j  on  E ̃ .

Thus we may replace h j by h ̃ j for all indices j. The functions h ̃ j are now drawn from a compact family of C ω functions determined by ψ, {φ j : jJ}. The same reasoning as in the above analyses of sublevel sets completes the proof of Theorem 16.1. □

The less singular case in which | J | < D / d can be treated by a simplified form of this reasoning. Details are omitted.

The intermediate conclusion that h j = H j + r j on a large set, with H j uniformly bounded and r j constant though not necessarily uniformly bounded, breaks down without the restriction | J | < 2 D / d . For an example, consider d = 1, D = 2 , and |J| = 4 with mappings φ 1(x) = φ 1(x 1, x 2) = x 1, φ 2(x) = x 2, φ 3(x) = x 1 + x 2, φ 4(x) = x 1x 2. Set G j (x) = 2x 2 for j = 1, 2, and = −x 2 for j = 3, 4. Then j = 1 4 G j φ j 0 . Therefore g j = G j satisfy j = 1 4 ( g j φ j ) D φ j 0 . Therefore any tuple of functions h j could be replaced by h j + tg j for any parameter t R , without changing the associated sublevel set E . Thus no upper bound at all holds for (h j : jJ) modulo constants r j , as in the above argument.

Conversely, in the context of the preceding paragraph, if ∑ j G j φ j ≡ 0 then each G j must be a polynomial of degree at most 2. Thus each g j is a polynomial of degree at most 1, though not necessarily constant. This suggests that when | J | 2 D / d , the reasoning should be modified by applying difference operators to ∇ψ − ∑ j (h j φ j ) ⋅ j , and that difference operators of higher degrees should be required as D / d increases.

17 A scalar-valued sublevel set inequality

The remainder of this paper is concerned with aspects of sublevel set inequalities. In Section 17 we reverse the flow of our analysis, deducing certain sublevel set inequalities from the oscillatory integral inequalities proved in earlier sections.

Let B R 2 be a ball of positive radius, and let φ j : B R 1 be real analytic for j ∈ {1, 2, 3}. Suppose that ∇φ j are pairwise linearly independent at each point in B. Let 0 ≤ ηC (B).

The functional equation f(x) + g(y) + h(x + y) = 0, has been widely studied. Its solutions are the ordered triples (f(x), g(y), h(x + y)) = (ax + c 1, ay + c 2, a(x + y) − c 1c 2) with a, c 1, c 2 all constant, and no others. Approximate solutions, in a certain sense, have been studied in Ref. [28]. We consider here the more general functional equation

(17.1) j = 1 3 ( f j φ j ) = 0  almost everywhere

where the mappings φ j need not be linear, and the functions f j are real-valued. We discuss related sublevel sets

(17.2) S ( f , r ) = x B : j = 1 3 ( f j φ ) ( x ) r

associated to ordered triples f of scalar-valued functions. The inequality (17.2) differs from corresponding inequalities studied and exploited in various proofs above in two ways: it is homogeneous rather than inhomogeneous, and it is a single scalar inequality, rather than a system of two scalar inequalities.

We show in this section that Theorem 4.2 has the following implication concerning the nonexistence of nontrivial solutions of the equation (17.1).

Corollary 17.1.

Let B R 2 be a closed ball of positive, finite radius. For j ∈ {1, 2, 3} let φ j C ω map a neighborhood of B to R , and suppose that ∇φ j are pairwise linearly independent at each point of B. Suppose that the curvature of the web defined by (φ j : j ∈ {1, 2, 3}) does not vanish identically on B. Let f be an ordered triple of Lebesgue measurable real-valued functions. Suppose that for each index j and each t R ,

(17.3) | { x : f j ( x ) = t } | = 0 .

Then { x B : j ( f j φ j ) ( x ) = 0 } = 0 .

A consequence is that any C ω solution f of (17.1) in any set of positive Lebesgue measure is constant. Indeed, one of the three component functions f j must fail to satisfy the hypothesis (17.3), and hence must be constant in φ j (B) since it is real analytic. Constancy of the other two component functions then follows immediately from the functional equation (17.1). □

A more quantitative statement is as follows.

Corollary 17.2.

Let B R 2 be a closed ball of positive, finite radius. For j ∈ {1, 2, 3} let φ j C ω map a neighborhood of B to R , and suppose that ∇φ j are pairwise linearly independent at each point of B. Suppose that the curvature of the web defined by (φ j : j ∈ {1, 2, 3}) does not vanish identically on B. There exist δ > 0 and C < ∞ such that for any ordered triple f of Lebesgue measurable real-valued functions and any r ∈ (0, ∞), the sublevel set S(f, r) satisfies

(17.4) | S ( f , r ) | C sup t R { x φ j ( B ) : | f j ( x ) t | r } δ

for each j ∈ {1, 2, 3}.

In Section 18 we discuss a related inequality for sublevel sets associated to expressions of the form j = 1 3 a j ( x ) ( f j φ j ) ( x ) with nonconstant coefficients a j , in the special case in which the mappings φ j are all linear.

Returning to the two corollaries formulated above, we will first prove Corollary 17.2, then will indicate how a modification of the proof gives Corollary 17.1. The following lemma will be used.

Lemma 17.3.

Let σ < 0. Let I R be a bounded interval. Then there exists C < ∞ such that for any real-valued function f L 2 ( R ) supported in a fixed bounded set, for any A ∈ (0, ∞),

(17.5) λ A 1 I e i λ f H σ 2 d λ C A sup t R { x I : | f ( x ) t | A 1 } | σ | .

Proof.

It suffices to treat the case A = 1, since the substitution λ = reduces the general case to this one.

Let h be a nonnegative Schwartz function satisfying h(y) ≥ 1 for all y ∈ [−1, 1], with h ̂ supported in [−1, 1].

