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

Fourier dimension of constant rank hypersurfaces

Junjie Zhu
Abstract.

Any hypersurface in d+1\mathbb{R}^{d+1} has a Hausdorff dimension of dd. However, the Fourier dimension depends on the finer geometric properties of the hypersurface. For example, the Fourier dimension of a hyperplane is 0, and the Fourier dimension of a hypersurface with non-vanishing Gaussian curvature is dd. Recently, Harris showed that the Euclidean light cone in d+1\mathbb{R}^{d+1} has a Fourier dimension of d1d-1, which leads one to conjecture that the Fourier dimension of a hypersurface equals the number of non-vanishing principal curvatures. We prove this conjecture for all constant rank hypersurfaces. Our method involves substantial generalizations of Harris’s strategy.

Key words and phrases:
Fourier dimension, Hausdorff dimension, Constant rank hypersurfaces, Oscillatory integrals, Stationary phase
2020 Mathematics Subject Classification:
42B10, 42B20, 28A12, 53A07.

1. Introduction

The decay rate of the Fourier transforms of measures supported on manifolds has been of central interest in harmonic analysis. The Fourier transform of the normalized surface measure σ\sigma on the unit sphere 𝕊dd+1\mathbb{S}^{d}\subset\mathbb{R}^{d+1} is given by

σ^(ξ)=cd|ξ|1d2Jd12(2π|ξ|),\widehat{\sigma}(\xi)=c_{d}|\xi|^{\frac{1-d}{2}}J_{\frac{d-1}{2}}(2\pi|\xi|),

where Jd12J_{\frac{d-1}{2}} is the Bessel function of integral order d12\frac{d-1}{2}, and cdc_{d} is the normalizing constant [18]. It satisfies

(1) |σ^(ξ)|Cd|ξ|d2 for a Cd>0.|\widehat{\sigma}(\xi)|\leq C_{d}|\xi|^{-\frac{d}{2}}\text{ for a }C_{d}>0.

Describing sets in d+1\mathbb{R}^{d+1}, which may not be manifolds, using the Fourier decay of measures supported on them is one motivation for studying the Fourier dimension. Let (A)\mathcal{M}(A) be the set of measures μ\mu supported on AA with finite total mass μ1:=μ(A)<\left\lVert\mu\right\rVert_{1}:=\mu(A)<\infty. The Fourier transform of a measure μ\mu at ξn\xi\in\mathbb{R}^{n} is defined as μ^(ξ):=e2πixξ𝑑μ(x)\widehat{\mu}(\xi):=\int e^{-2\pi ix\xi}d\mu(x). The Fourier dimension for a Borel Ad+1A\subset\mathbb{R}^{d+1} is defined as

dimF(A):=sup{s[0,d+1]:μ(A),supξd+1|ξ|s2|μ^(ξ)|<}.\dim_{F}(A):=\sup\{s\in[0,d+1]:\exists\mu\in\mathcal{M}(A),\sup_{\xi\in\mathbb{R}^{d+1}}|\xi|^{\frac{s}{2}}|\widehat{\mu}(\xi)|<\infty\}.

The notion of the Fourier dimension is ubiquitous in harmonic analysis and geometric measure theory. It provides a lower bound on the Hausdorff dimension, which is discussed further in Section 4.1. Studies of Fourier dimensions include probabilistic [17, 10, 1, 12] and deterministic constructions [11, 6] of Salem sets with equal Fourier and Hausdorff dimensions and properties satisfied by any Borel sets of large Fourier dimensions [13].

The objective of this paper is to study the case where AA is a constant rank hypersurface and compute its Fourier dimension.

