This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Two principles of decoupling

Jianhui Li and Tongou Yang Department of Mathematics, Northwestern University
Evanston, IL 60208, United States
jianhui.li@northwestern.edu Department of Mathematics, University of California
Los Angeles, CA 90095, United States
tongouyang@math.ucla.edu
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 LpL^{p} local smoothing estimates for large pp. In [Wol00], the decoupling was formulated in the setting of the truncated light cone in 3\mathbb{R}^{3}. Later, observed by Pramanik and Seeger [PS07], the decoupling inequalities for light cones in 3\mathbb{R}^{3} can be reduced to that of circles in 2\mathbb{R}^{2}. 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 n\mathbb{R}^{n} is then reduced to decoupling for the unit sphere in n1\mathbb{R}^{n-1}. 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 n\mathbb{R}^{n}: [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 nn kk non-degeneracy critial exponents
Bourgain-Demeter nn n1n-1 D2ϕ(x)0D^{2}\phi(x)\succ 0 2(L2(n+1)n1)\ell^{2}(L^{\frac{2(n+1)}{n-1}})
[BD15]
Bourgain-Demeter nn n1n-1 detD2ϕ(x)0\det D^{2}\phi(x)\neq 0 2(n+1)n1(L2(n+1)n1)\ell^{\frac{2(n+1)}{n-1}}(L^{\frac{2(n+1)}{n-1}})
[BD17]
Bourgain-Demeter- nn 11 det(ϕ′′,,ϕ(n))\det\left(\phi^{\prime\prime},\dots,\phi^{(n)}\right) 2(Ln(n+1))\ell^{2}(L^{n(n+1)})
Guth [BDG16] 0\neq 0
Guth-Maldague- 33 22 detD2ϕ(x)0\det D^{2}\phi(x)\neq 0 2(L4)\ell^{2}(L^{4})
Oh [GMO24]
Table 1. Decoupling inequalities for non-degenerate manifolds in n\mathbb{R}^{n} as graphs of ϕ:[1,1]knk\phi:[-1,1]^{k}\to\mathbb{R}^{n-k}.

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 n>3n>3 or higher co-dimensions nk2n-k\geq 2.

Results nn descriptions
Biswas-Gilula-Li-
Schwend-Xi [BGL+20]; 2 analytic/smooth planar curves
Demeter [Dem20];
Y. [Yan21]
Bourgain-Demeter 33 some surfaces of revolution: ϕ(x)=γ(|x|)\phi(x)=\gamma(|x|)
-Kemp [BDK20]
Kemp [Kem21] 3 surfaces lacking planar points
Kemp [Kem22] 3 Gaussian curvature vanishing identically
L.-Y. [LY22] 3 mixed-homogeneous polynomials ϕ\phi
L.-Y. [LY23]; Guth- 3 smooth surfaces
Maldague-Oh [GMO24]
Gao-Li-Zhao-Zheng nn ϕ(x)=iϕi(xi)\phi(x)=\sum_{i}\phi_{i}(x_{i}), ϕi′′1\phi_{i}^{\prime\prime}\sim 1 or ϕi(xi)=xim\phi_{i}(x_{i})=x_{i}^{m}.
[GLZZ22]
Table 2. Decoupling inequalities for curves/surfaces in n\mathbb{R}^{n} as graphs of ϕ:[1,1]n1\phi:[-1,1]^{n-1}\to\mathbb{R}.

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 SnS\subseteq\mathbb{R}^{n} and a finite collection \mathcal{R} of boundedly overlapping111To deal with technicalities, we will impose a slightly stronger overlapping condition. See Definition 1.7. parallelograms RnR\subseteq\mathbb{R}^{n}. For Lebesgue exponents p,q[2,]p,q\in[2,\infty] and α\alpha\in\mathbb{R}, we define the (q(Lp),α)(\ell^{q}(L^{p}),\alpha)-decoupling constant Dec(S,,p,q,α)\mathrm{Dec}(S,\mathcal{R},p,q,\alpha) to be the smallest constant Dec\mathrm{Dec} such that

(1.1) RfRLp(n)Dec(#)121q+αfRLp(n)q(R)\left\lVert\sum_{R}f_{R}\right\rVert_{L^{p}(\mathbb{R}^{n})}\leq\mathrm{Dec}\,\,(\#\mathcal{R})^{\frac{1}{2}-\frac{1}{q}+\alpha}\left\lVert\left\lVert f_{R}\right\rVert_{L^{p}(\mathbb{R}^{n})}\right\rVert_{\ell^{q}(R\in\mathcal{R})}

for all smooth test function fRf_{R} Fourier supported on RSR\cap S.

Given a subset SnS\subseteq\mathbb{R}^{n}, we say SS can be (q(Lp),α)(\ell^{q}(L^{p}),\alpha)-decoupled into the parallelograms RR\in\mathcal{R} at the cost of KK, if SS\subseteq\cup\mathcal{R} and Dec(S,,p,q,α)K\mathrm{Dec}(S,\mathcal{R},p,q,\alpha)\leq K.

When p,qp,q are given, the exponent α\alpha is most commonly taken to be 0. Hence, we also say SS can be q(Lp)\ell^{q}(L^{p})-decoupled if SS can be (q(Lp),0)(\ell^{q}(L^{p}),0)-decoupled. To simplify the notation further, since SS and the exponents p,q,αp,q,\alpha are usually clear from the context, we will often abbreviate Dec()=Dec(S,,p,q,α)\mathrm{Dec}(\mathcal{R})=\mathrm{Dec}(S,\mathcal{R},p,q,\alpha), and simply say SS can be decoupled into parallelograms RR\in\mathcal{R}.

By Hölder’s inequality, q(Lp)\ell^{q}(L^{p}) decoupling implies q(Lp)\ell^{q^{\prime}}(L^{p}) decoupling if 2qq2\leq q^{\prime}\leq q. Thus, together with Plancherel’s identity, we always have q(L2)\ell^{q}(L^{2}) decoupling for q2q\geq 2. Also, for a fixed qq and ppp^{\prime}\geq p, readers shall keep in mind that q(Lp)\ell^{q}(L^{p^{\prime}}) decoupling is generally stronger than q(Lp)\ell^{q}(L^{p}) decoupling. This follows from interpolation whenever applicable, or can be usually observed from the proof of decoupling.

1.1.1. Decoupling for manifolds in n\mathbb{R}^{n}

For 0<δ10<\delta\ll 1, we often consider the case when SS is the δ\delta-normal neighbourhood of a (compact piece of a) manifold MnM\subseteq\mathbb{R}^{n}:

Nδ(M):={ξ+η:ξM,ηTξM,|η|<δ},TξM is the tangent space of M at ξ.N_{\delta}(M):=\{\xi+\eta:\xi\in M,\eta\perp T_{\xi}M,|\eta|<\delta\},\quad T_{\xi}M\text{ is the tangent space of }M\text{ at }\xi.

To formulate decoupling for manifolds, we first introduce the following general definition which is independent of the manifold MM.

Definition 1.2 (δ\delta-flat subsets).

Let 1mn1\leq m\leq n. We say a subset RnR\subseteq\mathbb{R}^{n} is δ\delta-flat in dimension m1m-1, if it is contained inside the δ\delta-normal neighbourhood of some (m1)(m-1)-plane.

With this, we return to the formulation of decoupling for manifolds MM. Fix mnm\leq n. We usually would like to decouple S=Nδ(M)S=N_{\delta}(M) into smaller pieces SS^{\prime}, such that the convex hull of SS^{\prime} is equivalent to a parallelogram RnR\subseteq\mathbb{R}^{n} that is δ\delta-flat in dimension m1m-1.

We remark that if MM is a hypersurface, then the convex hull of SS^{\prime} will be essentially the same as SS^{\prime}, and we only need to consider the case m=nm=n. On the other hand, if MM is a curve, then the convex hull of SS^{\prime} 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 MM. We might need to consider all m[1,n]m\in[1,n].

By a partition of unity, we may assume that MM is given by the graph of an nk\mathbb{R}^{n-k}-valued function ϕ\phi on a compact set Ωk\Omega\subseteq\mathbb{R}^{k} with bounded derivatives, the δ\delta-vertical neighbourhood of MnM\subseteq\mathbb{R}^{n}

(1.2) Nδϕ(Ω):={(ξ,ϕ(ξ)+c):ξΩ,cnk,|c|<δ}N_{\delta}^{\phi}(\Omega):=\{(\xi^{\prime},\phi(\xi^{\prime})+c):\xi^{\prime}\in\Omega,c\in\mathbb{R}^{n-k},|c|<\delta\}

is equivalent to the δ\delta-normal neighbourhood of MM for decoupling considerations (see Section 1.7 for the precise definition of equivalent subsets). In this case, we may consider the projection of the δ\delta-flat parallelograms RR onto the coordinate space.

Definition 1.3.

Let 1mn1\leq m\leq n. We say a parallelogram ωΩ\omega\subseteq\Omega is (ϕ,δ)(\phi,\delta)-flat in dimension m1m-1, if we have

(1.3) supxω|λ(x,ϕ(x))|δ,\sup_{x\in\omega}|\lambda(x,\phi(x))|\leq\delta,

for some affine function λ:nnm+1\lambda:\mathbb{R}^{n}\to\mathbb{R}^{n-m+1} given by xAx+bx\mapsto Ax+b where A(nm+1)×nA\in\mathbb{R}^{(n-m+1)\times n} has orthonormal rows, and bnm+1b\in\mathbb{R}^{n-m+1}.333See Appendix 4.3 for more properties of (ϕ,δ)(\phi,\delta)-flatness.

To simplify notation, when MM is given by the graph of ϕ\phi over Ω\Omega, we say Ω\Omega can be graphically decoupled into (ϕ,δ)(\phi,\delta)-flat parallelograms ω\omega\in\mathcal{R}, if Nδϕ(Ω)N_{\delta}^{\phi}(\Omega) can be decoupled into parallelograms {Rω:ω}\{R_{\omega}:\omega\in\mathcal{R}\}, such that each RωR_{\omega} is equivalent to the convex hull of Nδϕ(ω)N_{\delta}^{\phi}(\omega).

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

(s,r(s)t,r(s)ψ(t)),r(s)1(s,r(s)t,r(s)\psi(t)),\quad r(s)\sim 1

can be reduced to the decoupling for

(s,t,r(s)+ψ(t)).(s,t,r(s)+\psi(t)).

The precise statement of the radial principle of decoupling is in Section 2. We give some examples.

  1. (1)

    Consider the surface MM given by the truncated cone {(t,|t|),1|t|2}n\{(t,|t|),1\leq|t|\leq 2\}\subseteq\mathbb{R}^{n}. After a finite partition and rotation, we may re-parameterise each part of the cone by

    {(r,rt,rψ(t)):1r2,|t|1/10},\{(r,rt^{\prime},r\psi(t^{\prime})):1\leq r\leq 2,|t^{\prime}|\leq 1/10\},

    where ψ(t)=1|t|2\psi(t^{\prime})=\sqrt{1-|t^{\prime}|^{2}} and |t|1/10|t^{\prime}|\leq 1/10. By the radial principle, it suffices to study decoupling for the following cylinder

    (r,t,r+ψ(t))(r,t^{\prime},r+\psi(t^{\prime}))

    which is precisely the reduction given in [BD15] using Pramanik-Seeger iteration.

  2. (2)

    Consider the radial surface MM generated by a smooth curve γ:\gamma:\mathbb{R}\to\mathbb{R}, that is, M={(t,γ(|t|)):1|t|2}nM=\{(t,\gamma(|t|)):1\leq|t|\leq 2\}\subset\mathbb{R}^{n} under an additional assumption that γ(r)1\gamma^{\prime}(r)\sim 1, where SS can be thought of as perturbations of the light cone. The case n=3n=3 has been studied in [BDK20]. Similarly, each part of the surface can be reparametrised by {(γ(r),rt,rψ(t)),1r2,|t|1/10}\{(\gamma(r),rt^{\prime},r\psi(t^{\prime})),1\leq r\leq 2,|t^{\prime}|\leq 1/10\}, where ψ(t)=1|t|2\psi(t)=\sqrt{1-|t^{\prime}|^{2}}. By the implicit function theorem, we further reparametrise the surface by

    (s,γ1(s)t,γ1(s)ψ(t)),γ1(s)1.(s,\gamma^{-1}(s)t^{\prime},\gamma^{-1}(s)\psi(t^{\prime})),\quad\gamma^{-1}(s)\sim 1.

    Now, the radial principle of decoupling reduces the original decoupling problem to the decoupling for the following surface:

    (s,t,γ1(s)+ψ(t)).(s,t^{\prime},\gamma^{-1}(s)+\psi(t^{\prime})).