λ 1 1 I e i λ f H σ 2 d λ h ( λ ) 1 I e i λ f H σ 2 d λ = h ( λ ) R e i λ f ( x ) e i x ξ 1 I ( x ) d x 2 ( 1 + ξ 2 ) σ d ξ d λ = R h ( λ ) R I × I e i λ [ f ( x ) f ( y ) ] e i ( x y ) ξ d x d y ( 1 + ξ 2 ) σ d ξ d λ = I × I R e i ( x y ) ξ ( 1 + ξ 2 ) σ d ξ A h ̂ ( A ( f ( y ) f ( x ) ) ) d x d y C A I × I | x y | 1 σ h ̂ ( A ( f ( y ) f ( x ) ) ) d x d y .

Since σ < 0, this is majorized by

C A I 2 | x y | 1 + | σ | 1 | f ( x ) f ( y ) | A 1 ( x , y ) d x d y C A sup y I I | x y | 1 + | σ | 1 | f ( x ) f ( y ) | A 1 ( x ) d x C A sup t { x I : | f ( x ) t | A 1 } | σ | .

Proof of Corollary 17.2.

It suffices to establish the conclusion in the special case in which r = 1, since replacing f j by r −1 f j reduces the general case to this one.

Fix a nonnegative C 0 cutoff function ζ. We aim for an upper bound for R 2 1 S ( f ) , 1 ζ d x d y . Let h : R [ 0 , ) be C and compactly supported, and be ≡ 1 on [−1, 1]. Consider instead the majorant

(17.6) h j ( f j φ j ) ζ .

By implementing a partition of unity, we may introduce C 0 cutoff functions satisfying j = 1 3 η j ( φ j ( x , y ) ) 1 on the support of ζ, with η j supported on an interval I j . Then (17.6) is equal to

(17.7) c R h ̂ ( λ ) R 2 j ( η j φ j ) e i λ f j φ j ζ ( x , y ) d x d y d λ .

By Theorem 4.2, there exists σ < 0 for which (17.7) is majorized by

(17.8) C r R ( 1 + λ ) 2 j = 1 3 η j e i λ f j H σ d λ C j = 1 3 R ( 1 + λ ) 2 η j e i λ f j H σ 3 d λ 1 / 3 C j = 1 3 R ( 1 + λ ) 2 η j e i λ f j H σ 2 d λ 1 / 3

since η j e i λ f j H σ η j e i λ f j L 2 = O ( 1 ) uniformly in all parameters because each f j is real-valued and η j has bounded support.

For any index j ∈ {1, 2, 3},

R ( 1 + λ ) 2 η j e i λ f j H σ 2 d λ C k = 0 2 2 k | λ | 2 k η j e i λ f j H σ 2 d λ .

To each term in this sum, apply Lemma 17.3 with A = 2 k to obtain a majorization by

C k = 0 2 2 k 2 k sup t x I j : | f j ( x ) t | 2 k | σ | C sup t { x I j : | f j ( x ) t | 1 } | σ | .

Inserting this bound into (17.8) gives

| S ( f , 1 ) | C j = 1 3 sup t j R { x φ j ( B ) : | f j ( x ) t j | 1 } | σ | / 3 .

We have implicitly proved a lemma that may be useful in future work:

Lemma 17.4.

Let σ < 0. Let η C ( R ) be supported in a closed bounded interval I R . There exists C < ∞, depending on σ, η, |I|, such that for any measurable function f : R R ,

(17.9) R ( 1 + λ 2 ) 1 η e i λ f H σ 2 d λ C sup t | { x I : | f ( x ) t | 1 } | | σ | .

Proof of Corollary 17.1.

Defining a measure μ on I j 2 by dμ(x, y) = |xy|−1+γ  dx dy, we have shown that

(17.10) λ 2 k r 1 η j e i λ f j H σ 2 d λ C 2 k r 1 μ ( x , y ) : | f j ( x ) f j ( y ) | 2 k r .

By summing over all nonnegative integers k we deduce that

(17.11) R r ( 1 + r λ ) 2 η j e i λ f j H σ 2 d λ C μ ( x , y ) I j 2 : | f j ( x ) f j ( y ) | r .

If f j satisfies the hypothesis (17.3), then μ ( ( x , y ) I j 2 : | f j ( x ) f j ( y ) | r ) 0 as r → 0+. Therefore |S(f, r)| → 0 as r → 0+. The set of points at which the equation (17.1) holds exactly is contained in S(f, r) for every r > 0, hence is a Lebesgue null set. □

18 A scalar sublevel set inequality with variable coefficients

Throughout this section, φ j : R 2 R 1 are assumed to be linear and surjective. Let Ω R 2 be a nonempty bounded open ball or parallelepiped. For j ∈ {1, 2, 3} let a j : Ω ̄ R be C ω functions. By this we mean that a j extends to a real analytic function defined in some neighborhood of Ω ̄ . To any three-tuple f = (f j : j ∈ {1, 2, 3}) of Lebesgue measurable functions f j : Ω R , and to any ɛ > 0, associate the sublevel set

(18.1) S ( f , ε ) = x Ω : j = 1 3 a j ( x ) ( f j φ j ) ( x ) < ε .

Theorem 18.1.

Let φ j : R 2 R 1 be pairwise linearly independent linear mappings. Let Ω, a j be as above. Suppose that for each j ∈ {1, 2, 3}, a j (x) ≠ 0 for every x Ω ̄ . Finally, suppose that for any nonempty open set U ⊂ Ω and any C ω functions F j : U R satisfying j = 1 3 a j ( x ) ( F j φ j ) ( x ) = 0 for every xU, all three functions F j vanish identically on U. Then there exist γ > 0 and C < ∞ such that for every ɛ > 0 and every three-tuple f of Lebesgue measurable functions satisfying

(18.2) | f 1 ( y ) | 1 y φ 1 ( Ω ) ,

the sublevel set S(f, ɛ) satisfies

(18.3) | S ( f , ε ) | C ε γ .

The conclusion seems likely to remain valid if the hypothesis that a j vanish nowhere, is relaxed to a j not vanishing identically. We emphasize that the mappings φ j are assumed in Theorem 18.1 to be linear. We plan to treat the nonlinearizable case in future work, by a quite different argument.