1.1. Constant rank hypersurfaces

Let Md+1M\subset\mathbb{R}^{d+1} be an orientable smooth hypersurface, which is a topological manifold of dimension dd with a normal direction N:M𝕊dN:M\to\mathbb{S}^{d}. A way to describe MM is through notions of curvatures, defined through the eigenvalues of the Weingarten map [19]. Let TpMT_{p}M be the tangent space of MM at pMp\in M. The Weingarten map Lp:TpMTpML_{p}:T_{p}M\to T_{p}M at pMp\in M is the linear map Lp(v)=DvN=ddt(Nγ)(0)L_{p}(v)=-D_{v}N=-\frac{d}{dt}(N\circ\gamma)(0), where γ:IM\gamma:I\to M is a curve with γ(0)=p\gamma(0)=p, γ(0)=v\gamma^{\prime}(0)=v. The principal curvatures of MM at pp are the eigenvalues of the map LpL_{p}, and the Gaussian curvature of MM at pp is the product of the eigenvalues, which equals the determinant of LpL_{p}. We note that Gaussian and principal curvatures at pMp\in M are independent of the parametrization of MM and the choice of a basis for TpMT_{p}M. The Weingarten map is self-adjoint, so the eigenvalues of LpL_{p} are real.

In this manuscript, we focus on hypersurfaces of constant rank kk where at all points pMp\in M, kk principal curvatures are non-zero, and dkd-k principal curvatures are zero. For example, in addition to cones and cylinders, tangent surfaces are hypersurfaces of constant rank 11 in 3\mathbb{R}^{3}. Two tangent surfaces are shown in Section 2. A hypersurface in higher dimensions not classified as a cone or a cylinder is

{γ(t)+j=1d1vjdjγdtj(t)|t,vj}d+1,\left\{\left.\gamma(t)+\sum_{j=1}^{d-1}v_{j}\frac{d^{j}\gamma}{dt^{j}}(t)\right\rvert t,v_{j}\in\mathbb{R}\right\}\subset\mathbb{R}^{d+1},

where γ(t)=(t,t2,,td,td+1)\gamma(t)=(t,t^{2},\cdots,t^{d},t^{d+1}). It is of constant rank 11 that generalizes tangent surfaces in 3\mathbb{R}^{3} to higher dimensions. Another hypersurface of constant rank 22 in 4\mathbb{R}^{4} is

{(t,sin(t)es,sin(2t)e4s,sin(3t)e9s)+v(1,cos(t)es,2cos(2t)e4s,3cos(3t)e9s)|t,s,v}.\left\{(t,\sin(t)e^{-s},\sin(2t)e^{-4s},\sin(3t)e^{-9s})+v(1,\cos(t)e^{-s},2\cos(2t)e^{-4s},3\cos(3t)e^{-9s})|t,s,v\in\mathbb{R}\right\}.

Some more examples are cylinders of cones {(kv,k,h)|k,h,vS}d+1\{(kv,k,h)|k,h\in\mathbb{R},v\in S\}\subset\mathbb{R}^{d+1}, where SS is a hypersurface in d1\mathbb{R}^{d-1} with non-vanishing Gaussian curvature.

Although surfaces of constant rank have been studied extensively in differential geometry [20], they are less examined in geometric measure theory, where geometric objects are analyzed via measures supported on them.

1.2. Main Result

This manuscript focuses on computing dimF(M)\dim_{F}(M), interpreting the properties of MM that dimF(M)\dim_{F}(M) captures, and differentiating between topological, Hausdorff, and Fourier dimensions on MM. Since the Hausdorff dimension is preserved under diffeomorphisms, which are bi-Lipschitz, dimH(M)=d\dim_{H}(M)=d [4, Proposition 3.1].

By (1), the unit sphere 𝕊d\mathbb{S}^{d} has Fourier dimension dd. Hlawka [9] showed that hypersurfaces with non-vanishing Gaussian curvature also have a Fourier dimension of dd, which suggests that the Fourier dimension does not depend on the symmetry of the hypersurface. Harris [5] showed that the Euclidean light cone in d+1\mathbb{R}^{d+1}, which is a hypersurface with zero Gaussian curvature but of constant rank d1d-1, has a Fourier dimension of d1d-1. The proof relies on the rotational symmetry of the light cone. The author’s previous work [21] adapts the proof of [5] and shows that cones and cylinders of rank d1d-1 without rotational symmetry also have a Fourier dimension of d1d-1. Table 1 summarizes these results.