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 ss and tt 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 \mathcal{M} be a family of manifolds MϕM_{\phi} given by the graph of a vector valued function ϕ:Rϕnk\phi:R_{\phi}\to\mathbb{R}^{n-k} for some parallelogram Rϕ[1,1]kR_{\phi}\subseteq[-1,1]^{k}. We say that HH is a degeneracy determinant if it quantitatively detects the degeneracy of the manifold. More precisely, HH maps MϕM_{\phi}\in\mathcal{M} to a smooth function on RϕR_{\phi} such that the following holds:

  • if supRϕ|HMϕ|σ1\sup_{R_{\phi}}|HM_{\phi}|\leq\sigma\ll 1, then MϕM_{\phi} restricted to R×nkR\times\mathbb{R}^{n-k} is within a CσβC\sigma^{\beta} neighbourhood of for some totally degenerate manifold MψM_{\psi}\in\mathcal{M}, i.e. HMψ0HM_{\psi}\equiv 0, for some β>0,C1\beta>0,C\geq 1 depending on \mathcal{M}.

Let us consider the family \mathcal{M} consisting of graphs of polynomials ϕ\phi of degree at most dd with bounded coefficients. For k=1k=1 and polynomial space curves (t,ϕ1(t),,ϕn1(t)),tI[1,1](t,\phi_{1}(t),\dots,\phi_{n-1}(t)),t\subseteq I\subseteq[-1,1], it is easy to check that supj=1,,n1|ϕj′′(t)|\sup_{j=1,\dots,n-1}|\phi_{j}^{\prime\prime}(t)| and the Wronskian of ϕ′′\phi^{\prime\prime}, namely, det(ϕ′′,ϕ(3),,ϕ(n))\det\left(\phi^{\prime\prime},\phi^{(3)},\dots,\phi^{(n)}\right), are examples of degeneracy determinants. For k=2k=2 and nk=1n-k=1, HMϕ:=detD2ϕHM_{\phi}:=\det D^{2}\phi 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 MϕM_{\phi}\in\mathcal{M} into parallelograms, it suffices to understand the decoupling of these cases at a cost uniform in \mathcal{M}:

  • decoupling of the sublevel set

    {xRϕ:|HMϕ(x)|σ},\{x\in R_{\phi}:|HM_{\phi}(x)|\leq\sigma\},

    where MϕM_{\phi}\in\mathcal{M}, σ1\sigma\leq 1, into parallelograms.

  • decoupling of the totally degenerate manifolds MϕM_{\phi}\in\mathcal{M}, i.e. HMϕ0HM_{\phi}\equiv 0, into parallelograms.

  • decoupling of the non-degenerate manifolds MϕM_{\phi}\in\mathcal{M}, i.e. HMϕ1HM_{\phi}\sim 1 on RϕR_{\phi}, 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 k=1k=1, the sublevel set decoupling for polynomial family \mathcal{M} of degree at most dd and bounded coefficients is a direct corollary of fundamental theorem of algebra. For special cases when k=2k=2 and n=3n=3 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

D(δ)Cεδε2D(δ1ε).D(\delta)\leq C_{\varepsilon}\delta^{-\varepsilon^{2}}D(\delta^{1-\varepsilon}).

Each time the step δε\delta^{-\varepsilon} changes as δ\delta gets larger, and the number of steps is of the order ε1loglogδ1\varepsilon^{-1}\log\log\delta^{-1}.

For the degeneracy locating principle, essentially we are iterating a bootstrap inequality of the form

D(δ)CεKεD(Kδ),D(\delta)\leq C_{\varepsilon}K^{\varepsilon}D(K\delta),

where the step K=δεK=\delta^{-\varepsilon} is fixed throughout the iteration. The number of steps is of the order ε1\varepsilon^{-1}.

In both cases, we are able to prove that the cost of such reduction D(δ)εδεD(\delta)\lesssim_{\varepsilon}\delta^{-\varepsilon} for every ε(0,1)\varepsilon\in(0,1).

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 2p2(n+1)n12\leq p\leq\frac{2(n+1)}{n-1}. Let γ:[1,2]\gamma:[1,2]\to\mathbb{R} be a smooth function with γ1\gamma^{\prime}\sim 1. Let MM be the hypersurface in n\mathbb{R}^{n} given by the graph of the function ϕ(t)=γ(|t|)\phi(t)=\gamma(|t|) over

Ω={1|t|2,|t1||t|/2},t:=(t1,t).\Omega=\{1\leq|t|\leq 2,|t_{1}|\geq|t|/2\},\quad\quad t:=(t_{1},t^{\prime}).

For any δ>0\delta>0, let δ\mathcal{I}_{\delta} be a partition of [1,2][1,2] into (γ,δ)(\gamma,\delta)-flat intervals II such that 2I2I is not (γ,δ)(\gamma,\delta)-flat. Let δ\mathcal{R}_{\delta} be a covering of Ω\Omega by parallelograms RR, each of which is equivalent to the region

{(r1|t|2,rt):rI,tT}\{(r\sqrt{1-|t^{\prime}|^{2}},rt^{\prime}):r\in I,t^{\prime}\in T\}

for IδI\in\mathcal{I}_{\delta} and δ1/2\delta^{1/2}-squares TT.

Then Ω\Omega can be graphically p(Lp)\ell^{p}(L^{p}) decoupled into (ϕ,δ)(\phi,\delta)-flat parallelograms RδR\in\mathcal{R}_{\delta} at a cost of Oε,γ(δε)O_{\varepsilon,\gamma}(\delta^{-\varepsilon}), for any ε>0\varepsilon>0. Moreover, if ϕ\phi is convex, this p(Lp)\ell^{p}(L^{p}) decoupling can be improved to 2(Lp)\ell^{2}(L^{p}) 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 MM can be broken into O(1)O(1) many pieces, each of which can be rotated and reparameterised by

(1.4) {(s,γ1(s)t,γ1(s)ψ(t)):|t|1/10,sγ([1,2])},\{(s,\gamma^{-1}(s)t^{\prime},\gamma^{-1}(s)\psi(t^{\prime})):|t^{\prime}|\leq 1/10,s\in\gamma([1,2])\},

where ψ(t)=1|t|2\psi(t^{\prime})=\sqrt{1-|t^{\prime}|^{2}}. By the radial principle of decoupling, it reduces to the decoupling of the manifold given by

(1.5) {(s,t,γ1(s)+ψ(t)):|t|1/10,sγ([1,2])}.\{(s,t^{\prime},\gamma^{-1}(s)+\psi(t^{\prime})):|t^{\prime}|\leq 1/10,s\in\gamma([1,2])\}.

Since γ\gamma is smooth with γ1\gamma^{\prime}\sim 1 on [1,2][1,2], γ1\gamma^{-1} is smooth on γ([1,2])\gamma([1,2]). By the smooth approximation argument in Section 2.3 of [LY23], it suffices to prove the uniform decoupling result for the family =(d)\mathcal{M}=\mathcal{M}(d) consisting of manifolds MϕM_{\phi} with ϕ(s,t)=P(s)+ψ(t)\phi(s,t^{\prime})=P(s)+\psi(t^{\prime}) where PP is a polynomial of degree at most dd and with bounded coefficients bounded by 11. We pick HMϕ(s,t)=P′′(s)HM_{\phi}(s,t^{\prime})=P^{\prime\prime}(s). By the degeneracy locating principle, it suffices to check the following.

  • HH is a degeneracy determinant. This follows from the mean value theorem, since ss\in\mathbb{R}.

  • The sublevel set

    {(s,t):|P′′(s)|σ}\{(s,t^{\prime}):|P^{\prime\prime}(s)|\leq\sigma\}

    is already a union of Od(1)O_{d}(1) many rectangular strips.

  • The totally degenerate manifolds MϕM_{\phi}\in\mathcal{M} are cylinders because PP must be an affine function. By cylindrical decoupling, it suffices to decouple

    {(t,ψ(t)):|t|1/10}\{(t^{\prime},\psi(t^{\prime})):|t^{\prime}|\leq 1/10\}

    which has non-vanishing Gaussian curvature everywhere. The decoupling inequalities for these cases are given in [BD15].

  • The non-degenerate manifolds MϕM_{\phi}\in\mathcal{M} has non-vanishing Gaussian curvature everywhere. Their decoupling inequalities are given in [BD15] and [BD17] (for hyperbolic cases).

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 δ\delta-flat parallelograms as desired.

It is straightforward to check by the full version of the radial principle, Theorem 2.2, that the family δ\mathcal{R}_{\delta} is exactly the decoupled δ\delta-flat parallelograms.

Corollary 1.5 (Decoupling for graphs of additively separable functions).

Let 2p2(n+1)n12\leq p\leq\frac{2(n+1)}{n-1}. For j=1,,Jj=1,\dots,J, let nj{1,2}n_{j}\in\{1,2\} such that jnj=n1\sum_{j}n_{j}=n-1 and ϕj:[1,1]nj\phi_{j}:[-1,1]^{n_{j}}\to\mathbb{R} be smooth. Let ϕ:[1,1]n1\phi:[-1,1]^{n-1}\to\mathbb{R} be such that

(1.6) ϕ(x1,,xJ)=j=1Jϕj(xj),xj[1,1]nj.\phi(x_{1},\dots,x_{J})=\sum_{j=1}^{J}\phi_{j}(x_{j}),\quad x_{j}\in[-1,1]^{n_{j}}.

For any ε,δ>0\varepsilon,\delta>0, there exists a BB-overlapping444See Definition 1.7; here BB depends on ϕ,ε\phi,\varepsilon but not δ\delta. cover \mathcal{R} of [1,1]n1[-1,1]^{n-1} by (ϕ,δ)(\phi,\delta)-flat parallelograms RR (in dimension n1n-1), such that [1,1]n1[-1,1]^{n-1} can be graphically p(Lp)\ell^{p}(L^{p}) decoupled into RR\in\mathcal{R} at a cost of Oε,ϕ(δε)O_{\varepsilon,\phi}(\delta^{-\varepsilon}). Moreover, if ϕ\phi is convex, this p(Lp)\ell^{p}(L^{p}) decoupling can be improved to 2(Lp)\ell^{2}(L^{p}) decoupling.

Proof of Corollary 1.5 using the simplified version of degeneracy locating principle.

Fix dd. For each 0lJ0\leq l\leq J, let ln\mathcal{M}_{l}^{n} be the collection of manifolds MϕM_{\phi} given by graphs of polynomials of the form (1.6), such that each ϕj\phi_{j} has degree at most dd and coefficients bounded by 1, and |detD2ϕj|1|\det D^{2}\phi_{j}|\sim 1 on [1,1]nj[-1,1]^{n_{j}} for j=1,,lj=1,\dots,l. It suffices to prove the decoupling for Mϕ0nM_{\phi}\in\mathcal{M}_{0}^{n} at a cost of Oε,d(δε)O_{\varepsilon,d}(\delta^{-\varepsilon}) uniform in the coefficients of the polynomials, by the smooth approximation argument in Section 2.3 of [LY23].

We now induce on nn and ll. The base case Jn\mathcal{M}_{J}^{n} is exactly given in [BD15, BD17] at a cost of Oε,d(δε)O_{\varepsilon,d}(\delta^{-\varepsilon}). The other base cases 01,02\mathcal{M}_{0}^{1},\mathcal{M}_{0}^{2} are exactly in [Yan21] and [LY23] respectively. By induction, we may assume the decoupling of Mϕ1n,0n1,0n2M_{\phi}\in\mathcal{M}_{1}^{n},\mathcal{M}_{0}^{n-1},\mathcal{M}_{0}^{n-2} at a cost of Oε,d(δε)O_{\varepsilon,d}(\delta^{-\varepsilon}).

We apply the degeneracy locating principle on 0n\mathcal{M}_{0}^{n} with H(Mϕ)=detD2ϕ1H(M_{\phi})=\det D^{2}\phi_{1}.

First, we consider the easy case when n1=1n_{1}=1. By the mean value theorem, HH is a degeneracy determinant. By the fundamental theorem of algebra, the sublevel set

{x[1,1]:|ϕ1′′(x)|σ}\{x\in[-1,1]:|\phi_{1}^{\prime\prime}(x)|\leq\sigma\}

is a union of Od(1)O_{d}(1) many intervals. Totally degenerate Mϕ0nM_{\phi}\in\mathcal{M}^{n}_{0} with ϕ1′′0\phi_{1}^{\prime\prime}\equiv 0 are cylinders, which can be solved by cylindrical decoupling and the induction hypothesis on 0n1\mathcal{M}^{n-1}_{0}. Non-degenerate Mϕ0nM_{\phi}\in\mathcal{M}^{n}_{0} are those with |ϕJ′′|1|\phi_{J}^{\prime\prime}|\sim 1, hence also in 1n\mathcal{M}^{n}_{1}. This follows from the induction hypothesis.

Second, we consider the remaining case when n1=2n_{1}=2. By Proposition 4.8, HH is a degeneracy determinant when n1=2n_{1}=2. The sublevel set decoupling is exactly Theorem 3.1 of [LY23]. Totally degenerate Mϕ0nM_{\phi}\in\mathcal{M}^{n}_{0} with detD2ϕ10\det D^{2}\phi_{1}\equiv 0 are cylinders, which again can be solved by cylindrical decoupling and the induction hypothesis on 0n2\mathcal{M}^{n-2}_{0}. The non-degenerate case is similar to the case when nJ=1n_{J}=1. ∎

Corollary 1.6 (Decoupling for analytic curves in n\mathbb{R}^{n}).