Several results related to Theorem 18.1 are known, for the case in which all a j are constant and all φ j are linear. Firstly, whenever ∑ j f j φ j vanishes Lebesgue almost everywhere, each f j must agree almost everywhere with an affine function. Secondly, if |∑ j f j φ j (x)| ≤ ɛ for all x ∈ Ω\E, and if |E| is sufficiently small, then there exist affine functions L j satisfying |f j (y) − L j (y)| ≤ for all yφ j (Ω)\E j with |E j | ≤ C|E|. Thirdly, for this linear constant coefficient situation, an inequality with a far weaker dependence on ɛ can be deduced from a structural theorem of Freĭman.

Proof of Theorem 18.1.

In the proof of this theorem, it suffices to treat the special case in which |f j (y)| ≤ 2 for every yφ j (Ω) and each j ∈ {1, 2, 3}, and |f 1(y)| ∈ [1, 2] for every yφ 1(Ω). Indeed, for k ≥ 0 define E k to be the set of all x E that satisfy 2 k ≤ max j |f j φ j (x)| < 2 k+1. Then E k = S(2k f, 2k ɛ). Therefore the conclusion of the special case gives |E k | ≲ 2γk ɛ γ . Summing over k yields the desired bound for | E | .

We may assume that Ω ̄ = [ 0,1 ] 2 , by partitioning a small neighborhood of Ω ̄ into finitely many cubes, making an affine change of coordinates in each, and treating each cube separately. In part of the proof we use coordinates (x, y) ∈ [0, 1] × [0, 1], and write D 1 = x and D 2 = y . By making a linear change of variables in R 2 , We may also assume without loss of generality that φ 1(x, y) ≡ x, φ 2(x, y) ≡ y, and φ 3(x, y) = x + y.

It suffices to show that there exists ɛ 0 such that the conclusion holds for all ɛ ∈ (0, ɛ 0]. It is no loss of generality to assume, as we will, that

(18.4) | E | ε δ 0

for a sufficiently small exponent δ 0 > 0. Indeed, if this assumption fails to hold then we have the stated conclusion, with γ = δ 0 and C = 1.

Rewrite the inequality characterizing E = S ( f , ε ) as

(18.5) f 3 ( x + y ) + a ( x , y ) f 1 ( x ) = b ( x , y ) f 2 ( y ) + O ( ε ) ( x , y ) E

with a = a 1/a 3 and b = −a 2/a 3. Let c 0 > 0 be small and define

(18.6) E ̃ 1 = y [ 0,1 ] : | { x [ 0,1 ] : ( x , y ) E } | c 0 | E | .

If c 0 is sufficiently small then by Fubini’s theorem and the Cauchy-Schwarz inequality,

(18.7) | E ̃ 1 | | E | 2 ε 2 δ 0 .

Henceforth we replace E by its subset E 1 = { ( x , y ) E : y E ̃ 1 } .

Let E 2 R 3 be the set of all ordered triples ( x , y , s ) R 3 such that ( x s , y + s ) E 1 and ( x , y ) E 1 . This set satisfies | E 2 | | E 1 | 2 | E | 4 by the Cauchy-Schwarz inequality. Indeed,

| E 1 | = C I | { ( x , y ) E 1 : x + y = t } | d t

where I is a bounded subinterval of R and with |⋅| denoting one-dimensional Lebesgue measure in the integral. Therefore

| E 1 | 2 C I | { ( x , y ) E 1 : x + y = t } | 2 d t = C | ( ( x , y ) , ( x , y ) ) E 1 × E 1 : x + y = x + y | ,

with the last |⋅| denoting the natural three-dimensional Lebesgue measure on the hyperplane in R 4 defined by this equation, with the constant C permitted to change from one occurrence to the next. The set of all pairs ((x, y), (x′, y′)) that satisfy x + y = x′ + y = is in measure-preserving one-to-one correspondence with E 2 via the relation (x′, y′) = (xs, y + s).

For any ( x , y , s ) E 2 ,

(18.8) f 3 ( x + y ) + a ( x s , y + s ) f 1 ( x s ) = b ( x s , y + s ) f 2 ( y + s ) + O ( ε ) ( x , y , s ) E 2 .

For any ( x , y , s ) E 2 we have the two approximate relations (18.5), (18.8). The contributions of f 3 cancel when these two relations are subtracted, leaving

(18.9) a ( x s , y + s ) f 1 ( x s ) a ( x , y ) f 1 ( x ) = b ( x s , y + s ) f 2 ( y + s ) b ( x , y ) f 2 ( y ) + O ( ε ) ( x , y , s ) E 2 .

The set E 3 of all ( x , x , s , y ) R 4 such that both (x, y, s) and (x′, y, s) belong to E 2 satisfies

(18.10) | E 3 | | E 2 | 2 | E | 8 ε 8 δ 0 .

Consider any such (x, x′, s, y). Consider the conjunction of (18.9) with the corresponding relation with (x, y, s) replaced by (x′, y, s). Express this pair of relations as the approximate matrix equation

(18.11) B ( x , x , s , y ) f 2 ( y ) f 2 ( y + s ) = A ( x , x , s , y ) + O ( ε )

in which the coefficient matrices A, B are the square matrix

(18.12) B ( x , x , s , y ) = b ( x s , y + s ) b ( x , y ) b ( x s , y + s ) b ( x , y )

and the column matrix

(18.13) A ( x , x , s , y ) = a ( x s , y + s ) f 1 ( x s ) a ( x , y ) f 1 ( x ) a ( x s , y + s ) f 1 ( x s ) a ( x , y ) f 1 ( x ) ,

respectively.

Lemma 18.2.

As a function of (x, x′, s, y), the determinant det(B) does not vanish identically.

Proof.

Assume to the contrary that det(B) ≡ 0. Then the ratio b(xs, y + s)/b(x′ − s, y + s) is independent of s, whence

2 s x ln | b ( x s , y + s ) | 0 .

