Two principles of decoupling
Abstract.
We put forward a conical principle and a degeneracy locating principle of decoupling. The former generalises the Pramanik-Seeger argument used in the proof of decoupling for the light cone. The latter locates the degenerate part of a manifold and effectively reduces the decoupling problem to two extremes: non-degenerate case and totally degenerate case. Both principles aim to provide a new algebraic approach of reducing decoupling for new manifolds to decoupling for known manifolds.
1. Introduction
Fourier decoupling was first introduced by Wolff in [Wol00] as a tool to prove local smoothing estimates for large . In [Wol00], the decoupling was formulated in the setting of the truncated light cone in . Later, observed by Pramanik and Seeger [PS07], the decoupling inequalities for light cones in can be reduced to that of circles in . The technique is now known as Pramanik-Seeger iteration. In simple terms, a light cone can be approximated by cylinders at intermediate scales. By projection and induction on scales, decoupling for a light cone in is then reduced to decoupling for the unit sphere in . Therefore, together with the seminal work of Bourgain and Demeter [BD15] on the sharp decoupling inequalities for the unit sphere (equivalent to that of a compact piece of a paraboloid), satisfactory decoupling results for light cones are obtained.
Based upon decoupling for spheres and light cones, decoupling inequalities for general manifolds have also been largely studied, and we consider two main types of such manifolds. The first type of manifolds carry certain non-degenerate conditions. See Table 1 below for some important examples. See also the following list of decoupling for non-degenerate manifolds in : [BD16][Oh18][DGS19][GOZZK23][GZ19]. Their nondengenacy conditions are slightly harder to state, and we encourage the reader to their papers for details. In general, these results rely on the corresponding multilinear decoupling inequalities derived from the geometry of the manifolds.
Results | non-degeneracy | critial exponents | ||
---|---|---|---|---|
Bourgain-Demeter | ||||
[BD15] | ||||
Bourgain-Demeter | ||||
[BD17] | ||||
Bourgain-Demeter- | ||||
Guth [BDG16] | ||||
Guth-Maldague- | ||||
Oh [GMO24] |
The second type of manifolds shown in Table 2 below are more general as the non-degenerate conditions may not be satisfied. The major approach to these results is to deal with the decoupling for pieces of the manifold on which the non-degenerate factor is small. For instance, for the case of hypersurfaces, the non-degenerate factor is the Gaussian curvature. When compared to Table 1, very limited results are known in higher dimensions or higher co-dimensions .
Results | descriptions | |
Biswas-Gilula-Li- | ||
Schwend-Xi [BGL+20]; | 2 | analytic/smooth planar curves |
Demeter [Dem20]; | ||
Y. [Yan21] | ||
Bourgain-Demeter | some surfaces of revolution: | |
-Kemp [BDK20] | ||
Kemp [Kem21] | 3 | surfaces lacking planar points |
Kemp [Kem22] | 3 | Gaussian curvature vanishing identically |
L.-Y. [LY22] | 3 | mixed-homogeneous polynomials |
L.-Y. [LY23]; Guth- | 3 | smooth surfaces |
Maldague-Oh [GMO24] | ||
Gao-Li-Zhao-Zheng | , or . | |
[GLZZ22] |
In this paper, we first generalise the cone-to-sphere reduction to a radial principle that applies to more general manifolds that exhibits some conical structure. Then, we form the degeneracy locating principle by abstracting the idea of reducing the manifolds with possible degeneracy to the non-degenerate counterpart. These results provide an algebraic approach of reducing decoupling for more complicated manifolds to simpler manifolds. In particular, we provide immediate corollaries of these principles, namely, decoupling inequalities for new manifolds.
1.1. Decoupling inequalities
We first formulate decoupling inequalities in a more general setting.
Definition 1.1.
Given a compact subset and a finite collection of boundedly overlapping111To deal with technicalities, we will impose a slightly stronger overlapping condition. See Definition 1.7. parallelograms . For Lebesgue exponents and , we define the -decoupling constant to be the smallest constant such that
(1.1) |
for all smooth test function Fourier supported on .
Given a subset , we say can be -decoupled into the parallelograms at the cost of , if and .
When are given, the exponent is most commonly taken to be . Hence, we also say can be -decoupled if can be -decoupled. To simplify the notation further, since and the exponents are usually clear from the context, we will often abbreviate , and simply say can be decoupled into parallelograms .
By Hölder’s inequality, decoupling implies decoupling if . Thus, together with Plancherel’s identity, we always have decoupling for . Also, for a fixed and , readers shall keep in mind that decoupling is generally stronger than decoupling. This follows from interpolation whenever applicable, or can be usually observed from the proof of decoupling.
1.1.1. Decoupling for manifolds in
For , we often consider the case when is the -normal neighbourhood of a (compact piece of a) manifold :
To formulate decoupling for manifolds, we first introduce the following general definition which is independent of the manifold .
Definition 1.2 (-flat subsets).
Let . We say a subset is -flat in dimension , if it is contained inside the -normal neighbourhood of some -plane.