Let m[1,n]m\in[1,n] be an intermediate dimension and 2pm(m+1)2\leq p\leq m(m+1). Let ϕ:[1,1]n1\phi:[-1,1]\to\mathbb{R}^{n-1} be real-analytic, and let MϕM_{\phi} be the smooth curve in n\mathbb{R}^{n} given by

{(s,ϕ1(s),,ϕn1(s)):s[1,1]}.\{(s,\phi_{1}(s),\dots,\phi_{n-1}(s)):s\in[-1,1]\}.

For each 0<δϕ10<\delta\ll_{\phi}1, let δ\mathcal{I}_{\delta} be a collection of disjoint intervals II that are (ϕ,δ)(\phi,\delta)-flat but not (ϕ,δ/2)(\phi,\delta/2)-flat in dimension (m1)(m-1). Then, [1,1][-1,1] can be graphically 2(Lp)\ell^{2}(L^{p}) decoupled into IδI\in\mathcal{I}_{\delta} at a cost Oε,ϕ(δε)O_{\varepsilon,\phi}(\delta^{-\varepsilon}) for any ε>0\varepsilon>0.

We remark that unless the whole curve lies in an affine space of dimension m1m-1, such partition δ\mathcal{I}_{\delta} always exists for δϕ1\delta\ll_{\phi}1. The interested reader may refer to Proposition 3.1 of [Yan21] for a proof in the case n=2n=2; the higher dimensional cases are similar.

To prove the new version, we observe that if WnϕW_{n}\phi has rank at most m2m-2, ϕ\phi stays in some (m1)(m-1) dimensional affine subspace. In this case, [1,1][-1,1] is (ϕ,δ)(\phi,\delta)-flat for all δ>0\delta>0. It is impossible to have such a collection δ\mathcal{I}_{\delta}.

Proof of Corollary 1.6 using the simplified version of degeneracy locating principle..

The case m=1m=1 is trivial. For each 2mn2\leq m\leq n, we define the generalised m×nm\times n Wronskian matrix

(1.7) Wmϕ(s):=[ ϕ′′(s)  ϕ(3)(s)  ϕ(m)(s) ].W_{m}\phi(s):=\left[\begin{array}[]{ccc}\rule[2.15277pt]{10.76385pt}{0.5pt}&\phi^{\prime\prime}(s)&\rule[2.15277pt]{10.76385pt}{0.5pt}\\ \rule[2.15277pt]{10.76385pt}{0.5pt}&\phi^{(3)}(s)&\rule[2.15277pt]{10.76385pt}{0.5pt}\\ &\vdots&\\ \rule[2.15277pt]{10.76385pt}{0.5pt}&\phi^{(m)}(s)&\rule[2.15277pt]{10.76385pt}{0.5pt}\end{array}\right].

Since ϕ\phi is real analytic, it follows from elementary linear algebra (see for instance [B.00]) that det(Wmϕ(Wmϕ)T)(s)0\det(W_{m}\phi(W_{m}\phi)^{T})(s)\equiv 0 if and only if the curve ϕ\phi lies in an (m1)(m-1)-plane.

Let \mathcal{M} be the collection of manifolds MϕM_{\phi} given by graphs of polynomials ϕ:[1,1]n1\phi:[-1,1]\to\mathbb{R}^{n-1} of bounded coefficients of degree at most dd satisfying either of the followings:

  • MϕM_{\phi} is a straight line;

  • the determinant of the (m1)×(m1)(m-1)\times(m-1) matrix Wmϕ(s)(Wmϕ(s))TW_{m}\phi(s)(W_{m}\phi(s))^{T}, as a polynomial of ss, has some coefficient \sim 1. Here and throughout this proof, all implicit constants depend on n,dn,d only.

Define HMϕ:=det(Wmϕ(Wmϕ)T)HM_{\phi}:=\det(W_{m}\phi(W_{m}\phi)^{T}). To apply the degeneracy locating principle, we need to check the followings:

  • HH is a degeneracy determinant. Suppose that 0<supI|HMϕ|σ0<\sup_{I}|HM_{\phi}|\leq\sigma for some interval II. Since HMϕHM_{\phi} is a univariate polynomial of degree at most 2dn2dn having some coefficients 1\sim 1, by Corollary 4.3,

    1sup[1,1]|HMϕ||I|2dnsupI|HMϕ||I|2dnσ.1\sim\sup_{[-1,1]}|HM_{\phi}|\lesssim|I|^{-2dn}\sup_{I}|HM_{\phi}|\leq|I|^{-2dn}\sigma.

    Hence, |I|σ1/(2dn)|I|\lesssim\sigma^{1/(2dn)}. On the other hand, supsIϕ(s)ϕ(s0)d|I|\sup_{s\in I}\|\phi(s)-\phi(s_{0})\|_{\infty}\lesssim_{d}|I|, for some s0Is_{0}\in I. Thus, MϕM_{\phi} restricted to I×nkI\times\mathbb{R}^{n-k} is inside O(σ1/(2dn))O(\sigma^{1/(2dn)}) neighbourhood of MψM_{\psi}, where ψ(s)ϕ(s0)\psi(s)\equiv\phi(s_{0}) and HMψ0HM_{\psi}\equiv 0.

  • Decoupling of sublevel sets. The sublevel set {s:|HMϕ(s)|σ}\{s:|HM_{\phi}(s)|\leq\sigma\} is an union of Od(1)O_{d}(1) many intervals. No decoupling is needed.

  • Decoupling of totally degenerate manifolds MϕM_{\phi}\in\mathcal{M}. These manifolds are straight lines. No decoupling is needed.

  • Decoupling of non-degenerate manifolds MϕM_{\phi}\in\mathcal{M}. By partition of unity into On,d(1)O_{n,d}(1) many pieces, we may assume that W(ϕ1′′,,ϕm1′′)W(\phi^{\prime\prime}_{1},\dots,\phi^{\prime\prime}_{m-1}) has determinant 1\sim 1. The decoupling of these manifolds follows from direct applications of cylindrical decoupling and decoupling for moment curves in m\mathbb{R}^{m} by Bourgain-Demeter-Guth [BDG16].

With all conditions checked, we conclude that Nδ(Mϕ)N_{\delta}(M_{\phi}), MϕM_{\phi}\in\mathcal{M}, can be uniformly 2(Lp)\ell^{2}(L^{p}) decoupled into δ\delta-flat parallelograms Nδϕ(I)N^{\phi}_{\delta}(I) (see (1.2)), IδI\in\mathcal{I}_{\delta}, at a cost Oε,(δε)O_{\varepsilon,\mathcal{M}}(\delta^{-\varepsilon}) in dimension m1m-1. 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

In Section 2 we state and prove the radial principle. In Section 3 we state and prove the degeneracy locating principle. Section 4 is the appendix, which discusses some fundamental concepts about parallelograms, polynomials and flatness in decoupling.

1.7. Notation

We first introduce some standard notation that will be used in this paper.

  1. (1)

    We use the standard notation a=OC(b)a=O_{C}(b), or |a|Cb|a|\lesssim_{C}b to mean there is a constant CC such that |a|Cb|a|\leq Cb. In many cases, such as results related to polynomials, the constant CC will be taken to be absolute or depend on the polynomial degree only, and so we may simply write a=O(b)a=O(b) or |a|b|a|\lesssim b. Similarly, we define \gtrsim and \sim.

  2. (2)

    Given a manifold MnM\subseteq\mathbb{R}^{n}, we define the δ\delta-normal neighbourhood Nδ(M)N_{\delta}(M) to be

    Nδ(M):={ξ+η:ξM,ηTξM,|η|<δ},N_{\delta}(M):=\{\xi+\eta:\xi\in M,\eta\perp T_{\xi}M,|\eta|<\delta\},

    where TξMT_{\xi}M is the tangent space of MM at ξ\xi.

  3. (3)

    Given a continuous function ϕ:ΩkRnk\phi:\Omega\subseteq\mathbb{R}^{k}\to R^{n-k}, we define the δ\delta-vertical neighbourhood Nδϕ(Ω)N^{\phi}_{\delta}(\Omega) to be

    Nδϕ(Ω):={(ξ,ϕ(ξ)+c):ξΩ,cnk,|c|<δ}.N_{\delta}^{\phi}(\Omega):=\{(\xi^{\prime},\phi(\xi^{\prime})+c):\xi^{\prime}\in\Omega,c\in\mathbb{R}^{n-k},|c|<\delta\}.

1.7.1. Equivalence of geometric objects

Decoupling inequalities are formulated using parallelograms in n\mathbb{R}^{n}. To deal with technicalities, it is crucial that we make clear some fundamental geometric concepts.

  1. (1)

    A parallelogram RnR\subseteq\mathbb{R}^{n} is defined to be a set of the form

    {x+bn:|xui|li},\{x+b\in\mathbb{R}^{n}:|x\cdot u_{i}|\leq l_{i}\},

    where {ui:1in}\{u_{i}:1\leq i\leq n\} is a basis of unit vectors in n\mathbb{R}^{n}, bnb\in\mathbb{R}^{n} and li0l_{i}\geq 0. If {ui:1in}\{u_{i}:1\leq i\leq n\} is in addition orthogonal, we say RR is a rectangle. We say two parallelograms are disjoint if their interiors are disjoint.

  2. (2)

    Unless otherwise specified, for a parallelogram SnS\subseteq\mathbb{R}^{n} and C>0C>0, we denote by CSCS the concentric dilation of SS by a factor of CC.

  3. (3)

    Given a subset SnS\subseteq\mathbb{R}^{n}, a parallelogram RR and C1C\geq 1, we say S,RS,R are CC-equivalent if C1RSCRC^{-1}R\subseteq S\subseteq CR. If the constant CC is unimportant, we simply say S,RS,R are equivalent.

  4. (4)

    By the John ellipsoid theorem [Joh14], every parallelogram in n\mathbb{R}^{n} is CnC_{n}-equivalent to a rectangle in n\mathbb{R}^{n}, for some dimensional constant CnC_{n}. 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 \mathcal{R} be a family of parallelograms. In all decoupling inequalities in literature, we hope the parallelograms in \mathcal{R} 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 \mathcal{R} of parallelograms in n\mathbb{R}^{n}. Given a function B:[1,)[1,)B:[1,\infty)\to[1,\infty), we say the family is BB-overlapping, if for every constant C1C\geq 1 we have

R1CRB(C).\sum_{R\in\mathcal{R}}1_{CR}\leq B(C).

For all decoupling at some scale δ(0,1)\delta\in(0,1) in this paper, the function BB depends on many parameters such as ε\varepsilon and the family of the manifolds we are decoupling, but most importantly, it is independent of δ\delta. We simply say \mathcal{R} is boundedly overlapping if the function BB 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 δ2\delta\in 2^{-\mathbb{N}}, let δ\mathcal{R}_{\delta} be a cover of [1,1]n[-1,1]^{n} by parallelograms of variable directions and dimensions. Given a function B:[1,)[1,)B:[1,\infty)\to[1,\infty), we say the collection {δ:δ2}\{\mathcal{R}_{\delta}:\delta\in 2^{-\mathbb{N}}\} is a dyadic mesh with BB-overlap, if the followings hold:

  1. (1)

    Each δ\mathcal{R}_{\delta} is BB-overlapping as in Definition 1.7.

  2. (2)

    For dyadic 0<δ<σ<10<\delta<\sigma<1, C1C\geq 1 and each subset RδδR_{\delta}\in\mathcal{R}_{\delta}, there exists C=C(C,n)C^{\prime}=C^{\prime}(C,n) such that CRδCR_{\delta} is contained in the union of at most OB,C,n(1)O_{B,C,n}(1) subsets CRσC^{\prime}R_{\sigma} where RσσR_{\sigma}\in\mathcal{R}_{\sigma}.

  3. (3)

    #δδCn\#\mathcal{R}_{\delta}\leq\delta^{-C_{n}} for some dimensional constant CnC_{n}.

The third condition is very mild. For instance, it is trivially true when each RδR_{\delta} has all dimensions δ\geq\delta.

Let k,l1k,l\geq 1. Given a C1C^{1} function r(s)r(s) defined in [1,1]k[-1,1]^{k} and map into [1,2][1,2]. Given a C2C^{2} function ψ(t)\psi(t) defined in [1,1]l[-1,1]^{l} and map into \mathbb{R}.

The canonical example of ψ(t)\psi(t) is given by ψ(t)=1|t|2\psi(t)=\sqrt{1-|t|^{2}}. A slightly different interesting example is ψ(t)=1|t|2\psi(t)=1-|t|^{2}. Both motivate the terminology “radial principle”. However, for future use, we will actually study much more general functions ψ\psi that may not have radial structure; see the beginning of Theorem 2.2 right below.