Therefore b takes the form

b ( x , y ) h ( x + y ) k ( y )

for some smooth nowhere vanishing functions h, k.

Choosing f 1(x) ≡ 0, f 2(y) = k(y)−1, and f 3(z) = h(z), we have

(18.14) f 3 ( x + y ) + a ( x , y ) f 1 ( x ) b ( x , y ) f 2 ( y )

on a nonempty open set. This contradicts the hypothesis of Theorem 18.1 that the functional equation has no solution except the trivial solution f 1f 2f 3 ≡ 0. □

For any (x, x′, s, y), multiply both sides of the approximate matrix equation (18.11) by the cofactor matrix of B(x, x′, s, y) to conclude that

(18.15) det ( B ) ( x , x , s , y ) g ( y ) = A ( x , x , s , y ) + O ( ε ) ,

where A ( x , x , s , y ) is one of the two components of the product of the cofactor matrix of B(x, x′, s, y) with A(x, x′, s, y). Thus A is a linear combination of products of the given coefficients a, b, evaluated at points that are functions of (x, x′, s, y), with coefficients in [−2, 2]4. Those coefficients are the quantities f 1(xs), f 1(x), f 1(x′ − s), f 1(x′), whose dependence on (x, x′, s) is merely Lebesgue measurable and is unknown. However, A depends linearly, hence real analytically, on those coefficients.

By partitioning [0, 1]2 into finitely many smaller cubes, and identifying each subcube again with [0, 1]2 via an affine change of variables, we may assume that each coefficient a j is defined and analytic in a large fixed ball that contains [0, 1]2. Define K R 7 to be the set of all tuples θ = ( x , x , s , r ) = x , x , s , r 1 , r 2 , r 3 , r 4 such that r ∈ [−2, 2]4, (x, x′) ∈ [0, 1]2, and s ∈ [−2, 2]. K is compact and connected. Let (y, θ) vary over [0, 1] × K. Equation (18.15) can thus be written as

(18.16) det ( B ) ( y , θ ) f 2 ( y ) = A * ( y , θ ) + O ( ε )

for all (y, θ) ∈ K for which ( x , x , s , y ) E 3 , with A * a real analytic function of (y, θ) in a neighborhood of [0, 1] × K.

The set of all ( x , x , s , y ) E 3 has Lebesgue measure | E | C 0 ε C 0 δ 0 . On the other hand, since the C ω function (x, x′, s, y) ↦  det(B)(x, x′, s, y) does not vanish identically, there exists η > 0 such that

(18.17) { ( x , x , s , y ) : | det ( B ) ( x , x , s , y ) | r } r η r ( 0,1 ] .

Choose a constant C 1 R + that satisfies ηC 1 > C 0. Applying the preceding inequality with r = ε C 1 δ 0 , r η is small relative to ε C 0 δ 0 , and thus we may conclude that there exists (x, x′, s) satisfying

(18.18) y [ 0,1 ] : ( x , x , s , y ) E 3  and  | det ( B ) ( x , x , s , y ) | ε C 1 δ 0 ε δ 0 .

The conclusion is that there exists θ ̄ = θ ̄ ( f , ε ) K satisfying (18.16) for the indicated set of pairs (y, θ), with

(18.19) | det ( B ) ( x , x , s , y ) | ε C 1 δ 0 .

For such θ ̄ ,

(18.20) | f 2 ( y ) det ( B ) ( y , θ ̄ ) 1 A * ( y , θ ̄ ) | = O ( ε )

for all y in a set of measure ε δ 0 .

Revert to the initial notation, with mappings φ j and coefficients a j . The conclusion proved thus far can be summarized as follows. Let a j , φ j satisfy the hypotheses of Theorem 18.1. Let δ 0, ɛ 0 > 0 be sufficiently small. There exist a compact connected set K R 7 , and a function F 2 : [ 0,1 ] × K R that extends meromorphically to a neighborhood of [0, 1] × K, with the following property. Let f and ɛ ∈ (0, ɛ 0] satisfy the hypotheses of the theorem, as well as the auxiliary condition | S ( f , ε ) | ε δ 0 . Then there exist E S ( f , ε ) satisfying | E | | S ( f , ε ) | C , and θ ̄ = θ ̄ ( f , ε ) K , such that the triple ( f 1 , f ̃ 2 , f 3 ) defined by f ̃ 2 ( y ) = F 2 ( y , θ ̄ ) satisfies

(18.21) a 2 ( x ) f ̃ 2 ( φ 2 ( x ) ) + j 2 a j ( x ) f j ( φ j ( x ) ) = O ( ε ) x E

and

(18.22) | f 2 ( y ) f ̃ 2 ( y ) | = O ( ε ) y φ 2 ( E ) .

Moreover, the function F 2 factors as F 2(y, θ) = α(y, θ)/β(y, θ) with α, β both analytic in a neighborhood of [0, 1] × K and satisfying

(18.23) | β ( y , θ ̄ ) | ε C 1 δ 0 y φ 2 ( E ) .

This reasoning can be applied twice more in succession, with the roles of the indices j ∈ {1, 2, 3} permuted, to approximate each of f 1, f 3 by C ω functions in the same way as has been done for f 2. With each iteration, E is replaced by a subset, and one of the functions f k is replaced by an approximating meromorphic function f ̃ k ; these replacements are retained through subsequent iterations. The conclusion may be summarized as follows, incorporating a change in the meaning of the auxiliary space K.