With this, we return to the formulation of decoupling for manifolds . Fix . We usually would like to decouple into smaller pieces , such that the convex hull of is equivalent to a parallelogram that is -flat in dimension .
We remark that if is a hypersurface, then the convex hull of will be essentially the same as , and we only need to consider the case . On the other hand, if is a curve, then the convex hull of will be nearly equivalent to the corresponding piece of an anisotropic neighbourhood222See, for instance, [GLYZK21] or Chapter 11 of [Dem20] for definitions of anisotropic neighbourhoods of a curve. of . We might need to consider all .
By a partition of unity, we may assume that is given by the graph of an -valued function on a compact set with bounded derivatives, the -vertical neighbourhood of
(1.2) |
is equivalent to the -normal neighbourhood of for decoupling considerations (see Section 1.7 for the precise definition of equivalent subsets). In this case, we may consider the projection of the -flat parallelograms onto the coordinate space.
Definition 1.3.
Let . We say a parallelogram is -flat in dimension , if we have
(1.3) |
for some affine function given by where has orthonormal rows, and .333See Appendix 4.3 for more properties of -flatness.
To simplify notation, when is given by the graph of over , we say can be graphically decoupled into -flat parallelograms , if can be decoupled into parallelograms , such that each is equivalent to the convex hull of .
1.2. A radial principle of decoupling
To push the ideas of Pramanik-Seeger iteration [PS07] further, we formulate and prove a radial principle of decoupling. This is a generalisation of the arguments used in Section 8 of [BD15] (see also Appendix B of [GGG+22]). Roughly speaking, under certain non-degenerate assumptions, the decoupling for the manifold parameterised by
can be reduced to the decoupling for
The precise statement of the radial principle of decoupling is in Section 2. We give some examples.
-
(1)
Consider the surface given by the truncated cone . After a finite partition and rotation, we may re-parameterise each part of the cone by
where and . By the radial principle, it suffices to study decoupling for the following cylinder
which is precisely the reduction given in [BD15] using Pramanik-Seeger iteration.
-
(2)
Consider the radial surface generated by a smooth curve , that is, under an additional assumption that , where can be thought of as perturbations of the light cone. The case has been studied in [BDK20]. Similarly, each part of the surface can be reparametrised by , where . By the implicit function theorem, we further reparametrise the surface by
Now, the radial principle of decoupling reduces the original decoupling problem to the decoupling for the following surface:
Although for these two examples, we can use elementary arguments to study decoupling for the reduced manifolds, we will instead put them under a more general framework introduced right below.
Note that a common feature of the reduced manifolds is that the degenerate set is separated in the and variables. It facilitates the needs of the following degeneracy locating principle of decoupling.
1.3. A degeneracy locating principle of decoupling
We make abstract the idea of reducing the manifolds with possible degeneracy to totally degenerate and non-degenerate cases. To avoid technicalities, we present a simple version here. See Section 3 for the detailed definitions and formulation.
Let be a family of manifolds given by the graph of a vector valued function for some parallelogram . We say that is a degeneracy determinant if it quantitatively detects the degeneracy of the manifold. More precisely, maps to a smooth function on such that the following holds:
-
•
if , then restricted to is within a neighbourhood of for some totally degenerate manifold , i.e. , for some depending on .
Let us consider the family consisting of graphs of polynomials of degree at most with bounded coefficients. For and polynomial space curves , it is easy to check that and the Wronskian of , namely, , are examples of degeneracy determinants. For and , is a degeneracy determinant as shown in Proposition 4.8. It generalises Theorem 3.2 of [LY23].
Our degeneracy locating principle says that to decouple into parallelograms, it suffices to understand the decoupling of these cases at a cost uniform in :
-
•
decoupling of the sublevel set
where , , into parallelograms.
-
•
decoupling of the totally degenerate manifolds , i.e. , into parallelograms.
-
•
decoupling of the non-degenerate manifolds , i.e. on , into parallelograms.
In the examples given in Table 2, the totally degenerate cases are the same problems of at least one dimensional lower and the non-degenerate case follows from the famous theorem of Bourgain and Demeter [BD15, BD17]. When , the sublevel set decoupling for polynomial family of degree at most and bounded coefficients is a direct corollary of fundamental theorem of algebra. For special cases when and in [BDK20] and [Kem21], the sublevel sets are neighbourhoods of circles and parabolas. In [LY23], exactly the same framework is used, where the sublevel set decoupling was called “generalised 2D uniform decoupling”.
1.4. Two different scales of induction
We remark that the proofs of both principles are based on the method of induction on scales. More precisely, for the radial principle, essentially we are iterating a bootstrap inequality of the form
Each time the step changes as gets larger, and the number of steps is of the order .
For the degeneracy locating principle, essentially we are iterating a bootstrap inequality of the form
where the step is fixed throughout the iteration. The number of steps is of the order .
In both cases, we are able to prove that the cost of such reduction for every .