Hypersurface MM Constant Rank dimF(M)\dim_{F}(M) Reference
non-vanishing Gaussian curvature (𝕊d\mathbb{S}^{d}, etc) dd dd [9]
Light cone {h(x,1)|x𝕊d1,h}\{h(x,1)|x\in\mathbb{S}^{d-1},h\in\mathbb{R}\} d1d-1 d1d-1 [5]
Generalized cone {h(x,1)|xS,h}\{h(x,1)|x\in S,h\in\mathbb{R}\} d1d-1 d1d-1 [21]
(d1)(d-1)-cylinder S×S\times\mathbb{R} d1d-1 d1d-1 [21]
Hyperplane {x:xd+1=0}\{x:x_{d+1}=0\} 0 0
Table 1. Fourier dimension of some hypersurfaces, where SdS\subset\mathbb{R}^{d} is a hypersurface with non-zero Gaussian curvature.

Previous results suggest that the Fourier dimension of MM equals its rank. In this manuscript, we confirm that such a hypothesis holds

Theorem 1.1.

Let MM be a smooth hypersurface of constant rank kk in d+1\mathbb{R}^{d+1}, then dimF(M)=k\dim_{F}(M)=k.

Remark.
  1. a.

    Theorem 1.1 shows a direct link between the principal curvatures and the Fourier dimension for constant rank hypersurfaces. We can view the Fourier dimension as a generalization of the number of non-zero principal curvatures defined on hypersurfaces. It is a generalization of the results in [5, 21]. It includes all hypersurfaces of constant rank kk for k<d1k<d-1 and hypersurfaces of constant rank d1d-1 that are neither cones nor cylinders.

  2. b.

    Previously known bounds on the Fourier dimension of a constant rank kk hypersurface Md+1M\subset\mathbb{R}^{d+1} are

    kdimF(M)k(d+1k+1).k\leq\dim_{F}(M)\leq k\left(\frac{d+1}{k+1}\right).

    The lower bound from [14] arises from the Fourier transform of measures supported on the hypersurface induced by the Lebesgue measure. The upper bound is obtained from the study of the pp-thin problem [7]. If μ(M)\mu\in\mathcal{M}(M) satisfies |μ^(ξ)|C(1+|ξ|)α2|\widehat{\mu}(\xi)|\leq C(1+|\xi|)^{-\frac{\alpha}{2}} for C,α>0C,\alpha>0, μ^Lp(d+1)\widehat{\mu}\in L^{p}(\mathbb{R}^{d+1}) for all p>2(d+1)αp>\frac{2(d+1)}{\alpha}. Such pp must satisfy p>2(k+1)kp>\frac{2(k+1)}{k}, so αk(d+1k+1)\alpha\leq k\left(\frac{d+1}{k+1}\right).

1.3. Overview of the proof

To establish Theorem 1.1, we need to show that the Fourier dimension of a hypersurface MM of constant rank kk is bounded below and above by kk. For the upper bound (Proposition 1.2), the main strategy adapted from [5] is to create a new measure by averaging a series of push-forward of an old measure supported on MM.

An adaptation of the methodology from [5] is needed. In previous works [5, 21], it was believed that the specific maps used to push-forward measures are required to map MM to itself, so the new measure is still supported on the hypersurface. While such assumptions pose no issue for cones and cylinders, similar maps may not exist for more general hypersurfaces for us to apply the same machinery.

Fortunately, we note that the new measure does not have to be supported on the original MM. We propose new maps that push forward an old measure to a series of similar hypersurfaces, and we still create a new measure as the average of the series of measures. The proof in Section 3 shows that such a new measure may not be supported on MM, but it has a Fourier decay exponent in one direction that is not less than that of the old measure. Moreover, the Fourier transform of the new measure is related to the second fundamental form of the hypersurface MM, which prevents the Fourier decay exponent of the new measure from being greater than k/2k/2.

1.4. Proof of the main result and organization of the manuscript

Theorem 1.1 follows from Theorem A and Proposition 1.2.

Theorem A ([14]).