Let a j , φ j be as in the statement of Theorem 18.1. Let δ 0, ɛ 0 > 0 be sufficiently small. There exist a compact connected set K R 21 = ( R 7 ) 3 and three C ω functions F j : [ 0,1 ] × K R , such that for any f and any ɛ ∈ (0, ɛ 0] satisfying the hypotheses of the theorem with associated sublevel set S(f, ɛ) satisfying | S ( f , ε ) | ε δ 0 , there exist a subset E S ( f , ε ) [ 0,1 ] 2 satisfying | E | | S ( f , ε ) | C and an associated parameter θ ̄ = θ ̄ ( f , ε ) K , such that the ordered triple of approximating functions ( f ̃ j : j { 1,2,3 } ) defined by f ̃ j ( y ) = F j ( y , θ ̄ ) satisfies

(18.24) j = 1 3 a j ( x ) f ̃ j ( φ j ( x ) ) = O ( ε ) x E

and

(18.25) | f j ( y ) f ̃ j ( y ) | = O ( ε ) y φ j ( E ) .

Moreover, for each j ∈ {1, 2, 3}, the function F j factors almost everywhere in its domain [0, 1] × K as

F j ( y , θ ) = α j ( y , θ ) / β j ( y , θ )

with α j , β j analytic in a neighborhood of [0, 1] × K. The denominators β j satisfy

(18.26) | β j ( y , θ ̄ ( f , ε ) ) | ε C 1 δ 0 y φ j ( E ) .

The exponents C, C 1 depend only on the data a j , φ j and the choice of ɛ 0, δ 0.

Consider the function of (x, θ) ∈ [0, 1]2 × K defined by

(18.27) H ( x , θ ) = j = 1 3 a j ( x ) α j ( φ j ( x ) , θ ) i j β i ( φ i ( x ) , θ )

along with the partial derivatives α x α H ( x , θ ) with respect to x of H, indexed by α ∈ {0, 1, 2, …}2. This function H is arrived at by multiplying j = 1 3 a j ( x ) F j ( φ j ( x ) , θ ) by i = 1 3 β i ( φ i ( x ) , θ ) in order to arrive at a function that is holomorphic, rather than merely meromorphic.

If x E then

| H ( x , θ ) | j a j ( x ) F j ( φ j ( x , θ ) ) = O ( ε )

since the functions β i are bounded. Thus in order to majorize the Lebesgue measure of the sublevel set S(f, ɛ), it will suffice to produce a satisfactory majorization of the measure of a sublevel set of x H ( x , θ ̄ ( f , ε ) ) .

To analyze sublevel sets associated to H requires information concerning H, and information concerning θ ̄ ( f , ε ) . But first, we review a happy general property (18.28) of real analytic functions that depend real analytically on auxiliary parameters. See Bourgain [29], and Stein and Street [30]. There exist N, C < ∞ such that for any multi-index satisfying |α| = N + 1, for every (x, θ) ∈ [0, 1]2 × K,

(18.28) α x α H ( x , θ ) C | β | N β x β H ( x , θ ) .

Introducing the nonnegative C ω function

(18.29) H ̃ ( x , θ ) = | β | N α H ( x , θ ) x α 2 ,

it follows from the Cauchy-Schwarz inequality that H ̃ satisfies the differential inequality

(18.30) | x H ̃ ( x , θ ) | C | H ̃ ( x , θ ) |

uniformly for all (x, θ) ∈ [0,1]2 × K. This differential inequality allows us to replace H ̃ ( x , θ ) by a function of θ alone; it implies that there exists C ∈ (0, ∞) such that the function G ( θ ) = H ̃ ( ( 0,0 ) , θ ) satisfies

(18.31) C 1 G ( θ ) H ̃ ( x , θ ) C G ( θ )  uniformly for all  ( x , θ ) [ 0,1 ] 2 × K .

GC ω in a neighborhood of K, and H(x, θ) = 0 for every x ∈ [0, 1]2 if and only if G(θ) = 0.

The following result, a variant of a lemma often attributed to van der Corput, is essentially well known.

Lemma 18.3.

Let N < ∞. Let C 1, C 2 ∈ (0, ∞). There exist C < ∞ and ρ > 0 with the following property. Let ψC N+1([0,1]2) satisfy ψ C N + 1 C 2 and

0 | α | N | α ψ ( x ) | C 1 x [ 0,1 ] 2 .

Then for any ɛ > 0,

(18.32) { x [ 0,1 ] 2 : | ψ ( x ) | ε } C ε ρ .

The upper bound on the C N+1 norm cannot be dispensed with entirely in this formulation. Consider for instance the example ψ ( x ) = ε sin ε 1 x 1 , with N = 2.

A consequence of the lemma is for any θ for which G(θ) ≠ 0, for any η ∈ (0, ∞),

(18.33) { x [ 0,1 ] 2 : | H ( x , θ ) | η G ( θ ) } C η ρ .

To complete the proof of the theorem, it would be desirable to know that G does not vanish identically on K. We will not actually prove that this is the case. Instead, note that if | S ( f , ε ) | ε δ 0 for every datum (f, ɛ) satisfying the hypotheses of the theorem, then the desired conclusion holds with γ = δ 0. Thus it suffices to treat the case in which there exists at least one datum (f, ɛ) that satisfies the reverse inequality | S ( f , ε ) | > ε δ 0 , along with the hypotheses of the theorem. We will prove that G ( θ ̄ ( f , ε ) ) 0 for any such datum, and hence may assume in the remainder of the proof that G does not vanish identically on K.

To prove that G ( θ ̄ ) 0 in this situation, with θ ̄ = θ ̄ ( f , ε ) , observe first that none of the factors β j ( y , θ ̄ ) vanishes identically as a function of y. Indeed, each such factor is ε C 1 δ 0 on a set whose Lebesgue measure is minorized by a positive quantity. By dividing by ∏ i β i (φ i (x), θ) in the definition of H, we conclude that if G ( θ ̄ ) = 0 then j = 1 3 a j ( x ) F j ( φ j ( x ) , θ ̄ ) = 0 almost everywhere as a function of x ∈ [0,1]2. By the main hypothesis of Theorem 18.1, j = 1 3 a j ( x ) F j ( φ j ( x ) , θ ̄ ) vanishes on an open set of values of x only if each function x F j ( φ j ( x ) , θ ̄ ) vanishes identically. However, the construction has | f 1 ( y ) F 1 ( y , θ ̄ ) | = O ( ε ) for y in a subset of positive measure, and by hypothesis, |f 1(y)| ∈ [1, 2] for almost every y. Therefore F 1 ( y , θ ̄ ) 0 .