Given S[1,1]kS\subseteq[-1,1]^{k} and T[1,1]lT\subseteq[-1,1]^{l} in the parameter space, we define

(2.1) Σ0(S,T)\displaystyle\Sigma_{0}(S,T) :=Nδ({(s,t,r(s)+ψ(t)):sS,tT})\displaystyle:=N_{\delta}(\{(s,t,r(s)+\psi(t)):s\in S,t\in T\})
(2.2) Π0(S,T)\displaystyle\Pi_{0}(S,T) :=Nδ({(s,r(s)t,r(s)ψ(t)):sS,tT}).\displaystyle:=N_{\delta}(\{(s,r(s)t,r(s)\psi(t)):s\in S,t\in T\}).

By Proposition 4.7, if SS is (r,δ)(r,\delta)-flat and TT is (ψ,δ)(\psi,\delta)-flat, all in the first sense (4.1), then there exist parallelograms Σ(S,T)\Sigma(S,T), Π(S,T)\Pi(S,T) which contain and are equivalent to Σ0(S,T)\Sigma_{0}(S,T), Π0(S,T)\Pi_{0}(S,T), respectively.

Theorem 2.2 (Radial principle of decoupling).

Suppose that either

  1. (1)

    r(s)r(s) is affine, or

  2. (2)

    Lψ(t):=ψ(t)tψ(t)L\psi(t):=\psi(t)-t\cdot\nabla\psi(t) is away from 0.

Assume there are BB-overlapping dyadic meshes {𝒮δ:δ2}\{\mathcal{S}_{\delta}:\delta\in 2^{-\mathbb{N}}\}, {𝒯δ:δ2}\{\mathcal{T}_{\delta}:\delta\in 2^{-\mathbb{N}}\}, covering [1,1]k[-1,1]^{k}, [1,1]l[-1,1]^{l} respectively, and obeying the following conditions:

  1. (1)

    Each S𝒮δS\in\mathcal{S}_{\delta} is (r,δ)(r,\delta)-flat in the sense of (4.1), namely, there exist affine functions λ(s)\lambda(s), Λ(s)\Lambda(s) such that

    (2.3) |r(s)λ(s)|δ,sS.|r(s)-\lambda(s)|\leq\delta,\quad\forall s\in S.
  2. (2)

    For each T𝒯δT\in\mathcal{T}_{\delta}, 2T2T is (ψ,δ)(\psi,\delta)-flat in the sense of (4.3), namely,

    (2.4) supt2T,v𝕊l1a:t+av2T|vTD2ψ(t)v||a|2δ.\sup_{\begin{subarray}{c}t\in 2T,v\in\mathbb{S}^{l-1}\\ a\in\mathbb{R}:t+av\in 2T\end{subarray}}\left|v^{T}D^{2}\psi(t)v\right||a|^{2}\leq\delta.
  3. (3)

    Let Σ\Sigma be the subset of k+l+1\mathbb{R}^{k+l+1} given by

    Σ={(s,t,r(s)+ψ(t)):(s,t)[1,1]k×[1,1]l}.\Sigma=\{(s,t,r(s)+\psi(t)):(s,t)\in[-1,1]^{k}\times[-1,1]^{l}\}.

    Thus the collection {Σ(S,T):S𝒮δ,T𝒯δ}\{\Sigma(S,T):S\in\mathcal{S}_{\delta},T\in\mathcal{T}_{\delta}\} is a B~\tilde{B}-overlapping cover of Σ\Sigma by parallelograms, where B~\tilde{B} is independent of δ\delta.

    Fix p,q,αp,q,\alpha. Assume for each δ>0\delta>0, the following q(Lp)\ell^{q}(L^{p}) decoupling inequality holds: for test functions fS,Tf_{S,T} each Fourier supported in Σ(S,T)\Sigma(S,T), we have

    (2.5) S𝒮δT𝒯δfS,TpCεδε(#𝒮δ#𝒯δ)αfS,Tpq(S𝒮δ,T𝒯δ),ε>0,\left\lVert\sum_{S\in\mathcal{S}_{\delta}}\sum_{T\in\mathcal{T}_{\delta}}f_{S,T}\right\rVert_{p}\leq C_{\varepsilon}\delta^{-\varepsilon}(\#\mathcal{S}_{\delta}\#\mathcal{T}_{\delta})^{\alpha}\left\lVert\left\lVert f_{S,T}\right\rVert_{p}\right\rVert_{\ell^{q}(S\in\mathcal{S}_{\delta},T\in\mathcal{T}_{\delta})},\quad\forall\varepsilon>0,

    for some constant CεC_{\varepsilon} depending possibly on other parameters but independent of the test functions fS,Tf_{S,T}. Here and throughout this section, we abbreviate p=Lp(k+l+1)\left\lVert\cdot\right\rVert_{p}=\left\lVert\cdot\right\rVert_{L^{p}(\mathbb{R}^{k+l+1})}.

Consider the subset Πk+l+1\Pi\subseteq\mathbb{R}^{k+l+1} given by

(2.6) Π={(s,r(s)t,r(s)ψ(t)),(s,t)[1,1]k×[1,1]l}.\Pi=\{(s,r(s)t,r(s)\psi(t)),\,(s,t)\in[-1,1]^{k}\times[-1,1]^{l}\}.

Thus the collection {Π(S,T):S𝒮δ,T𝒯δ}\{\Pi(S,T):S\in\mathcal{S}_{\delta},T\in\mathcal{T}_{\delta}\} is a B¯\bar{B}-overlapping cover of Π\Pi by parallelograms, where B¯\bar{B} is independent of δ\delta. Moreover, for each δ>0\delta>0, the following q(Lp)\ell^{q}(L^{p}) decoupling inequality holds: for test functions gS,Tg_{S,T} Fourier supported in Π(S,T)\Pi(S,T), we have

(2.7) S𝒮δT𝒯δgS,TpCεδε(#𝒮δ#𝒯δ)αgS,Tpq(S𝒮δ,T𝒯δ),ε>0,\left\lVert\sum_{S\in\mathcal{S}_{\delta}}\sum_{T\in\mathcal{T}_{\delta}}g_{S,T}\right\rVert_{p}\leq C^{\prime}_{\varepsilon}\delta^{-\varepsilon}(\#\mathcal{S}_{\delta}\#\mathcal{T}_{\delta})^{\alpha}\left\lVert\left\lVert g_{S,T}\right\rVert_{p}\right\rVert_{\ell^{q}(S\in\mathcal{S}_{\delta},T\in\mathcal{T}_{\delta})},\quad\forall\varepsilon>0,

where CεC^{\prime}_{\varepsilon} depends at most on

ε,Cε,rC1,ψC2,inf|Lψ(t)|,k,l,B,B~,B¯,p,q,α.\varepsilon,C_{\varepsilon},\left\lVert r\right\rVert_{C^{1}},\left\lVert\psi\right\rVert_{C^{2}},\inf|L\psi(t)|,k,l,B,\tilde{B},\bar{B},p,q,\alpha.

Moreover, all implicit constants in the big-O notation above depend at most on rC1,ψC2,inf|Lψ(t)|,k,l,B,B~,B¯\left\lVert r\right\rVert_{C^{1}},\left\lVert\psi\right\rVert_{C^{2}},\inf|L\psi(t)|,k,l,B,\tilde{B},\bar{B}.

For example, in the case of the standard truncated light cone in 3\mathbb{R}^{3} studied in Section 8 of [BD15], we have the following choices:

k=l=1,r(s)=s+2,ψ(t)=1t2,\displaystyle\cdot k=l=1,\quad r(s)=s+2,\quad\psi(t)=\sqrt{1-t^{2}},
𝒮δ is the trivial partition {[1,1]},\displaystyle\cdot\mathcal{S}_{\delta}\text{ is the trivial partition $\{[-1,1]\}$},
𝒯δ is the partition of [1,1] into intervals of length δ1/2,\displaystyle\cdot\mathcal{T}_{\delta}\text{ is the partition of $[-1,1]$ into intervals of length $\delta^{1/2}$},
B(C)=B~(C)=B¯(C)=(100C)100,\displaystyle\cdot B(C)=\tilde{B}(C)=\bar{B}(C)=(100C)^{100},
α=0,p=6,q=2.\displaystyle\cdot\alpha=0,\quad p=6,\quad q=2.

Thus we have the following chain of reduction of decoupling:

(s,st,s1t2)\displaystyle(s,st,s\sqrt{1-t^{2}})
\displaystyle\implies (s,t,s+1t2)by radial principle,\displaystyle(s,t,s+\sqrt{1-t^{2}})\quad\text{by radial principle},
\displaystyle\implies (s,t,1t2)by affine equivalence,\displaystyle(s,t,\sqrt{1-t^{2}})\quad\text{by affine equivalence},
\displaystyle\implies (t,1t2)by projection principle,\displaystyle(t,\sqrt{1-t^{2}})\quad\text{by projection principle},

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 D(δ)D(\delta) the smallest constant such that for test functions gS,Tg_{S,T} Fourier supported in Π(S,T)\Pi(S,T), we have

(2.8) S𝒮δT𝒯δgS,TpD(δ)(#𝒮δ#𝒯δ)αgS,Tpq(S𝒮δ,T𝒯δ).\left\lVert\sum_{S\in\mathcal{S}_{\delta}}\sum_{T\in\mathcal{T}_{\delta}}g_{S,T}\right\rVert_{p}\leq D(\delta)(\#\mathcal{S}_{\delta}\#\mathcal{T}_{\delta})^{\alpha}\left\lVert\left\lVert g_{S,T}\right\rVert_{p}\right\rVert_{\ell^{q}(S\in\mathcal{S}_{\delta},T\in\mathcal{T}_{\delta})}.

Our goal is to show that

(2.9) D(δ)εδε,ε>0.D(\delta)\lesssim_{\varepsilon}\delta^{-\varepsilon},\quad\forall\varepsilon>0.

Let ε>0\varepsilon>0. We first partition [1,1]k[-1,1]^{k} into cubes R0R_{0} of length δε\delta^{\varepsilon}. Since we allow ε\varepsilon losses, we may just decouple each R0R_{0}. By translation in ss, we may assume without loss of generality that the centre of R0R_{0} is the origin. By dividing the last two coordinates by r(0)r(0), we may assume without loss of generality that r(0)=1r(0)=1.

We record that

(2.10) |r(s)1|δε,sR0.|r(s)-1|\lesssim\delta^{\varepsilon},\quad\forall s\in R_{0}.

By abuse of notation we may also regard D(δ)D(\delta) as the best decoupling constant adapted to the smaller cubes.

2.3. Intermediate scale decoupling

Let σ=δ1ε\sigma=\delta^{1-\varepsilon}. Since Nδ(Π)Nσ(Π)N_{\delta}(\Pi)\subseteq N_{\sigma}(\Pi), applying the definition of D(σ)D(\sigma) gives

(2.11) S𝒮δT𝒯δgS,Tp(#𝒮σ#𝒯σ)αD(σ)gS,Tpq(S𝒮σ,T𝒯σ),\left\lVert\sum_{S\in\mathcal{S}_{\delta}}\sum_{T\in\mathcal{T}_{\delta}}g_{S,T}\right\rVert_{p}\leq(\#\mathcal{S}_{\sigma}\#\mathcal{T}_{\sigma})^{\alpha}D(\sigma)\left\lVert\left\lVert g_{S^{\prime},T^{\prime}}\right\rVert_{p}\right\rVert_{\ell^{q}(S^{\prime}\in\mathcal{S}_{\sigma},T^{\prime}\in\mathcal{T}_{\sigma})},

where gS,T:=S𝒮δ(S)T𝒯δ(T)gS,Tg_{S^{\prime},T^{\prime}}:=\sum_{S\in\mathcal{S}_{\delta}(S^{\prime})}\sum_{T\in\mathcal{T}_{\delta}(T^{\prime})}g_{S,T}, and where 𝒮δ(S)\mathcal{S}_{\delta}(S^{\prime}) the collection of subsets S𝒮δS\in\mathcal{S}_{\delta} that lie within SS^{\prime}, and similarly for TT^{\prime}. By Part (2) in the definition of a dyadic mesh, it then suffices to decouple gS,Tg_{S^{\prime},T^{\prime}} for each fixed pair of intermediate scale subsets S,TS^{\prime},T^{\prime}.

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 TT^{\prime} be centred at 0, and that ψ(0)=0\nabla\psi(0)=0. Next, also by affine invariance, we can let r(0)=0\nabla r(0)=0. Then by Taylor expansion, we can write

r(s)ψ(t)=(1+E(s))(ψ(0)+Q(t)),r(s)\psi(t)=(1+E(s))(\psi(0)+Q(t)),

where E(s)E(s) and Q(t)Q(t) are σ\sigma-flat over SS^{\prime} and TT^{\prime}, respectively, and their gradients are 0 at 0. Thus the last coordinate can be further simplified to E(s)ψ(0)+Q(t)E(s)\psi(0)+Q(t), since E(s)Q(t)=O(σ2)δE(s)Q(t)=O(\sigma^{2})\leq\delta. But then reversing the affine invariance, the decoupling of E(s)ψ(0)+Q(t)E(s)\psi(0)+Q(t) is equivalent to r(s)Lψ(0)+ψ(t)r(s)L\psi(0)+\psi(t). More precisely, we now need to decouple the manifold parametrised by

(s,t,r(s)Lψ(0)+ψ(t)).(s,t,r(s)L\psi(0)+\psi(t)).