1.5. Corollaries
As an application of the two principles, we present decoupling inequalities of two new types of manifolds. The proofs below assume the simplified versions above to avoid technicalities. Readers are encouraged to check that a direct modification of the arguments in this subsection applies to the detailed versions of the two principles in Sections 2 and 3 below.
Corollary 1.4 (Decoupling for radial surfaces).
Let . Let be a smooth function with . Let be the hypersurface in given by the graph of the function over
For any , let be a partition of into -flat intervals such that is not -flat. Let be a covering of by parallelograms , each of which is equivalent to the region
for and -squares .
Then can be graphically decoupled into -flat parallelograms at a cost of , for any . Moreover, if is convex, this decoupling can be improved to decoupling.
Proof of Corollary 1.4 using the simplified versions of the two principles.
Following the arguments in Example 2 in Subsection 1.2, the manifold can be broken into many pieces, each of which can be rotated and reparameterised by
(1.4) |
where . By the radial principle of decoupling, it reduces to the decoupling of the manifold given by
(1.5) |
Since is smooth with on , is smooth on . By the smooth approximation argument in Section 2.3 of [LY23], it suffices to prove the uniform decoupling result for the family consisting of manifolds with where is a polynomial of degree at most and with bounded coefficients bounded by . We pick . By the degeneracy locating principle, it suffices to check the following.
-
•
is a degeneracy determinant. This follows from the mean value theorem, since .
-
•
The sublevel set
is already a union of many rectangular strips.
-
•
The totally degenerate manifolds are cylinders because must be an affine function. By cylindrical decoupling, it suffices to decouple
which has non-vanishing Gaussian curvature everywhere. The decoupling inequalities for these cases are given in [BD15].
- •
We have all assumptions for the degeneracy locating principle checked. Therefore, we have obtained the decoupling of the manifold given in (1.5), and hence (1.4), into -flat parallelograms as desired.
It is straightforward to check by the full version of the radial principle, Theorem 2.2, that the family is exactly the decoupled -flat parallelograms.
∎
Corollary 1.5 (Decoupling for graphs of additively separable functions).
Let . For , let such that and be smooth. Let be such that
(1.6) |
For any , there exists a -overlapping444See Definition 1.7; here depends on but not . cover of by -flat parallelograms (in dimension ), such that can be graphically decoupled into at a cost of . Moreover, if is convex, this decoupling can be improved to decoupling.
Proof of Corollary 1.5 using the simplified version of degeneracy locating principle.
Fix . For each , let be the collection of manifolds given by graphs of polynomials of the form (1.6), such that each has degree at most and coefficients bounded by 1, and on for . It suffices to prove the decoupling for at a cost of uniform in the coefficients of the polynomials, by the smooth approximation argument in Section 2.3 of [LY23].
We now induce on and . The base case is exactly given in [BD15, BD17] at a cost of . The other base cases are exactly in [Yan21] and [LY23] respectively. By induction, we may assume the decoupling of at a cost of .
We apply the degeneracy locating principle on with .
First, we consider the easy case when . By the mean value theorem, is a degeneracy determinant. By the fundamental theorem of algebra, the sublevel set
is a union of many intervals. Totally degenerate with are cylinders, which can be solved by cylindrical decoupling and the induction hypothesis on . Non-degenerate are those with , hence also in . This follows from the induction hypothesis.
Second, we consider the remaining case when . By Proposition 4.8, is a degeneracy determinant when . The sublevel set decoupling is exactly Theorem 3.1 of [LY23]. Totally degenerate with are cylinders, which again can be solved by cylindrical decoupling and the induction hypothesis on . The non-degenerate case is similar to the case when . ∎
Corollary 1.6 (Decoupling for analytic curves in ).
Let be an intermediate dimension and . Let be real-analytic, and let be the smooth curve in given by
For each , let be a collection of disjoint intervals that are -flat but not -flat in dimension . Then, can be graphically decoupled into at a cost for any .
We remark that unless the whole curve lies in an affine space of dimension , such partition always exists for . The interested reader may refer to Proposition 3.1 of [Yan21] for a proof in the case ; the higher dimensional cases are similar.
To prove the new version, we observe that if has rank at most , stays in some dimensional affine subspace. In this case, is -flat for all . It is impossible to have such a collection .
Proof of Corollary 1.6 using the simplified version of degeneracy locating principle..
The case is trivial. For each , we define the generalised Wronskian matrix
(1.7) |
Since is real analytic, it follows from elementary linear algebra (see for instance [B.00]) that if and only if the curve lies in an -plane.
Let be the collection of manifolds given by graphs of polynomials of bounded coefficients of degree at most satisfying either of the followings:
-
•
is a straight line;
-
•
the determinant of the matrix , as a polynomial of , has some coefficient 1. Here and throughout this proof, all implicit constants depend on only.
Define . To apply the degeneracy locating principle, we need to check the followings:
-
•
is a degeneracy determinant. Suppose that for some interval . Since is a univariate polynomial of degree at most having some coefficients , by Corollary 4.3,
Hence, . On the other hand, , for some . Thus, restricted to is inside neighbourhood of , where and .