Define the zero variety

(18.34) Z = { θ K : G ( θ ) = 0 } .

G is and nonnegative in a neighborhood of K, G does not vanish identically on K, and K is connected. Therefore by a theorem of Łojasiewicz [31], there exist c, τ > 0 such that

(18.35) G ( θ ) c distance ( θ , Z ) τ θ K .

If f, ɛ, S(f, ɛ) satisfy the hypotheses, then θ ̄ = θ ̄ ( f , ε ) satisfies distance ( θ ̄ , Z ) ε C δ 0 . Indeed, consider any x E . Then for y = φ 1(x), | f 1 ( y ) F 1 ( y , θ ̄ ) | = O ( ε ) and |f 1(y)| ∈ [1, 2], so | F 1 ( y , θ ̄ ) | 1 O ( ε ) 1 2 . Since F 1 = α 1/β 1, it follows that

(18.36) | α 1 ( y , θ ̄ ) | 1 2 | β 1 ( y , θ ̄ ) | ε C 1 δ 0 .

The function α 1 is real analytic with respect to both variables, hence is Lipschitz, and vanishes identically on Z. Therefore distance ( θ ̄ , Z ) ε C δ 0 , and consequently G ( θ ̄ ) ε C δ 0 . Applying (18.33) gives

(18.37) { x [ 0,1 ] 2 : | H ( x , θ ̄ ) | = O ( ε ) } = O ( ε 1 C δ 0 ρ ) .

If δ 0 is chosen to be sufficiently small then 1 − 0 > 0, so this inequality becomes

(18.38) { x [ 0,1 ] 2 : | H ( x , θ ̄ ) | = O ( ε ) } = O ( ε γ ) ,

for a certain exponent γ > 0 that depends only on the coefficients a j and the mappings φ j . This completes the proof of Theorem 18.1. □

19 A remark and a question

Continuing to assume linearity of the mappings φ j , more can be deduced from the analysis in Section 18. Drop the assumption that no nontrivial solution exists, and ask whether for any f = (f 1, f 2, f 3) and any ɛ, f can be approximated within O(ɛ) on some subset S′ ⊂ S(f, ɛ) satisfying |S′| ≳ |S(f, ɛ)| C , by an R 3 -valued function g drawn from a finite-dimensional family of C ω functions that depends only on the data (φ i , a i : i ∈ {1, 2, 3}). More generously, in light of that analysis, we allow meromorphic approximants by asking whether there exist g j and β j , drawn from such a family, such that β j does not vanish identically and β j f j g j = O(ɛ 1−ρ ) on φ j (S′). We refer to this as the approximability property.

It suffices to approximate f k by a component g k of such a g for a single index k, for then a rather simple analysis can be applied to the relation ∑ jk a j (f j φ j ) = −a k (g k φ k ) + O(ɛ); restrict this equation to level curves of φ i for each of the two indices ik in turn and exploit the transversality hypothesis.

The analysis in Section 18 shows that f 2 can be so approximated, except possibly in the special case in which a 2(x, y)/a 3(x, y) can be factored in the form h(x + y)/k(y), that is, (hφ 3)/(kφ 2). This reasoning can be repeated for any permutation of the indices 1, 2, 3. The conclusion, in invariant form with the mappings φ j assumed to be linear, is that the approximability property holds, and follows from the analysis sketched, for all but a small family of exceptional cases. Each of those exceptional cases can be transformed, by application of symmetries of the problem, to one of the two examples

(19.1) f 1 ( x ) + f 2 ( y ) + f 3 ( x + y ) = 0 .

(19.2) f 1 ( x ) + f 2 ( y ) + e x f 3 ( x + y ) = 0 .

These symmetries are linear changes of variables in R 2 and in the domains R 1 of the three mappings φ j , multiplication of the equation by an arbitrary nowhere vanishing C ω function b(x, y), and incorporation of coefficients into functions f j via multiplicative substitutions f ̃ j ( x ) = f j ( x ) u j ( x ) , with u j C ω vanishing nowhere in the relevant domain. The equation (19.1) has a three-dimensional space of exact solutions f, while (19.2) has a two-dimensional space of solutions defined by f 3(x) = c 1 e x + c 2 for arbitrary coefficients c 1 , c 2 R . In an appropriate coordinate system, the graph of an exact solution f j of (19.2) is a one-parameter subgroup of the ax + b group. □

Question 19.1.

Let ɛ > 0, and let f be measurable and satisfy |f| = O(1). Let φ j (x 1, x 2) = x 1, = x 2, and = x 1 + x 2 for j = 1, 2, 3, respectively. Let a 1(x) = 1, a 2(x) = 1, and a 3 ( x ) = e x 1 . Let S(f, ɛ) be the set of all xB satisfying | j = 1 3 a j ( x ) ( f j φ j ) ( x ) | < ε and j = 1 3 | ( f j φ j ) ( x ) | 1 .

Do there exist an exact C ω solution f* of (19.2) and a subset S′ ⊂ S(f, ɛ) satisfying |S′| ≥ c|S(f, ɛ)| C such that | f j φ j ( x ) f j * φ j ( x ) | C ε for each index j for every xS′? The constants c, C are to be independent of f, ɛ.

The answer is well known to be negative for the corresponding question for equation (19.1); counterexamples can be based on multiprogressions of ranks greater than 1.

Theorem 18.1 remains valid for mappings φ j that are real analytic with pairwise transverse gradients, rather than linear, as shown in a sequel [32] to the present work. That result is used to establish a quadrilinear analogue of Theorem 4.2 in Ref. [33]. It would be desirable to go farther, dropping the hypothesis that no exact C ω solutions of the underlying equation exist, and weakening the conclusion to approximability by exact solutions, as in Question 19.1.