If r(s)r(s) is affine, then the decoupling problem reduces to that for

(s,t,ψ(t)).(s,t,\psi(t)).

If Lψ(0)0L\psi(0)\neq 0, the decoupling problem reduces that for

(s,t,r(s)+ψ(t)).(s,t,r(s)+\psi(t)).

We now provide the details. Let TT^{\prime} be centred at cc. Denote by ϕ(t)=ψ(t+c)\phi(t)=\psi(t+c). By a translation, the manifold (s,r(s)t,r(s)ψ(t))(s,r(s)t,r(s)\psi(t)) can be written as

(2.12) (s,r(s)(t+c),r(s)ϕ(t)),(s,r(s)(t+c),r(s)\phi(t)),

where tTt\in T^{\prime} and TT^{\prime} is centred at 0.

2.4.1. Affine invariance in ϕ\phi

Write ϕ(t)=ϕ(0)+ϕ(0)t+Q(t)\phi(t)=\phi(0)+\nabla\phi(0)\cdot t+Q(t). By a linear transformation, we may transform the decoupling problem to the manifold

(2.13) (s,r(s)(t+c),r(s)(ϕ(0)ϕ(0)c+Q(t))).(s,r(s)(t+c),r(s)(\phi(0)-\nabla\phi(0)\cdot c+Q(t))).

Note that ϕ(0)ϕ(0)c=ψ(c)cψ(t)=Lψ(c)\phi(0)-\nabla\phi(0)\cdot c=\psi(c)-c\cdot\nabla\psi(t)=L\psi(c). Also, note that Q(0)=0\nabla Q(0)=0 by this reduction.

2.4.2. Affine invariance in rr

Write r(s)=1+r(0)s+E(s)r(s)=1+\nabla r(0)s+E(s), where |E(s)|σ|E(s)|\leq\sigma. Using an affine transformation, we may transform the decoupling problem to the manifold

(2.14) (s,r(s)t+E(s)c,r(s)(Lψ(c)+Q(t))).(s,r(s)t+E(s)c,r(s)(L\psi(c)+Q(t))).

2.4.3. Approximation

We now denote τ=r(s)t+E(s)c\tau=r(s)t+E(s)c. We first claim that τ2T\tau\in 2T^{\prime}. Indeed, for any v𝕊l1v\in\mathbb{S}^{l-1}, we compute

|(τt)v|\displaystyle|(\tau-t)\cdot v| =|(r(s)1)tv+E(s)cv|\displaystyle=|(r(s)-1)t\cdot v+E(s)c\cdot v|
δε|projv(T)|+σ\displaystyle\lesssim\delta^{\varepsilon}|\mathrm{proj}_{v}(T^{\prime})|+\sigma
δε|projv(T)|,\displaystyle\sim\delta^{\varepsilon}|\mathrm{proj}_{v}(T^{\prime})|,

since TT^{\prime} has all dimensions at least σ1/2=δ1ε2\sigma^{1/2}=\delta^{\frac{1-\varepsilon}{2}}. This means that τtδεT\tau-t\in\delta^{\varepsilon}T^{\prime}, and in particular, τ2T\tau\in 2T^{\prime}.

We then approximate the manifold in (2.14) by

(2.15) (s,τ,r(s)Lψ(c)+Q(τ)).(s,\tau,r(s)L\psi(c)+Q(\tau)).

Indeed, one need to compute the difference

r(s)(Lψ(c)+Q(t))(r(s)Lψ(c)+Q(τ))\displaystyle r(s)(L\psi(c)+Q(t))-(r(s)L\psi(c)+Q(\tau))
=(r(s)1)Q(t)+Q(t)Q(τ).\displaystyle=(r(s)-1)Q(t)+Q(t)-Q(\tau).

By (2.4), we have in particular Q(t)=O(σ)Q(t)=O(\sigma). Thus by (2.10), we have (r(s)1)Q(t)=O(σδε)=O(δ)(r(s)-1)Q(t)=O(\sigma\delta^{\varepsilon})=O(\delta).

It remains to estimate Q(t)Q(τ)Q(t)-Q(\tau). We first apply an affine rescaling l:[1,1]l2Tl^{\prime}:[-1,1]^{l}\to 2T^{\prime}, and denote Q=QlQ^{\prime}=Q\circ l^{\prime}. Also, denote t=l1tt^{\prime}=l^{\prime-1}t, τ=l1τ\tau^{\prime}=l^{\prime-1}\tau. Since τtδεT\tau-t\in\delta^{\varepsilon}T^{\prime}, we have |τt|δε|\tau^{\prime}-t^{\prime}|\lesssim\delta^{\varepsilon}. Also, by (2.4) and affine invariance (see Proposition 4.4), we have that

(2.16) supv𝕊l1supx[1,1]l|vTD2Q(x)v|σ.\sup_{v\in\mathbb{S}^{l-1}}\sup_{x\in[-1,1]^{l}}|v^{T}D^{2}Q^{\prime}(x)v|\lesssim\sigma.

Then by the elementary Lemma 2.3 below, we have

|Q(t)Q(τ)|\displaystyle|Q(t)-Q(\tau)| =|Q(t)Q(τ)|\displaystyle=|Q^{\prime}(t^{\prime})-Q^{\prime}(\tau^{\prime})|
supv𝕊l1supx[1,1]l|vTD2Q(x)v||tτ|\displaystyle\lesssim\sup_{v\in\mathbb{S}^{l-1}}\sup_{x\in[-1,1]^{l}}|v^{T}D^{2}Q^{\prime}(x)v||t^{\prime}-\tau^{\prime}|
σδε=δ.\displaystyle\lesssim\sigma\delta^{\varepsilon}=\delta.

This finishes the approximation assuming the lemma.

Lemma 2.3.

Let T[1,1]lT\subseteq[-1,1]^{l} be a parallelogram centred at 0. Let Q:TQ:T\to\mathbb{R} be a C2C^{2} function with Q(0)=0\nabla Q(0)=0. For all tτTt\neq\tau\in T, denote by vv the unit vector in the same direction as tτt-\tau. Then we have

|Q(t)Q(τ)|suptT|vTD2Q(t)v||tτ|(|t|+|τ|).|Q(t)-Q(\tau)|\lesssim\sup_{t^{\prime}\in T}|v^{T}D^{2}Q(t^{\prime})v||t-\tau|(|t|+|\tau|).
Proof of lemma.

Given tτTt\neq\tau\in T. Then

|Q(t)Q(τ)|\displaystyle|Q(t)-Q(\tau)| =|01Q((1u)τ+ut)(tτ)𝑑u|\displaystyle=\left|\int_{0}^{1}\nabla Q((1-u)\tau+ut)\cdot(t-\tau)du\right|
=|01Q((1u)τ+ut)(tτ)𝑑u01Q(0)(tτ)𝑑u|\displaystyle=\left|\int_{0}^{1}\nabla Q((1-u)\tau+ut)\cdot(t-\tau)du-\int_{0}^{1}\nabla Q(0)\cdot(t-\tau)du\right|
|tτ|01|Q((1u)τ+ut)Q(0)|𝑑u\displaystyle\leq|t-\tau|\int_{0}^{1}|\nabla Q((1-u)\tau+ut)-\nabla Q(0)|du
suptT|vTD2Q(t)v||tτ|(|t|+|τ|).\displaystyle\lesssim\sup_{t^{\prime}\in T}|v^{T}D^{2}Q(t^{\prime})v||t-\tau|(|t|+|\tau|).

Therefore, we have reduced the decoupling of gS,Tg_{S^{\prime},T^{\prime}} to that for the manifold

(2.17) {(s,τ,r(s)Lψ(c)+Q(τ)):sS,τ2T}.\{(s,\tau,r(s)L\psi(c)+Q(\tau)):s\in S^{\prime},\tau\in 2T^{\prime}\}.

If either |Lψ(t)|1|L\psi(t)|\sim 1 or r(s)r(s) affine, then we can reverse the affine invariance in the τ\tau variable, and so we have reduced to the decoupling for r(s)+ψ(t)r(s)+\psi(t).

2.5. Applying decoupling assumption

Now we can apply (2.5) with ε2\varepsilon^{2} in place of ε\varepsilon to get that

(2.18) gS,TpCε2δε2(#𝒮δ(S)#𝒯δ(T))αgS,Tpq(S𝒮δ(S),T𝒯δ(T)),ε>0.\left\lVert g_{S^{\prime},T^{\prime}}\right\rVert_{p}\leq C_{\varepsilon^{2}}\delta^{-\varepsilon^{2}}(\#\mathcal{S}_{\delta}(S^{\prime})\#\mathcal{T}_{\delta}(T^{\prime}))^{\alpha}\left\lVert\left\lVert g_{S,T}\right\rVert_{p}\right\rVert_{\ell^{q}(S\in\mathcal{S}_{\delta}(S^{\prime}),T\in\mathcal{T}_{\delta}(T^{\prime}))},\quad\forall\varepsilon>0.

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 #𝒮δ(S)\#\mathcal{S}_{\delta}(S^{\prime}) is essentially a constant for each SS^{\prime} at a cost of logδ1εδε2\log\delta^{-1}\lesssim_{\varepsilon}\delta^{-\varepsilon^{2}}. Similarly, we may assume #𝒯δ(T)\#\mathcal{T}_{\delta}(T^{\prime}) is essentially a constant for each TT^{\prime}.

Combined with (2.11), we thus have for some C~ε\tilde{C}_{\varepsilon} that

(2.19) S𝒮δT𝒯δfS,TpC~εδε2D(σ)(#𝒮δ#𝒯δ)αfS,Tpq(S𝒮δ,T𝒯δ).\left\lVert\sum_{S\in\mathcal{S}_{\delta}}\sum_{T\in\mathcal{T}_{\delta}}f_{S,T}\right\rVert_{p}\leq\tilde{C}_{\varepsilon}\delta^{-\varepsilon^{2}}D(\sigma)(\#\mathcal{S}_{\delta}\#\mathcal{T}_{\delta})^{\alpha}\left\lVert\left\lVert f_{S,T}\right\rVert_{p}\right\rVert_{\ell^{q}(S\in\mathcal{S}_{\delta},T\in\mathcal{T}_{\delta})}.

Recall σ=δ1ε\sigma=\delta^{1-\varepsilon}. This implies the following bootstrap inequality:

(2.20) D(δ)D(δ1ε)C~εδε2.D(\delta)\leq D(\delta^{1-\varepsilon})\tilde{C}_{\varepsilon}\delta^{-\varepsilon^{2}}.

Iterating this bootstrap inequality gives us D(δ)εδ2εD(\delta)\lesssim_{\varepsilon}\delta^{-2\varepsilon}.

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 1kn11\leq k\leq n-1. Let 0=0(n,k)\mathcal{M}_{0}=\mathcal{M}_{0}(n,k) be the collection of manifolds in n\mathbb{R}^{n} consisting of MϕM_{\phi} given by the graph of a vector valued C1C^{1}555Typically, we consider smooth manifolds. The minimal requirement here is that the C1C^{1} norms of the functions ϕ\phi are bounded uniformly. function ϕ:Rϕnk\phi:R_{\phi}\to\mathbb{R}^{n-k} for some parallelogram Rϕ[1,1]kR_{\phi}\subseteq[-1,1]^{k}.

Throughout the rest of of this section, we fix some integer m[1,n]m\in[1,n], and all flatness conditions are stated in dimension m1m-1 (see Definitions 1.2 and 1.3).

Definition 3.1 (Affine equivalence).

Define an equivalence relation \sim on 0\mathcal{M}_{0} as follows. MϕMψM_{\phi}\sim M_{\psi} if there exist affine transformations L1,L2L_{1},L_{2} with |detL1|=1|\det L_{1}|=1 such that L1MϕL2=MψL_{1}\circ M_{\phi}\circ L_{2}=M_{\psi}. In other words, MϕMψM_{\phi}\sim M_{\psi} if there exist A(nk)×kA\in\mathbb{R}^{(n-k)\times k} with |detA|=1|\det A|=1 and bkb\in\mathbb{R}^{k} such that ϕ(L2(x))(Ax+b)=ψ(x)\phi(L_{2}(x))-(Ax+b)=\psi(x) for all xRψx\in R_{\psi}.

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 Mϕ,Mψ0M_{\phi},M_{\psi}\in\mathcal{M}_{0} are equivalent. For cnk\vec{c}\in\mathbb{R}^{n-k} with supi|ci|1\sup_{i}|c_{i}|\leq 1 and c:=infi|ci|>0c:=\inf_{i}|c_{i}|>0, denote by cψ\vec{c}\psi the nk\mathbb{R}^{n-k}-valued function

(c1ψ1,,cnkψnk).(c_{1}\psi_{1},\dots,c_{n-k}\psi_{n-k}).