Suppose that at least kk of the principal curvatures are not zero at all points in MM. Then, there exists a measure μ(M)\mu\in\mathcal{M}(M) such that supξd+1|ξ|k2|μ^(ξ)|<\sup_{\xi\in\mathbb{R}^{d+1}}|\xi|^{\frac{k}{2}}|\widehat{\mu}(\xi)|<\infty.

Therefore, if MM is a constant rank kk hypersurface, dimF(M)k\dim_{F}(M)\geq k. It remains to prove the following.

Proposition 1.2.

dimF(M)k\dim_{F}(M)\leq k.

We present examples of Proposition 1.2 in Section 2 and prove Proposition 1.2 in Section 3.

1.5. Notations

For two functions f,g:D0f,g:D\to\mathbb{R}_{\geq 0} with a domain DD, we write fgf\lesssim g to denote that there exists a c>0c>0, such that for all xDx\in D, f(x)cg(x)f(x)\leq cg(x).

For ϕ:nn\phi:\mathbb{R}^{n}\to\mathbb{R}^{n}, we denote the Jacobian of ϕ\phi as 𝐉ϕ{\bf J}\phi.

If (X,𝒳)(X,\mathcal{X}) and (Y,𝒴)(Y,\mathcal{Y}) are measurable spaces and the mapping F:XYF:X\to Y is (𝒳,𝒴)(\mathcal{X},\mathcal{Y})-measurable, the push-forward map F#F^{\#} is defined to be F#μ(U)=μ(F1(U))F^{\#}\mu(U)=\mu(F^{-1}(U)) for any set U𝒴U\in\mathcal{Y} and any measure μ\mu on (X,𝒳)(X,\mathcal{X}).

On (X,𝒳)(X,\mathcal{X}), if ff is a measurable function and μ\mu is a measure, we denote f,μ=Xf𝑑μ\langle f,\mu\rangle=\int_{X}fd\mu.

1.6. Acknowledgments

I thank Malabika Pramanik and Jialing Zhang for their discussions on this project. I thank Yuveshen Moorogen for editing suggestions for an earlier version of the manuscript.

2. Upper bounds on the Fourier dimension: Examples

In this section, we demonstrate the process of showing that the Fourier dimensions of two tangent surfaces, which are hypersurfaces of constant rank 11 in 3\mathbb{R}^{3} different from generalized cones and cylinders, are at most 1. We note that the normal vectors used in the section are not unit vectors, but the conclusion is not affected.

2.1. Tangent surface generated by a Helix

Let γ:3\gamma:\mathbb{R}\to\mathbb{R}^{3}, γ(t)=(t,t2,t3)\gamma(t)=(t,t^{2},t^{3}), and

(2) S={Φ(t,v):=γ(t)+vγ(t)|t,v0}S=\{\Phi(t,v):=\gamma(t)+v\gamma^{\prime}(t)|t\in\mathbb{R},v\neq 0\}

be the tangent surface generated by γ\gamma, where γ(t)=(1,2t,3t2)\gamma^{\prime}(t)=(1,2t,3t^{2}). γ′′(t)=(0,2,6t)\gamma^{\prime\prime}(t)=(0,2,6t), and a normal vector of SS at a point Φ(t,v)\Phi(t,v) is

(3) n(t):=(3t2,3t,1).\overrightarrow{n}(t):=(3t^{2},-3t,1).

The tangent surface SS is of rank 1 since the second fundamental form at Φ(t,v)\Phi(t,v) is

(ΦttnΦtvnΦvtnΦvvn)=(6v000)\begin{pmatrix}\Phi_{tt}\cdot\overrightarrow{n}&\Phi_{tv}\cdot\overrightarrow{n}\\ \Phi_{vt}\cdot\overrightarrow{n}&\Phi_{vv}\cdot\overrightarrow{n}\end{pmatrix}=\begin{pmatrix}6v&0\\ 0&0\end{pmatrix}

has rank 11 when v0v\neq 0.

Refer to caption
Figure 1. A 3D plot of SS from Matlab.
Proposition 2.1.

dimF(S)1\dim_{F}(S)\leq 1.