20 Large sublevel sets: an example

Consider the ordered triple of submersions [ 0,1 ] 2 R defined by (x, y) ↦ x, ↦y, and ↦x + y. To any ordered triple (f, g, h) of Lebesgue measurable functions associate the sublevel set

(20.1) E = { ( x , y ) [ 0,1 ] 2 : | g ( x ) h ( x + y ) | < ε  and  | y f ( x ) h ( x + y ) | < ε }

defined by the indicated inhomogeneous system of two approximate equations for (f, g, h). One has | E | = O ( ε ) uniformly for all affine functions f, g, h. Here we show, via a construction based on multiprogressions of rank 2, that no bound better than | E | = O ( ε 1 / 2 ) is valid.

Let ɛ > 0 be small, with ε 1 / 2 N . Set N = ɛ −1/2. For each k Z , define

(20.2) f ( x ) = k ε 1 / 2 x  whenever  | x k ε 1 / 2 | < 1 2 ε 1 / 2 .

Define

(20.3) g ( y ) = k ε 1 / 2 + k ε  whenever  | y k ε 1 / 2 | < 1 2 ε 1 / 2 .

For each t R there exist unique k , n Z with 0 ≤ n < N such that | t ( k ε 1 / 2 + n ε ) | < 1 2 ε . Define

(20.4) h ( t ) = n ε 1 / 2 + n ε  whenever  | t ( k ε 1 / 2 + n ε ) | < 1 2 ε .

For m , n Z satisfying 0 ≤ n < N, define E ( m , n ) to be the set of all ( x , y ) R 2 that satisfy the three inequalities

(20.5) | y n ε 1 / 2 | < 1 2 ε 1 / 2 , | x ( m n ) ε 1 / 2 | < 1 2 ε 1 / 2 , | x + y ( m ε 1 / 2 + n ε ) | < 1 2 ε .

The sets E ( m , n ) are pairwise disjoint and satisfy

(20.6) | E ( m , n ) | = ε 3 / 2 + O ( ε 2 ) .

The number of indices ( m , n ) Z × { 0,1,2 , , N 1 } for which E ( m , n ) [ 0,1 ] 2 is c ε 1 .

If E ( m , n ) [ 0,1 ] 2 , then E ( m , n ) E . Indeed, let ( x , y ) E ( m , n ) . Firstly,

(20.7) g ( y ) h ( x + y ) = 0

since both g(y) and h(x + y) are defined to be 1/2 + in this region. Secondly,

f ( x ) + h ( x + y ) y = ( ( m n ) ε 1 / 2 x ) + h ( x + y ) y = x + y m ε 1 / 2 n ε + h ( x + y ) n ε 1 / 2 n ε .

Since x + y lies in the strip indicated in the definition of E ( m , n ) ,

| x + y m ε 1 / 2 n ε | < 1 2 ε  and  h ( x + y ) = n ε 1 / 2 + n ε .

Consequently

(20.8) | y f ( x ) h ( x + y ) | < 1 2 ε .

Thus E ( m , n ) E whenever E ( m , n ) [ 0,1 ] 2 . There are c ε 1 such sets, pairwise disjoint and satisfying | E ( m , n ) | ε 3 / 2 O ( ε 2 ) . Therefore

(20.9) | E | c ε 1 / 2

for a certain constant c′ > 0. □


Corresponding author: Michael Christ, Department of Mathematics, University of California, Berkeley, CA 94720-3840, USA, E-mail: 

To Robert Fefferman, in gratitude.


  1. Research ethics: Not applicable.

  2. Informed consent: Not applicable.

  3. Author contributions: The author has accepted responsibility for the entire contentof this manuscript and approved its submission.

  4. Use of Large Language Models, AI and Machine Learning Tools: None declared.

  5. Conflict of interest: None.

  6. Research funding: The author’s research was supported in part by NSF grants DMS-13363724 and DMS-1901413.

  7. Data availability: Not applicable.

References

[1] L. Hörmander, “Oscillatory integrals and multipliers on FLp,” Ark. Mat., vol. 11, nos. 1–2, pp. 1–11, 1973.10.1007/BF02388505Search in Google Scholar

[2] D. H. Phong, E. M. Stein, and J. Sturm, “Multilinear level set operators, oscillatory integral operators, and Newton polyhedra,” Math. Ann., vol. 319, no. 3, pp. 573–596, 2001, https://doi.org/10.1007/BF02796120.Search in Google Scholar

[3] M. Gilula, P. T. Gressman, and L. Xiao, “Higher decay inequalities for multilinear oscillatory integrals,” Math. Res. Lett., vol. 25, no. 3, pp. 819–842, 2018.10.4310/MRL.2018.v25.n3.a5Search in Google Scholar

[4] E. M. Stein, Harmonic Analysis: Real-Variable Methods, Orthogonality, and Oscillatory Integrals. With the Assistance of Timothy S. Murphy, Princeton Mathematical Series, 43. Monographs in Harmonic Analysis, III, Princeton, NJ, Princeton University Press, 1993.10.1515/9781400883929Search in Google Scholar

[5] M. Christ, X. Li, T. Tao, and C. Thiele, “On multilinear oscillatory integrals, nonsingular and singular,” Duke Math. J., vol. 130, no. 2, pp. 321–351, 2005.10.1215/00127094-8229909Search in Google Scholar

[6] A. Carbery and J. Wright, “What is van der Corput’s lemma in higher dimensions?” in Proceedings of the 6th International Conference on Harmonic Analysis and Partial Differential Equations (El Escorial, 2000), Publ. Mat., vol. Extra, 2002, pp. 13–26.10.5565/PUBLMAT_Esco02_01Search in Google Scholar

[7] A. Carbery, M. Christ, and J. Wright, “Multidimensional van der Corput and sublevel set estimates,” J. Am. Math. Soc., vol. 12, no. 4, pp. 981–1015, 1999.10.1090/S0894-0347-99-00309-4Search in Google Scholar

[8] Z. Zhou, “On multilinear oscillatory integrals and associated maximal functions,” Ph.D. dissertation, University of California, Berkeley, 2024.Search in Google Scholar