For δ>0\delta>0 and fixed p,q,αp,q,\alpha, the following two statements are equivalent:

  1. (1)

    Nδϕ(Rϕ)N_{\delta}^{\phi}(R_{\phi}) can be (q(Lp),α)(\ell^{q}(L^{p}),\alpha) decoupled into parallelograms at cost CC.

  2. (2)

    Nδψ(Rψ)N_{\delta}^{\psi}(R_{\psi}) can be (q(Lp),α)(\ell^{q}(L^{p}),\alpha) decoupled into parallelograms at cost CC.

Either statement above implies:

(3)\quad\,\,\,\,\mathrm{(3)} Ncδcψ(Rψ)N^{\vec{c}\psi}_{c\delta}(R_{\psi}) can be (q(Lp),α)(\ell^{q}(L^{p}),\alpha) decoupled into parallelograms at cost CC.

Furthermore, if ci=cc_{i}=c for all 1ink1\leq i\leq n-k, then all the above are equivalent to

(4)\quad\,\,\,\,\mathrm{(4)} Ncδ(Mcψ)N_{c\delta}(M_{\vec{c}\psi}) can be (q(Lp),α)(\ell^{q}(L^{p}),\alpha) decoupled into parallelograms at cost CC.

Let 0\mathcal{R}_{0} be the collection of all parallelograms contained in [1,1]k[-1,1]^{k}.

Definition 3.3 (Rescaling invariance).

Let \mathcal{M} be a subfamily of 0\mathcal{M}_{0}. A collection 0\mathcal{R}\subseteq\mathcal{R}_{0} of parallelograms in [1,1]k[-1,1]^{k} is said to be rescaling invariant with respect to \mathcal{M} if the following holds. For every σ(0,1]\sigma\in(0,1], MϕM_{\phi}\in\mathcal{M} and every (ϕ,σ)(\phi,\sigma)-flat parallelogram RR\in\mathcal{R} with RRϕR\subseteq R_{\phi}, the manifold Mϕ(R×nk)M_{\phi}\cap(R\times\mathbb{R}^{n-k}), as an element of 0\mathcal{M}_{0}, is equivalent to MσψM_{\vec{\sigma}\psi} for some MψM_{\psi}\in\mathcal{M} and σnk\vec{\sigma}\in\mathbb{R}^{n-k} with infiσi=σ\inf_{i}\sigma_{i}=\sigma.

Moreover, a collection 0\mathcal{R}\subseteq\mathcal{R}_{0} of parallelograms in [1,1]k[-1,1]^{k} is said to be non-rotational rescaling invariant with respect to \mathcal{M} if we further assume that the transformation L2L_{2} in the equivalence definition Mϕ(R×nk)MσψM_{\phi}\cap(R\times\mathbb{R}^{n-k})\sim M_{\vec{\sigma}\psi} does not involve rotation, i.e. DL2DL_{2} is a diagonal matrix.

In simple words, every flat piece of a manifold in \mathcal{M} over RR\in\mathcal{R} can be rescaled to become another manifold in \mathcal{M}. We illustrate Lemma 3.2 and Definition 3.3 using the following example. Consider the moment curve MϕM_{\phi} given by (s,ϕ(s))=(s,s2,s3)(s,\phi(s))=(s,s^{2},s^{3}), s[1,1]s\in[-1,1] in 3\mathbb{R}^{3}. Let σ(0,1)\sigma\in(0,1).

  • The interval R=[0,σ1/3]R=[0,\sigma^{1/3}] is (ϕ,σ)(\phi,\sigma)-flat in dimension 22. The manifold Mϕ(R×2)M_{\phi}\cap(R\times\mathbb{R}^{2}) can be rescaled to become

    {(s,σ2/3s2,σs3):s[1,1]},\{(s,\sigma^{2/3}s^{2},\sigma s^{3}):s\in[-1,1]\},

    which is MσϕM_{\vec{\sigma}\phi} where σ:=(σ2/3,σ)\vec{\sigma}:=(\sigma^{2/3},\sigma) so that infiσi=σ\inf_{i}\sigma_{i}=\sigma. A similar argument shows that other (ϕ,σ)(\phi,\sigma)-flat intervals behave similarly. Therefore, the collection of all closed intervals is rescaling invariance with respect to the singleton family ={ϕ}\mathcal{M}=\{\mathcal{M}_{\phi}\}. In this case, Lemma 3.2 is exactly the key rescaling property of the moment curve used in [BDG16].

  • The interval R=[0,σ1/2]R=[0,\sigma^{1/2}] is (ϕ,σ)(\phi,\sigma)-flat in dimension 11. The manifold Mϕ(R×2)M_{\phi}\cap(R\times\mathbb{R}^{2}) can be rescaled to become

    {(s,σs2,σ3/2s3):s[1,1]},\{(s,\sigma s^{2},\sigma^{3/2}s^{3}):s\in[-1,1]\},

    which is MσϕM_{\vec{\sigma}\phi} where σ:=(σ,σ)\vec{\sigma}:=(\sigma,\sigma) so that infiσi=σ\inf_{i}\sigma_{i}=\sigma, and ψ(s)=(s2,σ1/2s3)\psi(s)=(s^{2},\sigma^{1/2}s^{3}). The idea here is that we are ignoring the torsion given by the third coordinate s3s^{3}. We can still get 2(L6)\ell^{2}(L^{6}) 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 \mathcal{M} containing all manifolds MϕM_{\phi} such that Rϕ=[1,1]R_{\phi}=[-1,1], and ϕ\phi is of the form (s2,ϕ2(s))(s^{2},\phi_{2}(s)), where ϕ2(s)\phi_{2}(s) is an arbitrary continuous function.

Definition 3.4 (Degeneracy determinant).

Let \mathcal{M} be a subfamily of 0\mathcal{M}_{0} and \mathcal{R} be rescaling invariant with respect to \mathcal{M}. We say an operator HH on \mathcal{M} that sends MϕM_{\phi}\in\mathcal{M} to a smooth function on RϕR_{\phi} is a degeneracy determinant if the following holds.

  1. (1)

    For every MϕM_{\phi}\in\mathcal{M}, HMϕHM_{\phi} is Lipschitz, namely,

    |HMϕ(x)HMϕ(y)|C0|xy|,x,yRϕ,|HM_{\phi}(x)-HM_{\phi}(y)|\leq C_{0}|x-y|,\quad\forall x,y\in R_{\phi},

    where C01C_{0}\geq 1 depends on ,\mathcal{M},\mathcal{R} only.

  2. (2)

    There are constants C1C\geq 1 and β(0,1]\beta\in(0,1], both depending on ,\mathcal{M},\mathcal{R} only, such that for any MϕM_{\phi}\in\mathcal{M} and RR\in\mathcal{R}, if we have HMϕL(R)σ\left\lVert HM_{\phi}\right\rVert_{L^{\infty}(R)}\leq\sigma, then there exists MψM_{\psi}\in\mathcal{M} such that RRψR\subseteq R_{\psi}, Hψ0H\psi\equiv 0, and

    (3.1) ϕiψiL(R)Cσβ,1ink.\|\phi_{i}-\psi_{i}\|_{L^{\infty}(R)}\leq C\,\sigma^{\beta},\quad\forall 1\leq i\leq n-k.

3.2. Statement of theorem

Theorem 3.5.

Let HH be a degeneracy determinant defined with respect to a pair ,\mathcal{M},\mathcal{R}. Let p,q[2,]p,q\in[2,\infty], α[0,)\alpha\in[0,\infty), ε>0\varepsilon>0, and suppose we have the following assumptions, where all implicit constants and overlap bounds BB below depend only on ,,ε,p,q,α\mathcal{M},\mathcal{R},\varepsilon,p,q,\alpha.

  1. (1)

    (Sublevel set decoupling) For each MϕM_{\phi}\in\mathcal{M} and each σ(0,1]\sigma\in(0,1], the set

    {xRϕ:|HMϕ(x)|σ}\{x\in R_{\phi}:|HM_{\phi}(x)|\leq\sigma\}

    can be (q(Lp),α)(\ell^{q}(L^{p}),\alpha) decoupled into parallelograms uniformly at a cost of Oε(σε)O_{\varepsilon}(\sigma^{-\varepsilon}), on each of which |HMϕ|σ|HM_{\phi}|\lesssim\sigma.

  2. (2)

    (Degenerate decoupling) If MϕM_{\phi}\in\mathcal{M} is such that HMϕ0HM_{\phi}\equiv 0, then for each σ(0,1]\sigma\in(0,1], RϕR_{\phi} can be graphically (q(Lp),α)(\ell^{q}(L^{p}),\alpha)-decoupled into (ϕ,σ)(\phi,\sigma)-flat parallelograms Rdeg(ϕ)R\in\mathcal{R}_{deg}(\phi)\subseteq\mathcal{R} at a cost of O(σεO(\sigma^{-\varepsilon}), and each deg(ϕ)\mathcal{R}_{deg}(\phi) is BB-overlapping.

  3. (3)

    (Nondegnerate decoupling) If MϕM_{\phi}\in\mathcal{M} such that infRϕ|HMϕ|K1\inf_{R_{\phi}}|HM_{\phi}|\geq K^{-1}, then for each δ>0\delta>0, RϕR_{\phi} can be graphically decoupled into (ϕ,δ)(\phi,\delta)-flat parallelograms Rnon(ϕ)R\in\mathcal{R}_{non}(\phi) at a cost of KO(1)δεK^{O(1)}\delta^{-\varepsilon}, and each non(ϕ)\mathcal{R}_{non}(\phi) is BB-overlapping.

Then for each MϕM_{\phi}\in\mathcal{M}, [1,1]k[-1,1]^{k} can be graphically (q(Lp),α)(\ell^{q}(L^{p}),\alpha) decoupled into (ϕ,δ)(\phi,\delta)-flat parallelograms (ϕ)\mathcal{R}(\phi), with the implicit constants and overlap bounds depending only on ,,ε,p,q,α\mathcal{M},\mathcal{R},\varepsilon,p,q,\alpha.

Moreover, if the collection \mathcal{R} is non-rotational rescaling invariant with respect to \mathcal{M}, and each collection deg(ϕ),non(ϕ)\mathcal{R}_{deg}(\phi),\mathcal{R}_{non}(\phi) in Assumptions (2) and (3) above has non-overlapping interiors, the final collection (ϕ)\mathcal{R}(\phi) has non-overlapping interiors.

Note that only the second assumption requires decoupling into the rescaling invariant subcollection \mathcal{R}.

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 MϕM_{\phi}\in\mathcal{M} Parallelograms in \mathcal{R} HMϕHM_{\phi}
1.4 ϕ(s,t)=P(s)+ψ(t)\phi(s,t^{\prime})=P(s)+\psi(t^{\prime}), (s,t)I×Q(s,t^{\prime})\in I\times Q, P′′(s)P^{\prime\prime}(s)
PP polynomial, ψ(t)=1|t|2\psi(t^{\prime})=\sqrt{1-|t^{\prime}|^{2}} QQ axis-parallel square
1.5 ϕ(x1,x2)=ϕ1(x1)+ϕ2(x2)\phi(x_{1},x_{2})=\phi_{1}(x_{1})+\phi_{2}(x_{2})666For simplicity, we only write down the case for the family 1n\mathcal{M}_{1}^{n}, J=2,n=3,4,5J=2,n=3,4,5. The other families are straightforward generalisations of this case. (x1,x2)R×Q(x_{1},x_{2})\in R\times Q, detD2ϕ1\det D^{2}\phi_{1}
polynomial, |detD2ϕ2|1|\det D^{2}\phi_{2}|\sim 1 QQ axis-parallel square
1.6 ϕ:n1\phi:\mathbb{R}\to\mathbb{R}^{n-1} polynomials, all intervals det((Wmϕ)\det((W_{m}\phi)
|HMϕ|1|HM_{\phi}|\sim 1 or MϕM_{\phi} is straight line i.e. 0\mathcal{R}_{0} (Wmϕ)T)(W_{m}\phi)^{T})
Table 3. Applications of the degeneracy locating principles in the corollaries. Here polynomial means a family of polynomial in appropriate number of variables of degree at most dd and bounded coefficients.

Regarding overlapping conditions, it is clear that the families \mathcal{R} in Corollaries 1.4 and 1.6 above are non-rotational invariant with respect to the corresponding families \mathcal{M}. 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], (ϕ,δ)(\phi,\delta)-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 KK (the scale MM in [LY23]), which allows us to improve the overlapping bound of the parallelograms from CεδεC_{\varepsilon}\delta^{-\varepsilon} to CεC_{\varepsilon}. This improvement is nontrivial, in view of Appendix A of [GMO24].

For the rest of this section, we suppress the subscripts j,k,p,q,αj,k,p,q,\alpha unless otherwise specified.

3.3. The base case

We may assume δε1\delta\ll_{\varepsilon}1, otherwise the decoupling inequality is trivial.