-
•
Decoupling of sublevel sets. The sublevel set is an union of many intervals. No decoupling is needed.
-
•
Decoupling of totally degenerate manifolds . These manifolds are straight lines. No decoupling is needed.
-
•
Decoupling of non-degenerate manifolds . By partition of unity into many pieces, we may assume that has determinant . The decoupling of these manifolds follows from direct applications of cylindrical decoupling and decoupling for moment curves in by Bourgain-Demeter-Guth [BDG16].
With all conditions checked, we conclude that , , can be uniformly decoupled into -flat parallelograms (see (1.2)), , at a cost in dimension . By the smooth approximation argument in Section 2.3 of [LY23], we upgrade the uniform decoupling results for families of polynomial curves to the desired decoupling results for smooth curves. ∎
1.6. Outline of the paper
1.7. Notation
We first introduce some standard notation that will be used in this paper.
-
(1)
We use the standard notation , or to mean there is a constant such that . In many cases, such as results related to polynomials, the constant will be taken to be absolute or depend on the polynomial degree only, and so we may simply write or . Similarly, we define and .
-
(2)
Given a manifold , we define the -normal neighbourhood to be
where is the tangent space of at .
-
(3)
Given a continuous function , we define the -vertical neighbourhood to be
1.7.1. Equivalence of geometric objects
Decoupling inequalities are formulated using parallelograms in . To deal with technicalities, it is crucial that we make clear some fundamental geometric concepts.
-
(1)
A parallelogram is defined to be a set of the form
where is a basis of unit vectors in , and . If is in addition orthogonal, we say is a rectangle. We say two parallelograms are disjoint if their interiors are disjoint.
-
(2)
Unless otherwise specified, for a parallelogram and , we denote by the concentric dilation of by a factor of .
-
(3)
Given a subset , a parallelogram and , we say are -equivalent if . If the constant is unimportant, we simply say are equivalent.
-
(4)
By the John ellipsoid theorem [Joh14], every parallelogram in is -equivalent to a rectangle in , for some dimensional constant . For this reason and in view of the enlarged overlap introduced right below, we do not distinguish rectangles and parallelograms.
1.7.2. Enlarged overlap
Let be a family of parallelograms. In all decoupling inequalities in literature, we hope the parallelograms in do not overlap too much. Since we often deal with constant enlargement of parallelograms, we will need a slightly stronger version of overlap condition.
Definition 1.7.
Given a family of parallelograms in . Given a function , we say the family is -overlapping, if for every constant we have
For all decoupling at some scale in this paper, the function depends on many parameters such as and the family of the manifolds we are decoupling, but most importantly, it is independent of . We simply say is boundedly overlapping if the function depends on unimportant parameters that are clear from the context.
1.8. Acknowledgement
The first author would like to thank Betsy Stovall for preliminary discussions on an early stage of the work. The first author is supported by AMS-Simons Travel Grants. The second author is supported by the Croucher Fellowships for Postdoctoral Research. He would like to thank Terry Tao for numerous suggestions on both the content and the presentation of this article.
2. The radial principle
In this section, we are going to introduce and prove the radial principle of decoupling.
2.1. Statement of theorem
We will need some notation to make the statement precise. First we impose a standard regularity assumption on the partitions.
Definition 2.1 (Dyadic mesh).
For each , let be a cover of by parallelograms of variable directions and dimensions. Given a function , we say the collection is a dyadic mesh with -overlap, if the followings hold:
-
(1)
Each is -overlapping as in Definition 1.7.
-
(2)
For dyadic , and each subset , there exists such that is contained in the union of at most subsets where .
-
(3)
for some dimensional constant .
The third condition is very mild. For instance, it is trivially true when each has all dimensions .
Let . Given a function defined in and map into . Given a function defined in and map into .
The canonical example of is given by . A slightly different interesting example is . Both motivate the terminology “radial principle”. However, for future use, we will actually study much more general functions that may not have radial structure; see the beginning of Theorem 2.2 right below.
Given and in the parameter space, we define
(2.1) | ||||
(2.2) |
By Proposition 4.7, if is -flat and is -flat, all in the first sense (4.1), then there exist parallelograms , which contain and are equivalent to , , respectively.
Theorem 2.2 (Radial principle of decoupling).
Suppose that either
-
(1)
is affine, or
-
(2)
is away from .
Assume there are -overlapping dyadic meshes , , covering , respectively, and obeying the following conditions:
-
(1)
Each is -flat in the sense of (4.1), namely, there exist affine functions , such that
(2.3) -
(2)
For each , is -flat in the sense of (4.3), namely,
(2.4) -
(3)
Let be the subset of given by
Thus the collection is a -overlapping cover of by parallelograms, where is independent of .
Fix . Assume for each , the following decoupling inequality holds: for test functions each Fourier supported in , we have
(2.5) for some constant depending possibly on other parameters but independent of the test functions . Here and throughout this section, we abbreviate .