Proof.

Let μ(S)\mu\in\mathcal{M}(S). Suppose that there exists β>0\beta>0, such that |μ^(ξ)|C|ξ|β2|\widehat{\mu}(\xi)|\leq C|\xi|^{-\frac{\beta}{2}} for ξ3\xi\in\mathbb{R}^{3}, C>0C>0. By Lemma 4.2, we may assume that S={Φ(t,v)||t|c,v[a,b]}S=\{\Phi(t,v)||t|\leq c,v\in[a,b]\} for 0<a<b0<a<b, c>0c>0.

We construct new measures. Let the reparametrization Ts:SST_{s}:S\to S be

(4) Ts(Φ(t,v))=Φ(ts,v).T_{s}(\Phi(t,v))=\Phi(t-s,v).
Φ(t,v)\Phi(t,v)Φ(ts,v)\Phi(t-s,v) Re-parametrize by TsT_{s}
Figure 2. Illustration of the re-parametrization by TsT_{s} on SS.

Let μs(S)\mu_{s}\in\mathcal{M}(S) be

(5) μs=Ts#μ.\mu_{s}=T_{s}^{\#}\mu.

Let ψC0()\psi\in C_{0}^{\infty}(\mathbb{R}) be non-negative bump function with ψ(s)=1\psi(s)=1 when |s|c|s|\leq c, ψ(s)=0\psi(s)=0 when |s|2c|s|\geq 2c. Then we let the average measure ν\nu be defined as

(6) f𝑑ν=f,μsψ(s)𝑑s\int fd\nu=\int\langle f,\mu_{s}\rangle\psi(s)ds

for non-negative Borel functions ff.

Now, we examine ν^\widehat{\nu} in the e3e_{3} direction, where e3e_{3} is also a normal vector at γ(0,v)\gamma(0,v). Proposition 2.1 is a consequence of the following two lemmas.

Lemma 2.2.

For ρ>0\rho>0, |ν^(ρe3)|ρβ2|\widehat{\nu}(\rho e_{3})|\lesssim\rho^{-\frac{\beta}{2}}.

Lemma 2.3.

There exists ρ0>0\rho_{0}>0, such that for ρ>ρ0\rho>\rho_{0}, |ν^(ρe3)|ρ12|\widehat{\nu}(\rho e_{3})|\gtrsim\rho^{-\frac{1}{2}}. ∎

Proof of Lemma 2.2.

First, we show that for each ss\in\mathbb{R}, |μs^(ρe3)|ρβ2|\widehat{\mu_{s}}(\rho e_{3})|\lesssim\rho^{-\frac{\beta}{2}}. By unwinding the definitions of push-forward measures, Fourier transform, and integration on the manifold, we have

(7) μs^(ρe3)\displaystyle\widehat{\mu_{s}}(\rho e_{3}) =e2πixρe3𝑑Ts#μ(x)\displaystyle=\int e^{-2\pi ix\cdot\rho e_{3}}dT_{s}^{\#}\mu(x) by (5)
=e2πiρTs(Φ(t,v))e3𝑑μΦ(t,v)\displaystyle=\int e^{-2\pi i\rho T_{s}(\Phi(t,v))\cdot e_{3}}d\mu\circ\Phi(t,v)
=e2πiρΦ(ts,v)e3𝑑μΦ(t,v)\displaystyle=\int e^{-2\pi i\rho\Phi(t-s,v)\cdot e_{3}}d\mu\circ\Phi(t,v) by (4).\displaystyle\text{ by (\ref{eq31_TS})}.

From (2) and (3),

(8) Φ(ts,v)e3=Φ(t,v)n(s)s3.\Phi(t-s,v)\cdot e_{3}=\Phi(t,v)\cdot\overrightarrow{n}(s)-s^{3}.