Let K=δεK=\delta^{-\varepsilon}. Our induction hypothesis is that the conclusion holds at all coarser scales δKβδ\delta^{\prime}\gtrsim K^{\beta}\delta, where β\beta is as in Definition 3.4.

3.4. Degenerate and nondegenerate parts

Partition RϕR_{\phi} into two parts RnonR_{\text{non}} and RdegenR_{\text{degen}} as follows:

Rnon:={xRϕ:|HMϕ(x)|>K1}\displaystyle R_{\text{non}}:=\{x\in R_{\phi}:|HM_{\phi}(x)|>K^{-1}\}
Rdegen:={xRϕ:|HMϕ(x)|K1}.\displaystyle R_{\text{degen}}:=\{x\in R_{\phi}:|HM_{\phi}(x)|\leq K^{-1}\}.

We first deal with the subset RnonR_{\text{non}}. We can cover it by squares QQ of side length δ1/2\delta^{1/2}, and if δ1\delta\ll 1, then by the Lipschitz assumption in Definition 3.4, we have |HMϕ(x)|K1/2|HM_{\phi}(x)|\geq K^{-1}/2 on each such square. Then we can decouple RnonR_{\text{non}} by these squares QQ, by the nondegenerate decoupling assumption (3).

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 σ:=K1\sigma:=K^{-1} to decouple SdegenS_{\text{degen}} into parallelograms RR on each of which HMϕL(R)K1\|HM_{\phi}\|_{L^{\infty}(R)}\lesssim K^{-1}. It remains to further decouple each RR into smaller (ϕ,δ)(\phi,\delta)-flat parallelograms.

3.5.2. Approximation and Degenerate decoupling

By the definition of degeneracy determinant HH, since HMϕL(R)K1\|HM_{\phi}\|_{L^{\infty}(R)}\lesssim K^{-1}, there exists MψM_{\psi}\in\mathcal{M} such that Hψ0H\psi\equiv 0 and ϕψKβ\left\lVert\phi-\psi\right\rVert_{\infty}\lesssim K^{-\beta}. Apply the degenerate decoupling assumption (2) at scale σKβ\sigma\sim K^{-\beta} to the function ψ\psi to graphically decouple RR into (ψ,O(Kβ))(\psi,O(K^{-\beta}))-flat parallelograms RR^{\prime}\in\mathcal{R}. Thus RR^{\prime} is also (ϕ,O(Kβ))(\phi,O(K^{-\beta}))-flat.

3.5.3. Rescaling

By the definition of rescaling invariance of \mathcal{R}, we see that Mϕ(R×nk)M_{\phi}\cap(R^{\prime}\times\mathbb{R}^{n-k}) is equivalent to MCKβψM_{CK^{-\beta}\psi} for some MψM_{\psi}\in\mathcal{M}. Hence, by Lemma 3.2, the graphical decoupling of RψR_{\psi} into (ψ,CKβδ)(\psi,CK^{\beta}\delta)-flat parallelograms is implied by the graphical decoupling of RR^{\prime} into (ϕ,δ)(\phi,\delta)-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 D(δ)D(\delta) the cost of graphical decoupling of [1,1]k[-1,1]^{k} into (ϕ,δ)(\phi,\delta)-flat parallelograms. Our goal is to show that D(δ)CεδεD(\delta)\leq C_{\varepsilon}\delta^{-\varepsilon}.

The first difference is that we now need to combine q(Lp)\ell^{q}(L^{p}) decoupling inequalities obtained at each step, which is done by a simple dyadic decomposition argument with a logarithmic loss O(logK)O(\log K) at each step (see Proposition 4.9). This is acceptable since we already lose KεK^{\varepsilon} at each step of the induction. The second difference is the choice of K=δεK=\delta^{-\varepsilon} instead of Kε1K\sim_{\varepsilon}1 in [LY23] (where we used MM in place of KK). 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) D(δ)Cεδε+CεD(Kβδ)KεCεδε+CεD(Kβδ)δε2.D(\delta)\leq C_{\varepsilon}\delta^{-\varepsilon}+C_{\varepsilon}D(K^{\beta}\delta)K^{\varepsilon}\leq C_{\varepsilon}\delta^{-\varepsilon}+C_{\varepsilon}D(K^{\beta}\delta)\delta^{-\varepsilon^{2}}.

(Here we have omitted the unimportant implicit constants.)

3.6.1. Iteration

Iterating inequality (3.2) for NN times, we obtain

D(δ)\displaystyle D(\delta) Cεδε(1+δε2+δ2ε2++δNε2)+δNε2D(KNβδ).\displaystyle\leq C_{\varepsilon}\delta^{-\varepsilon}(1+\delta^{-\varepsilon^{2}}+\delta^{-2\varepsilon^{2}}+\cdots+\delta^{-N\varepsilon^{2}})+\delta^{-N\varepsilon^{2}}D(K^{N\beta}\delta).

We will stop this iteration once we reach KNβδ1K^{N\beta}\delta\sim 1, that is, when

(3.3) N(βε)1.N\sim(\beta\varepsilon)^{-1}.

In this case we have D(KβNδ)1D(K^{\beta N}\delta)\sim 1, and

1+δε2+δ2ε2++δNε2+δNε2D(KNβδ)δO(ε),1+\delta^{-\varepsilon^{2}}+\delta^{-2\varepsilon^{2}}+\cdots+\delta^{-N\varepsilon^{2}}+\delta^{-N\varepsilon^{2}}D(K^{N\beta}\delta)\lesssim\delta^{-O(\varepsilon)},

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 K=δεK=\delta^{-\varepsilon} leads to N=O(1/ε)N=O(1/\varepsilon), and so we may just roughly bound the overlapping number by Oε(1)O_{\varepsilon}(1).

We remark that this is a slight improvement of the corresponding argument in [LY23], where we were only able to obtain O(δε)O(\delta^{-\varepsilon}) overlap.

Moreover, if the rescaling in Subsection 3.5.3 does not involve rotation, the final collection (ϕ)\mathcal{R}(\phi) 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. (1)

    If PP is a convex body in n\mathbb{R}^{n}, then there exist a parallelogram TT and a constant C=CnC=C_{n} such that C1TPCTC^{-1}T\subseteq P\subseteq CT.

  2. (2)

    Let T1,T2T_{1},T_{2} be parallelograms in n\mathbb{R}^{n} with nonempty intersection. Then there exists some degree constant C=10nC=10n and some parallelogram TT such that

    T1T2TCT1CT2,T_{1}\cap T_{2}\subseteq T\subseteq CT_{1}\cap CT_{2},

Here CTCT means the dilation of TT with respect to its centre by a factor of CC.

This follows from an easy application of the John Ellipsoid Theorem [Joh14]. Alternatively, the reader could read Proposition 4.10 of [LY23] for an elementary proof in the case n=2n=2.

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 nn-variate real polynomial P=αcαxαP=\sum_{\alpha}c_{\alpha}x^{\alpha} of degree at most dd, we have the following relation:

sup[1,1]n|P|n,dmaxα|cα|.\sup_{[-1,1]^{n}}|P|\sim_{n,d}\max_{\alpha}|c_{\alpha}|.

As a result, we also have the following relations:

sup[1,1]n|P|n,dα|cα|\displaystyle\sup_{[-1,1]^{n}}|P|\sim_{n,d}\sum_{\alpha}|c_{\alpha}|
sup[C,C]n|P|n,dsupBn(0,C)|P|n,dCdsup[1,1]n|P|,\displaystyle\sup_{[-C,C]^{n}}|P|\sim_{n,d}\sup_{B^{n}(0,C)}|P|\lesssim_{n,d}C^{d}\sup_{[-1,1]^{n}}|P|,

for any constant C1C\geq 1.

For a proof of this proposition, the reader may consult [Kel28].

Corollary 4.3.

For any parallelogram TnT\subset\mathbb{R}^{n} and any constant C1C\geq 1,

supT|P|supCT|P|n,dCdsupT|P|.\sup_{T}|P|\leq\sup_{CT}|P|\lesssim_{n,d}C^{d}\sup_{T}|P|.
Proof.

Let ll be an affine transformation that maps [1,1]n[-1,1]^{n} to TT. Apply Proposition 4.2 to PlP\circ l to get

sup[1,1]n|Pl|n,dsup[C,C]nCd|Pl|,\sup_{[-1,1]^{n}}|P\circ l|\lesssim_{n,d}\sup_{[-C,C]^{n}}C^{d}|P\circ l|,

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 n1n-1, namely, for hypersurfaces.

Let ϕ:[1,1]n1\phi:[-1,1]^{n-1}\to\mathbb{R} be a continuous function. Let Ω\Omega be a parallelogram contained in [1,1]n[-1,1]^{n} and δ>0\delta>0. We say Ω\Omega is (ϕ,δ)(\phi,\delta)-flat

  1. (1)

    in the first sense, if there exists an affine function λ:n\lambda:\mathbb{R}^{n}\to\mathbb{R} such that

    (4.1) supxΩ|ϕ(x)λ(x)|δ;\sup_{x\in\Omega}|\phi(x)-\lambda(x)|\leq\delta;
  2. (2)

    in the second sense, if ϕ\phi is C1C^{1}, and

    (4.2) supx,yΩ|ϕ(y)ϕ(x)ϕ(x)(yx)|δ;\sup_{x,y\in\Omega}|\phi(y)-\phi(x)-\nabla\phi(x)\cdot(y-x)|\leq\delta;
  3. (3)

    in the third sense, if ϕ\phi is C2C^{2}, and for each v𝕊n1v\in\mathbb{S}^{n-1} we have

    (4.3) supxΩ,v𝕊n1t:x+tvΩ|vTD2ϕ(x)v||t|2δ.\sup_{\begin{subarray}{c}x\in\Omega,v\in\mathbb{S}^{n-1}\\ t\in\mathbb{R}:x+tv\in\Omega\end{subarray}}\left|v^{T}D^{2}\phi(x)v\right||t|^{2}\leq\delta.

    Equivalently, this means that given any line segment \mathcal{L} in the direction of vv and contained Ω\Omega, we have

    (4.4) supx|vTD2ϕ(x)v||Λ|2δ.\sup_{x\in\mathcal{L}}\left|v^{T}D^{2}\phi(x)v\right||\Lambda|^{2}\leq\delta.

Note that all the above flatness are invariant under affine transformations. Namely, if L:nnL:\mathbb{R}^{n}\to\mathbb{R}^{n} is an affine transformation, then a set Ω\Omega is (ϕ,δ)(\phi,\delta)-flat if and only if LΩL\Omega is (ϕL,δ)(\phi\circ L,\delta)-flat, in all the three senses above. Thus in particular, we have

Proposition 4.4.

If ϕ\phi is δ\delta-flat over [1,1]n[-1,1]^{n} in the third sense, then we have supx[1,1]n|D2ϕ(x)|nδ\sup_{x\in[-1,1]^{n}}|D^{2}\phi(x)|\lesssim_{n}\delta.

Fix ϕ,Ω,δ\phi,\Omega,\delta. Below we denote by F1(δ),F2(δ),F3(δ)F1(\delta),F2(\delta),F3(\delta) the property that ϕ\phi is δ\delta-flat over Ω\Omega in the first, second, third sense, respectively.

Proposition 4.5.

We have F3(δ)F2(δ)F1(δ)F3(\delta)\implies F2(\delta)\implies F1(\delta). Conversely, if ϕ\phi is a polynomial of degree dd and Ω\Omega is a parallelogram, then we also have F1(δ)F3(Cd,nδ)F1(\delta)\implies F3(C_{d,n}\delta), where Cd,nC_{d,n} is a constant depending only on d,nd,n. However, F1(δ)F2(Cδ)F1(\delta)\implies F2(C\delta) and F2(δ)F3(Cδ)F2(\delta)\implies F3(C\delta) are both false, for any fixed constant C=CnC=C_{n} depending on nn only.

Proof.

F3(δ)F2(δ)F1(δ)F3(\delta)\implies F2(\delta)\implies F1(\delta) is trivial. For the implication F1(δ)F3(Cδ)F1(\delta)\implies F3(C\delta), we let Λ:[1,1]nΩ\Lambda:[-1,1]^{n}\to\Omega be an affine bijection. Then by the assumption,

(4.5) supx[1,1]n|ϕ(Λ(x))λ(Λ(x))|d,nδ.\sup_{x\in[-1,1]^{n}}|\phi(\Lambda(x))-\lambda(\Lambda(x))|\lesssim_{d,n}\delta.

Using Proposition 4.2, we see all coefficients of ϕ(Λ(x))λ(Λ(x))\phi(\Lambda(x))-\lambda(\Lambda(x)) are Od,n(δ)O_{d,n}(\delta). Thus, for any direction v𝕊n1v\in\mathbb{S}^{n-1}, we have

supx[1,1]n|vTD2(ϕΛ)(x)v|d,nδ.\sup_{x\in[-1,1]^{n}}|v^{T}D^{2}(\phi\circ\Lambda)(x)v|\lesssim_{d,n}\delta.