Consider the subset given by
(2.6) |
Thus the collection is a -overlapping cover of by parallelograms, where is independent of . Moreover, for each , the following decoupling inequality holds: for test functions Fourier supported in , we have
(2.7) |
where depends at most on
Moreover, all implicit constants in the big-O notation above depend at most on .
For example, in the case of the standard truncated light cone in studied in Section 8 of [BD15], we have the following choices:
Thus we have the following chain of reduction of decoupling:
which can be dealt with using Bourgain-Demeter decoupling Theorem (Theorem 1 of [BD15]).
The rest of this section is devoted to the proof of Theorem 2.2.
2.2. Preliminary reductions
Denote by the smallest constant such that for test functions Fourier supported in , we have
(2.8) |
Our goal is to show that
(2.9) |
Let . We first partition into cubes of length . Since we allow losses, we may just decouple each . By translation in , we may assume without loss of generality that the centre of is the origin. By dividing the last two coordinates by , we may assume without loss of generality that .
We record that
(2.10) |
By abuse of notation we may also regard as the best decoupling constant adapted to the smaller cubes.
2.3. Intermediate scale decoupling
Let . Since , applying the definition of gives
(2.11) |
where , and where the collection of subsets that lie within , and similarly for . By Part (2) in the definition of a dyadic mesh, it then suffices to decouple for each fixed pair of intermediate scale subsets .
2.4. Approximation
This subsection will be the key to this proof. The main idea as follows. First, we argue by affine invariance that we can let be centred at , and that . Next, also by affine invariance, we can let . Then by Taylor expansion, we can write
where and are -flat over and , respectively, and their gradients are at . Thus the last coordinate can be further simplified to , since . But then reversing the affine invariance, the decoupling of is equivalent to . More precisely, we now need to decouple the manifold parametrised by
If is affine, then the decoupling problem reduces to that for
If , the decoupling problem reduces that for
We now provide the details. Let be centred at . Denote by . By a translation, the manifold can be written as
(2.12) |
where and is centred at .
2.4.1. Affine invariance in
Write . By a linear transformation, we may transform the decoupling problem to the manifold
(2.13) |
Note that . Also, note that by this reduction.
2.4.2. Affine invariance in
Write , where . Using an affine transformation, we may transform the decoupling problem to the manifold
(2.14) |
2.4.3. Approximation
We now denote . We first claim that . Indeed, for any , we compute
since has all dimensions at least . This means that , and in particular, .
It remains to estimate . We first apply an affine rescaling , and denote . Also, denote , . Since , we have . Also, by (2.4) and affine invariance (see Proposition 4.4), we have that
(2.16) |
Then by the elementary Lemma 2.3 below, we have
This finishes the approximation assuming the lemma.
Lemma 2.3.
Let be a parallelogram centred at . Let be a function with . For all , denote by the unit vector in the same direction as . Then we have
Proof of lemma.
Given . Then
∎
Therefore, we have reduced the decoupling of to that for the manifold
(2.17) |
If either or affine, then we can reverse the affine invariance in the variable, and so we have reduced to the decoupling for .
2.5. Applying decoupling assumption
Now we can apply (2.5) with in place of to get that
(2.18) |
By Part (3) in the definition of a dyadic mesh, plus dyadic decomposition and the triangle inequality (see Proposition 4.9), we may assume that is essentially a constant for each at a cost of . Similarly, we may assume is essentially a constant for each .
Combined with (2.11), we thus have for some that
(2.19) |
Recall . This implies the following bootstrap inequality:
(2.20) |
Iterating this bootstrap inequality gives us .
3. The degeneracy locating principle
In this section, we introduce and prove the degeneracy locating principle of decoupling. We need a little notation to make this principle more general.
3.1. The setup
Let . Let be the collection of manifolds in consisting of given by the graph of a vector valued 555Typically, we consider smooth manifolds. The minimal requirement here is that the norms of the functions are bounded uniformly. function for some parallelogram .
Throughout the rest of of this section, we fix some integer , and all flatness conditions are stated in dimension (see Definitions 1.2 and 1.3).
Definition 3.1 (Affine equivalence).
Define an equivalence relation on as follows. if there exist affine transformations with such that . In other words, if there exist with and such that for all .
Note that by the formulation of decoupling in Definition 1.1, it is invariant under affine transformations. We summarise this into the following lemma.
Lemma 3.2.
Suppose that are equivalent. For with and , denote by the -valued function
For and fixed , the following two statements are equivalent:
-
(1)
can be decoupled into parallelograms at cost .
-
(2)
can be decoupled into parallelograms at cost .
Either statement above implies:
can be decoupled into parallelograms at cost .
Furthermore, if for all , then all the above are equivalent to
can be decoupled into parallelograms at cost .
Let be the collection of all parallelograms contained in .
Definition 3.3 (Rescaling invariance).
Let be a subfamily of . A collection of parallelograms in is said to be rescaling invariant with respect to if the following holds. For every , and every -flat parallelogram with , the manifold , as an element of , is equivalent to for some and with .