[9] J. Bourgain, “A nonlinear version of Roth’s theorem for sets of positive density in the real line,” J. Anal. Math., vol. 50, pp. 169–181, 1988, https://doi.org/10.1007/BF02796120.Search in Google Scholar

[10] S. Peluse, “Bounds for sets with no polynomial progressions,” Forum Math. Pi, vol. 8, 2020, Art. no. e16, https://doi.org/10.1017/fmp.2020.11.Search in Google Scholar

[11] S. Peluse and S. Prendiville, “Quantitative bounds in the non-linear Roth theorem,” Invent. Math., vol. 238, no. 3, pp. 865–903, 2024.10.1007/s00222-024-01293-xSearch in Google Scholar

[12] S. Peluse and S. Prendiville, “A polylogarithmic bound in the nonlinear Roth theorem,” Int. Math. Res. Not. IMRN, no. 8, pp. 5658–5684, 2022.10.1093/imrn/rnaa261Search in Google Scholar

[13] B. Krause, M. Mirek, S. Peluse, and J. Wright, “Polynomial progressions in topological fields,” Forum Math. Sigma, vol. 12, 2024, Art. no. e106, https://doi.org/10.1017/fms.2024.104.Search in Google Scholar

[14] T. Tao, Cut Norms and Degree Lowering, 2020. Available at: https://terrytao.wordpress.com/2020/03/08/cut-norms-and-degree-lowering/.Search in Google Scholar

[15] M. Christ, “On trilinear oscillatory integral inequalities and related topics,” 2020, arXiv:2007.12753.Search in Google Scholar

[16] M. Christ, P. Durcik, and J. Roos, “Trilinear smoothing inequalities and a variant of the triangular Hilbert transform,” Adv. Math., vol. 390, 2021, Art. no. 107863, https://doi.org/10.1016/j.aim.2021.107863.Search in Google Scholar

[17] M. Christ, P. Durcik, V. Kovač, and J. Roos, “Pointwise convergence of certain continuous-time double ergodic averages,” Ergod. Theory Dyn. Syst., vol. 42, no. 7, pp. 2270–2280, 2022.10.1017/etds.2021.45Search in Google Scholar

[18] M. Christ and Z. Zhou, “A class of singular bilinear maximal functions,” J. Funct. Anal., vol. 287, no. 8, 2024, Art. no. 110572.10.1016/j.jfa.2024.110572Search in Google Scholar

[19] M. Christ, “On implicitly oscillatory trilinear integrals with data of limited regularity,” Math. Res. Lett., https://intlpress.com/journals/journalList?p=4&id=1804416324302073857.Search in Google Scholar

[20] J. L. Joly, G. Métivier, and J. Rauch, “Trilinear compensated compactness and nonlinear geometric optics,” Ann. Math., vol. 142, no. 1, pp. 121–169, 1995.10.2307/2118612Search in Google Scholar

[21] M. Christ, “Multilinear oscillatory integral inequalities: best of the best,” in Madison Lectures on Fourier Analysis, 2019. Available at: https://www.math.wisc.edu/seeger2019/christ.Search in Google Scholar

[22] P. Gressman and E. Urheim, “Multilinear oscillatory integral operators and geometric stability,” J. Geom. Anal., vol. 31, no. 9, pp. 8710–8734, 2021.10.1007/s12220-020-00383-5Search in Google Scholar

[23] P. T. Gressman and L. Xiao, “Maximal decay inequalities for trilinear oscillatory integrals of convolution type,” J. Funct. Anal., vol. 271, no. 12, pp. 3695–3726, 2016.10.1016/j.jfa.2016.09.003Search in Google Scholar

[24] W. Blaschke and G. Bol, Geometrie der Gewebe, Berlin, Springer-Verlag, 1938.Search in Google Scholar

[25] L. C. Evans, preprint.Search in Google Scholar

[26] M. Christ, “Bounds for multilinear sublevel sets via Szemerédi’s theorem,” 2011, arXiv:1107.2350.Search in Google Scholar

[27] A. Zygmund, Trigonometric Series, vols. I–II, 3rd ed. With a foreword by Robert A. Fefferman, Cambridge, Cambridge Mathematical Library, Cambridge University Press, 2002.Search in Google Scholar

[28] M. Christ, “Near-extremizers of Young’s inequality for Euclidean groups,” Rev. Mat. Iberoam., vol. 35, no. 7, pp. 1925–1972, 2019.10.4171/rmi/1055Search in Google Scholar

[29] J. Bourgain, “A remark on the maximal function associated to an analytic vector field,” in Analysis at Urbana, Vol. I (Urbana, IL, 1986–1987), London Math. Soc. Lecture Note Ser., vol. 137, Cambridge, Cambridge Univ. Press, 1989, pp. 111–132.10.1017/CBO9780511662294.006Search in Google Scholar

[30] E. M. Stein and B. Street, “Multi-parameter singular Radon transforms III: real analytic surfaces,” Adv. Math., vol. 229, no. 4, pp. 2210–2238, 2012.10.1016/j.aim.2011.11.016Search in Google Scholar

[31] S. Łojasiewicz, “Sur le probléme de la division,” Stud. Math., vol. 18, no. 1, pp. 87–136, 1959.10.4064/sm-18-1-87-136Search in Google Scholar

[32] M. Christ, “A three term sublevel set inequality,” 2022, arXiv:2204.04346.Search in Google Scholar

[33] M. Christ, “On implicitly oscillatory quadrilinear integrals,” 2022, arXiv:2204.03780.Search in Google Scholar

Received: 2024-08-13
Accepted: 2025-04-04
Published Online: 2025-05-29

© 2025 the author(s), published by De Gruyter, Berlin/Boston

This work is licensed under the Creative Commons Attribution 4.0 International License.

Downloaded on 12.9.2025 from https://www.degruyterbrill.com/document/doi/10.1515/ans-2023-0181/html
Scroll to top button