Rescaling back, this gives F3(Cd,nδ)F3(C_{d,n}\delta).

Lastly, using the simple example ϕ(t)=|t|2d\phi(t)=|t|^{2d} when dd can be arbitrarily large, we observe that F1(δ)F2(Cδ)F1(\delta)\implies F2(C\delta) and F2(δ)F3(Cδ)F2(\delta)\implies F3(C\delta) are both false, for any fixed constant C=CnC=C_{n} depending on nn only. ∎

4.4. Flatness of parametrised surfaces

Definition 4.6.

We say a subset Ωn\Omega\subseteq\mathbb{R}^{n} is δ\delta-flat if there exists a hyperplane HH such that ΩNδ(H)\Omega\subseteq N_{\delta}(H).

Proposition 4.7.

Let r,ψr,\psi be continuous functions, and let δ>0\delta>0. Assume SS is (r,δ)(r,\delta)-flat in sense of (4.1). Assume also TT is (ψ,δ)(\psi,\delta)-flat in sense of (4.1). Then the subsets

{(s,r(s)t,r(s)ψ(t):sS,tT}\displaystyle\{(s,r(s)t,r(s)\psi(t):s\in S,t\in T\}
{(s,t,r(s)+ψ(t)):sS,tT}\displaystyle\{(s,t,r(s)+\psi(t)):s\in S,t\in T\}

are O(δ)O(\delta)-flat in the sense of Definition 4.6. Here the implicit constant depends only on the sup norms of r,ψr,\psi.

Proof.

By flatness in sense of (4.1), we may assume r,ψr,\psi are affine functions. For the first subset, we need to find a hyperplane HH such that

{(s,r(s)t,r(s)ψ(t):sS,tT}NO(δ)(H).\{(s,r(s)t,r(s)\psi(t):s\in S,t\in T\}\subseteq N_{O(\delta)}(H).

Write ψ(t)=At+b\psi(t)=At+b for a matrix AA and a vector bb. Defining t=r(s)tt^{\prime}=r(s)t, we compute directly that

r(s)ψ(tr(s))=At+r(s)b,r(s)\psi\left(\frac{t^{\prime}}{r(s)}\right)=At^{\prime}+r(s)b,

which is an affine function in s,ts,t^{\prime}. Thus we may just take the hyperplane HH to be parametrised by

(4.6) (s,t,At+r(s)b).(s,t^{\prime},At^{\prime}+r(s)b).

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 PP be a bivariate polynomial of degree at most dd and coefficients bounded by 11. Let R[1,1]2R\subseteq[-1,1]^{2} be a parallelogram, and suppose for some σ(0,1]\sigma\in(0,1] we have |detD2P|σ|\det D^{2}P|\leq\sigma on RR. Then there exists another bivariate polynomial QQ of degree at most dd with coefficients bounded by Od(1)O_{d}(1), such that detD2Q0\det D^{2}Q\equiv 0, and PQL(R)dσβ\left\lVert P-Q\right\rVert_{L^{\infty}(R)}\lesssim_{d}\sigma^{\beta}, where β=β(d)(0,1]\beta=\beta(d)\in(0,1]. Moreover, uQc\nabla_{u}Q\equiv\vec{c} for some |u|=1|u|=1.

The case R=[1,1]2R=[-1,1]^{2} is equivalent to the statement of Theorem 3.2 of [LY23].

Proof.

Let α=α(d)(0,1]\alpha=\alpha(d)\in(0,1] be the exponent given in Theorem 3.2 of [LY23]. We consider two cases according to whether |R|σαd|R|\leq\sigma^{\frac{\alpha}{d}} or not.

If |R|σαd|R|\leq\sigma^{\frac{\alpha}{d}}, then the shorter dimension of RR is σα2d\leq\sigma^{\frac{\alpha}{2d}}. Let LL be the unbounded straight line along one of the longer edges of RR, and let πL\pi_{L} be the orthogonal projection from 2\mathbb{R}^{2} onto LL. Then we simply define Q(x)=P(πLx)Q(x)=P(\pi_{L}x).

Then detD2Q0\det D^{2}Q\equiv 0, QQ has all coefficients bounded by Od(1)O_{d}(1), and PQL(R)dσα2d\left\lVert P-Q\right\rVert_{L^{\infty}(R)}\lesssim_{d}\sigma^{\frac{\alpha}{2d}}.

If |R|>σαd|R|>\sigma^{\frac{\alpha}{d}}, then both dimensions of RR must be at least σαd/10\sigma^{\frac{\alpha}{d}}/10. Let Λ:[1,1]2R\Lambda:[-1,1]^{2}\to R be an affine bijection. With this, denote

PR(x):=P(Λx)P(Λ0)P(Λ0)x.P_{R}(x):=P(\Lambda x)-P(\Lambda 0)-\nabla P(\Lambda 0)\cdot x.

Then PRP_{R} has degree at most dd with coefficients bounded by Od(1)O_{d}(1), and |detD2PR|σ|\det D^{2}P_{R}|\leq\sigma on [1,1]2[-1,1]^{2}. Apply Theorem 3.2 of [LY23] to find a rotation ρ:22\rho:\mathbb{R}^{2}\to\mathbb{R}^{2} such that

(PRρ)(x)=A(πx)+σαB(x1,x2),(P_{R}\circ\rho)(x)=A(\pi x)+\sigma^{\alpha}B(x_{1},x_{2}),

where πx:=x1\pi x:=x_{1} and A,BA,B are polynomials of degree at most dd and coefficients bounded by Od(1)O_{d}(1). This means PR(x)=A(ρ1πx)+σαB(ρ1x)P_{R}(x)=A(\rho^{-1}\pi x)+\sigma^{\alpha}B(\rho^{-1}x), and so

P(x)\displaystyle P(x) =PR(Λ1x)+P(Λ0)+P(Λ0)(Λ1x)\displaystyle=P_{R}(\Lambda^{-1}x)+P(\Lambda 0)+\nabla P(\Lambda 0)\cdot(\Lambda^{-1}x)
=A(ρ1πΛ1x)+σαB(ρ1Λ1x)+P(Λ0)+P(Λ0)(Λ1x).\displaystyle=A(\rho^{-1}\pi\Lambda^{-1}x)+\sigma^{\alpha}B(\rho^{-1}\Lambda^{-1}x)+P(\Lambda 0)+\nabla P(\Lambda 0)\cdot(\Lambda^{-1}x).

Let

Q(x)=A(ρ1πΛ1x)+P(Λ0)+P(Λ0)(Λ1x).Q(x)=A(\rho^{-1}\pi\Lambda^{-1}x)+P(\Lambda 0)+\nabla P(\Lambda 0)\cdot(\Lambda^{-1}x).

Then QQ is a polynomial of degree at most dd, detD2Q0\det D^{2}Q\equiv 0, and PQL(R)dσα\left\lVert P-Q\right\rVert_{L^{\infty}(R)}\lesssim_{d}\sigma^{\alpha}. It remains to show that all coefficients of QQ are bounded by Od(1)O_{d}(1). But since PP has coefficients bounded by 11, it remains to show all coefficients of σαB(ρ1Λ1x)\sigma^{\alpha}B(\rho^{-1}\Lambda^{-1}x) are bounded above by Od(1)O_{d}(1). But this follows from the assumption that both dimensions of RR are at least σαd/10\sigma^{\frac{\alpha}{d}}/10.

Thus, we see that taking β=α2d\beta=\frac{\alpha}{2d} suffices. ∎

4.6. Combining decoupling inequalities

We provide a proof of how we can combine different q(Lp)\ell^{q}(L^{p}) 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 𝒯\mathcal{T} be a partition of a parallelogram P0P_{0} by parallelograms TT, and for each TT, let 𝒮(T)\mathcal{S}(T) be a partition of TT by parallelograms SS. Denote 𝒮=T𝒯𝒮(T)\mathcal{S}=\cup_{T\in\mathcal{T}}\mathcal{S}(T), which is a partition of P0P_{0} at a finer level. Then for all p,q,αp,q,\alpha we have

Dec(P0,𝒮,p,q,α)log(#𝒮)Dec(P0,𝒯,p,q,α)supT𝒯Dec(T,𝒮(T),p,q,α).\mathrm{Dec}(P_{0},\mathcal{S},p,q,\alpha)\lesssim\log(\#\mathcal{S})\mathrm{Dec}(P_{0},\mathcal{T},p,q,\alpha)\sup_{T\in\mathcal{T}}\mathrm{Dec}(T,\mathcal{S}(T),p,q,\alpha).
Proof.

For each k1k\geq 1, denote

𝒯k:={T𝒯:#𝒮(T)[2k1,2k)}.\mathcal{T}_{k}:=\{T\in\mathcal{T}:\#\mathcal{S}(T)\in[2^{k-1},2^{k})\}.

Then there are O(log(#𝒮))O(\log(\#\mathcal{S})) many such kk’s. Given test functions fSf_{S} each Fourier supported on SS, and denote fT:=S𝒮(T)fSf_{T}:=\sum_{S\in\mathcal{S}(T)}f_{S}, f:=T𝒯fTf:=\sum_{T\in\mathcal{T}}f_{T}. Then we have

fp\displaystyle\left\lVert f\right\rVert_{p} kfkp\displaystyle\leq\sum_{k}\left\lVert f_{k}\right\rVert_{p}
kDec(P0,𝒯,p,q,α)(#𝒯k)αfTpq(T𝒯k)\displaystyle\leq\sum_{k}\mathrm{Dec}(P_{0},\mathcal{T},p,q,\alpha)(\#\mathcal{T}_{k})^{\alpha}\left\lVert\left\lVert f_{T}\right\rVert_{p}\right\rVert_{\ell^{q}(T\in\mathcal{T}_{k})}
kDec(P0,𝒯,p,q,α)(#𝒯k)α2kαsupT𝒯Dec(T,𝒮(T),p,q,α)fSpq(S𝒮)\displaystyle\leq\sum_{k}\mathrm{Dec}(P_{0},\mathcal{T},p,q,\alpha)(\#\mathcal{T}_{k})^{\alpha}2^{k\alpha}\sup_{T\in\mathcal{T}}\mathrm{Dec}(T,\mathcal{S}(T),p,q,\alpha)\left\lVert\left\lVert f_{S}\right\rVert_{p}\right\rVert_{\ell^{q}(S\in\mathcal{S})}
kDec(P0,𝒯,p,q,α)(#𝒮)αsupT𝒯Dec(T,𝒮(T),p,q,α)fSpq(S𝒮)\displaystyle\leq\sum_{k}\mathrm{Dec}(P_{0},\mathcal{T},p,q,\alpha)(\#\mathcal{S})^{\alpha}\sup_{T\in\mathcal{T}}\mathrm{Dec}(T,\mathcal{S}(T),p,q,\alpha)\left\lVert\left\lVert f_{S}\right\rVert_{p}\right\rVert_{\ell^{q}(S\in\mathcal{S})}
log(#𝒮)Dec(P0,𝒯,p,q,α)(#𝒮)αsupT𝒯Dec(T,𝒮(T),p,q,α)fSpq(S𝒮),\displaystyle\lesssim\log(\#\mathcal{S})\mathrm{Dec}(P_{0},\mathcal{T},p,q,\alpha)(\#\mathcal{S})^{\alpha}\sup_{T\in\mathcal{T}}\mathrm{Dec}(T,\mathcal{S}(T),p,q,\alpha)\left\lVert\left\lVert f_{S}\right\rVert_{p}\right\rVert_{\ell^{q}(S\in\mathcal{S})},

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 l2l^{2} decoupling conjecture. Ann. of Math. (2), 182(1):351–389, 2015.
  • [BD16] J. Bourgain and C. Demeter. Decouplings for surfaces in 4\mathbb{R}^{4}. 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. 2\ell^{2} decoupling in 2\mathbb{R}^{2} 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 5\mathbb{R}^{5}. 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 3\mathbb{R}^{3}. preprint, 2022. arXiv:2207.13844.
  • [GLYZK21] S. Guo, Z. K. Li, P.-L. Yung, and P. Zorin-Kranich. A short proof of 2\ell^{2} 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. l2l^{2} decoupling theorem for surfaces in 3\mathbb{R}^{3}. 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 3\mathbb{R}^{3}. Math. Ann., 383(3-4):1319–1351, 2022.
  • [LY23] J. Li and T. Yang. Decoupling for smooth surfaces in 3\mathbb{R}^{3}. Amer. J. Math., 2023. accepted.
  • [Oh18] C. Oh. Decouplings for three-dimensional surfaces in 6\mathbb{R}^{6}. Math. Z., 290(1-2):389–419, 2018.
  • [PS07] M. Pramanik and A. Seeger. LpL^{p} 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 LpL^{p} for large pp. Geom. Funct. Anal., 10(5):1237–1288, 2000.
  • [Yan21] T. Yang. Uniform l2l^{2}-Decoupling in 2\mathbb{R}^{2} for polynomials. J. Geom. Anal., 31(11):10846–10867, 2021.