Moreover, a collection of parallelograms in is said to be non-rotational rescaling invariant with respect to if we further assume that the transformation in the equivalence definition does not involve rotation, i.e. is a diagonal matrix.
In simple words, every flat piece of a manifold in over can be rescaled to become another manifold in . We illustrate Lemma 3.2 and Definition 3.3 using the following example. Consider the moment curve given by , in . Let .
-
•
The interval is -flat in dimension . The manifold can be rescaled to become
which is where so that . A similar argument shows that other -flat intervals behave similarly. Therefore, the collection of all closed intervals is rescaling invariance with respect to the singleton family . In this case, Lemma 3.2 is exactly the key rescaling property of the moment curve used in [BDG16].
-
•
The interval is -flat in dimension . The manifold can be rescaled to become
which is where so that , and . The idea here is that we are ignoring the torsion given by the third coordinate . We can still get decoupling in this case. This can be seen by projection principle and decoupling for the parabola. In this setting, the collection of all closed intervals is rescaling invariance with respect to the family containing all manifolds such that , and is of the form , where is an arbitrary continuous function.
Definition 3.4 (Degeneracy determinant).
Let be a subfamily of and be rescaling invariant with respect to . We say an operator on that sends to a smooth function on is a degeneracy determinant if the following holds.
-
(1)
For every , is Lipschitz, namely,
where depends on only.
-
(2)
There are constants and , both depending on only, such that for any and , if we have , then there exists such that , , and
(3.1)
3.2. Statement of theorem
Theorem 3.5.
Let be a degeneracy determinant defined with respect to a pair . Let , , , and suppose we have the following assumptions, where all implicit constants and overlap bounds below depend only on .
-
(1)
(Sublevel set decoupling) For each and each , the set
can be decoupled into parallelograms uniformly at a cost of , on each of which .
-
(2)
(Degenerate decoupling) If is such that , then for each , can be graphically -decoupled into -flat parallelograms at a cost of ), and each is -overlapping.
-
(3)
(Nondegnerate decoupling) If such that , then for each , can be graphically decoupled into -flat parallelograms at a cost of , and each is -overlapping.
Then for each , can be graphically decoupled into -flat parallelograms , with the implicit constants and overlap bounds depending only on .
Note that only the second assumption requires decoupling into the rescaling invariant subcollection .
It is straight forward to check that the following applications of the degeneracy locating principle in Corollaries 1.4 to 1.6 satisfy the full assumptions in Theorem 3.5.
Corollary | Manifolds | Parallelograms in | |
---|---|---|---|
1.4 | , | , | |
polynomial, | axis-parallel square | ||
1.5 | 666For simplicity, we only write down the case for the family , . The other families are straightforward generalisations of this case. | , | |
polynomial, | axis-parallel square | ||
1.6 | polynomials, | all intervals | |
or is straight line | i.e. |
Regarding overlapping conditions, it is clear that the families in Corollaries 1.4 and 1.6 above are non-rotational invariant with respect to the corresponding families . Therefore, these manifolds can be decoupled into a collection of rectangles having non-overlapping interiors. By a similar construction to in Proposition 3.1 of [Yan21], -flat rectangles are essentially those in the statements of the corollaries.
The rest of this section is devoted to the proof of Theorem 3.5. It features an induction on scales argument similar to that of Section 5 of [LY23]. The only essential difference here is that we choose a different scale (the scale in [LY23]), which allows us to improve the overlapping bound of the parallelograms from to . This improvement is nontrivial, in view of Appendix A of [GMO24].
For the rest of this section, we suppress the subscripts unless otherwise specified.
3.3. The base case
We may assume , otherwise the decoupling inequality is trivial.
Let . Our induction hypothesis is that the conclusion holds at all coarser scales , where is as in Definition 3.4.
3.4. Degenerate and nondegenerate parts
3.5. The degenerate subset
Now we deal with the degenerate subset, with the hope of reducing to the induction hypothesis.
3.5.1. Sublevel set decoupling
We apply the sublevel set decoupling assumption (1) at scale to decouple into parallelograms on each of which . It remains to further decouple each into smaller -flat parallelograms.
3.5.2. Approximation and Degenerate decoupling
By the definition of degeneracy determinant , since , there exists such that and . Apply the degenerate decoupling assumption (2) at scale to the function to graphically decouple into -flat parallelograms . Thus is also -flat.
3.5.3. Rescaling
By the definition of rescaling invariance of , we see that is equivalent to for some . Hence, by Lemma 3.2, the graphical decoupling of into -flat parallelograms is implied by the graphical decoupling of into -flat parallelograms. Thus we have reduced to decoupling at a larger scale, for which the induction hypothesis can be applied.
3.6. Decoupling inequality
We briefly describe the proof of the required decoupling inequality, and point out the differences from Section 5.3 of [LY23].
Denote by the cost of graphical decoupling of into -flat parallelograms. Our goal is to show that .
The first difference is that we now need to combine decoupling inequalities obtained at each step, which is done by a simple dyadic decomposition argument with a logarithmic loss at each step (see Proposition 4.9). This is acceptable since we already lose at each step of the induction. The second difference is the choice of instead of in [LY23] (where we used in place of ). Carrying out the same proof and combining the decoupling inequalities obtained from the nondegenerate and degenerate parts in Sections 3.4 and 3.5, respectively, we obtain the following bootstrap inequality:
(3.2) |
(Here we have omitted the unimportant implicit constants.)
3.6.1. Iteration
Iterating inequality (3.2) for times, we obtain
We will stop this iteration once we reach , that is, when
(3.3) |
In this case we have , and
which finishes the proof of the decoupling inequality.
3.7. Analysis of overlap
There are two main sources of overlaps between parallelograms, namely, overlaps between parallelograms created within one iteration, and overlaps between different iterations.
The former kind of overlap is guaranteed by the overlap condition in the assumption of Theorem 3.5. For the second kind of overlap, now we can see that the choice of leads to , and so we may just roughly bound the overlapping number by .
We remark that this is a slight improvement of the corresponding argument in [LY23], where we were only able to obtain overlap.
Moreover, if the rescaling in Subsection 3.5.3 does not involve rotation, the final collection will have non-overlapping interior if every decoupling steps generates families of parallelograms with non-overlapping interior.
4. Appendix
4.1. Facts about parallelograms
We also often need to transform many geometric objects into parallelograms, which are much easier to handle.
Proposition 4.1.
The following statements are true.
-
(1)
If is a convex body in , then there exist a parallelogram and a constant such that .
-
(2)
Let be parallelograms in with nonempty intersection. Then there exists some degree constant and some parallelogram such that
Here means the dilation of with respect to its centre by a factor of .
4.2. Facts about polynomials
The following facts about polynomials will be used extensively, and are the key to many of our analysis of polynomials.
Proposition 4.2.
For any -variate real polynomial of degree at most , we have the following relation:
As a result, we also have the following relations:
for any constant .
For a proof of this proposition, the reader may consult [Kel28].
Corollary 4.3.
For any parallelogram and any constant ,
Proof.
Let be an affine transformation that maps to . Apply Proposition 4.2 to to get
which implies the desired relation. ∎
4.3. Flatness of functions
Since we only need the different notions of flatness in Section 2, we just focus on flatness in dimension , namely, for hypersurfaces.
Let be a continuous function. Let be a parallelogram contained in and . We say is -flat
-
(1)
in the first sense, if there exists an affine function such that
(4.1) -
(2)
in the second sense, if is , and
(4.2) -
(3)
in the third sense, if is , and for each we have
(4.3) Equivalently, this means that given any line segment in the direction of and contained , we have
(4.4)
Note that all the above flatness are invariant under affine transformations. Namely, if is an affine transformation, then a set is -flat if and only if is -flat, in all the three senses above. Thus in particular, we have
Proposition 4.4.
If is -flat over in the third sense, then we have .
Fix . Below we denote by the property that is -flat over in the first, second, third sense, respectively.
Proposition 4.5.
We have . Conversely, if is a polynomial of degree and is a parallelogram, then we also have , where is a constant depending only on . However, and are both false, for any fixed constant depending on only.
Proof.
is trivial. For the implication , we let be an affine bijection. Then by the assumption,
(4.5) |
Using Proposition 4.2, we see all coefficients of are . Thus, for any direction , we have
Rescaling back, this gives .
Lastly, using the simple example when can be arbitrarily large, we observe that and are both false, for any fixed constant depending on only. ∎
4.4. Flatness of parametrised surfaces
Definition 4.6.
We say a subset is -flat if there exists a hyperplane such that .
Proposition 4.7.
Proof.
By flatness in sense of (4.1), we may assume are affine functions. For the first subset, we need to find a hyperplane such that
Write for a matrix and a vector . Defining , we compute directly that
which is an affine function in . Thus we may just take the hyperplane to be parametrised by
(4.6) |
The argument for the latter set is even easier. ∎
4.5. A more general small Hessian proposition
We provide a slight generalisation of Theorem 3.2 of [LY23].
Proposition 4.8.
Let be a bivariate polynomial of degree at most and coefficients bounded by . Let be a parallelogram, and suppose for some we have on . Then there exists another bivariate polynomial of degree at most with coefficients bounded by , such that , and , where . Moreover, for some .
The case is equivalent to the statement of Theorem 3.2 of [LY23].
Proof.
Let be the exponent given in Theorem 3.2 of [LY23]. We consider two cases according to whether or not.
If , then the shorter dimension of is . Let be the unbounded straight line along one of the longer edges of , and let be the orthogonal projection from onto . Then we simply define .
Then , has all coefficients bounded by , and .
If , then both dimensions of must be at least . Let be an affine bijection. With this, denote
Then has degree at most with coefficients bounded by , and on . Apply Theorem 3.2 of [LY23] to find a rotation such that
where and are polynomials of degree at most and coefficients bounded by . This means , and so
Let
Then is a polynomial of degree at most , , and . It remains to show that all coefficients of are bounded by . But since has coefficients bounded by , it remains to show all coefficients of are bounded above by . But this follows from the assumption that both dimensions of are at least .
Thus, we see that taking suffices. ∎
4.6. Combining decoupling inequalities
We provide a proof of how we can combine different decoupling inequalities at different levels. For simplicity of presentation, we only provide the case when we have strict partitions, as the case of boundedly overlapping covers is similar.
Proposition 4.9.
Let be a partition of a parallelogram by parallelograms , and for each , let be a partition of by parallelograms . Denote , which is a partition of at a finer level. Then for all we have
Proof.
For each , denote
Then there are many such ’s. Given test functions each Fourier supported on , and denote , . Then we have
from which the result follows. ∎
References
- [B.00] Maxime B. The theory of linear dependence. Annals of Mathematics, 2(1/4):81–96, 1900.
- [BD15] J. Bourgain and C. Demeter. The proof of the decoupling conjecture. Ann. of Math. (2), 182(1):351–389, 2015.
- [BD16] J. Bourgain and C. Demeter. Decouplings for surfaces in . J. Funct. Anal., 270(4):1299–1318, 2016.
- [BD17] J. Bourgain and C. Demeter. Decouplings for curves and hypersurfaces with nonzero Gaussian curvature. J. Anal. Math., 133:279–311, 2017.
- [BDG16] J. Bourgain, C. Demeter, and L. Guth. Proof of the main conjecture in Vinogradov’s mean value theorem for degrees higher than three. Ann. of Math. (2), 184(2):633–682, 2016.
- [BDK20] J. Bourgain, C. Demeter, and D. Kemp. Decouplings for real analytic surfaces of revolution. In Geometric aspects of functional analysis. Vol. I, volume 2256 of Lecture Notes in Math., pages 113–125. Springer, Cham, [2020] ©2020.
- [BGL+20] C. Biswas, M. Gilula, L. Li, J. Schwend, and Y. Xi. decoupling in for curves with vanishing curvature. Proc. Amer. Math. Soc., 148(5):1987–1997, 2020.
- [Dem20] C. Demeter. Fourier restriction, decoupling, and applications, volume 184 of Cambridge Studies in Advanced Mathematics. Cambridge University Press, Cambridge, 2020.
- [DGS19] C. Demeter, S. Guo, and F. Shi. Sharp decouplings for three dimensional manifolds in . Rev. Mat. Iberoam., 35(2):423–460, 2019.
- [GGG+22] S. Gan, S. Guo, L. Guth, T. L. J. Harris, D. Maldague, and H. Wang. On restricted projections to planes in . preprint, 2022. arXiv:2207.13844.
- [GLYZK21] S. Guo, Z. K. Li, P.-L. Yung, and P. Zorin-Kranich. A short proof of decoupling for the moment curve. Amer. J. Math., 143(6):1983–1998, 2021.
- [GLZZ22] C. Gao, Z. Li, T. Zhao, and J. Zheng. Decoupling for finite type phases in higher dimensions. 2022. arXiv:2202.11326.
- [GMO24] L. Guth, D. Maldague, and Changkeun Oh. decoupling theorem for surfaces in . 2024. arXiv:2403.18431.
- [GOZZK23] S. Guo, C. Oh, R. Zhang, and P. Zorin-Kranich. Decoupling inequalities for quadratic forms. Duke Math. J., 172(2):387–445, 2023.
- [GZ19] S. Guo and R. Zhang. On integer solutions of Parsell-Vinogradov systems. Invent. Math., 218(1):1–81, 2019.
- [Joh14] F. John. Extremum problems with inequalities as subsidiary conditions [reprint of mr0030135]. In Traces and emergence of nonlinear programming, pages 198–215. Birkhäuser/Springer Basel AG, Basel, 2014.
- [Kel28] O. D. Kellogg. On bounded polynomials in several variables. Math. Z., 27(1):55–64, 1928.
- [Kem21] D. Kemp. Decoupling via scale-based approximations of limited efficacy. 2021. arXiv:2104.04115.
- [Kem22] D. Kemp. Decouplings for surfaces of zero curvature. J. Funct. Anal., 283(12):Paper No. 109663, 23, 2022.
- [LY22] J. Li and T. Yang. Decoupling for mixed-homogeneous polynomials in . Math. Ann., 383(3-4):1319–1351, 2022.
- [LY23] J. Li and T. Yang. Decoupling for smooth surfaces in . Amer. J. Math., 2023. accepted.
- [Oh18] C. Oh. Decouplings for three-dimensional surfaces in . Math. Z., 290(1-2):389–419, 2018.
- [PS07] M. Pramanik and A. Seeger. regularity of averages over curves and bounds for associated maximal operators. Amer. J. Math., 129(1):61–103, 2007.
- [Wol00] T. Wolff. Local smoothing type estimates on for large . Geom. Funct. Anal., 10(5):1237–1288, 2000.
- [Yan21] T. Yang. Uniform -Decoupling in for polynomials. J. Geom. Anal., 31(11):10846–10867, 2021.