Then, (7) becomes

((7) c.) μs^(ρe3)\displaystyle\widehat{\mu_{s}}(\rho e_{3}) =e2πiρΦ(ts,v)e3𝑑μΦ(t,v)\displaystyle=\int e^{-2\pi i\rho\Phi(t-s,v)\cdot e_{3}}d\mu\circ\Phi(t,v)
=e2πiρ[Φ(t,v)n(s)s3]𝑑μΦ(t,v)\displaystyle=\int e^{-2\pi i\rho[\Phi(t,v)\cdot\overrightarrow{n}(s)-s^{3}]}d\mu\circ\Phi(t,v) by (8)
=eπiρs3μ^(ρn(s)).\displaystyle=e^{\pi i\rho s^{3}}\widehat{\mu}(\rho\overrightarrow{n}(s)).

Since |n(s)|1|\overrightarrow{n}(s)|\geq 1, |μs^(ρe3)|=|μ^(ρn(s))|C|ρn(s)|β2Cρβ2.|\widehat{\mu_{s}}(\rho e_{3})|=|\widehat{\mu}(\rho\overrightarrow{n}(s))|\leq C|\rho\overrightarrow{n}(s)|^{-\frac{\beta}{2}}\leq C\rho^{-\frac{\beta}{2}}. Then,

|ν^(ρe3)|\displaystyle|\widehat{\nu}(\rho e_{3})|\leq |μs^(ρn(s))|ψ(s)𝑑s\displaystyle\int|\widehat{\mu_{s}}(\rho\overrightarrow{n}(s))|\psi(s)ds by (6)
\displaystyle\leq CψL1ρβ2.\displaystyle C\left\lVert\psi\right\rVert_{L^{1}}\rho^{-\frac{\beta}{2}}.
Proof of Lemma 2.3.

An expression of ν^(ρe3)\widehat{\nu}(\rho e_{3}) is given by

(9) ν^(ρe3)\displaystyle\widehat{\nu}(\rho e_{3}) =e2πiρΦ(ts,v)e3𝑑μΦ(t,v)ψ(s)𝑑s\displaystyle=\int\int e^{-2\pi i\rho\Phi(t-s,v)\cdot e_{3}}d\mu\circ\Phi(t,v)\psi(s)ds by (6)
=e2πiρΦ(ts,v)e3ψ(s)𝑑s𝑑μΦ(t,v).\displaystyle=\int\int e^{-2\pi i\rho\Phi(t-s,v)\cdot e_{3}}\psi(s)dsd\mu\circ\Phi(t,v).

We apply the stationary phase method to study the inner integral

I(ρ;t,v):=e2πiρΦ(ts,v)e3ψ(s)𝑑s.I(\rho;t,v):=\int e^{-2\pi i\rho\Phi(t-s,v)\cdot e_{3}}\psi(s)ds.

Fixing t,vt,v, we define the phase function ϕ\phi of I(ρ;t,v)I(\rho;t,v) as

ϕ(s):=Φ(ts,v)e3=(ts)3+3v(ts)2 from (8).\phi(s):=\Phi(t-s,v)\cdot e_{3}=(t-s)^{3}+3v(t-s)^{2}\text{ from }(\ref{eq_helix_Phie3}).

Its derivatives are

ϕ(s)=3(ts)26v(ts),\phi^{\prime}(s)=-3(t-s)^{2}-6v(t-s),

and

ϕ′′(s)=6(ts)+6v.\phi^{\prime\prime}(s)=6(t-s)+6v.

The solutions to ϕ(s)=0\phi^{\prime}(s)=0 are

s=t,t+2v.s=t,t+2v.

Since v[a,b]v\in[a,b], if 2c+2a>2c-2c+2a>2c or 2c<a2c<a, the only critical point in spt ψ\text{spt }\psi is s=ts=t. Note that

ϕ(t)=0,ϕ(t)=0,ϕ′′(t)=6v.\phi(t)=0,\phi^{\prime}(t)=0,\phi^{\prime\prime}(t)=6v.

Then, by the stationary phase (part b, Theorem 4.6), there exists D0>0D_{0}>0 independent of ρ\rho, tt, and vv, such that when ρv\rho v is sufficiently large,

|I(ρ;t,v)(i6ρv)12ψ(t)|D0(ρv)1.\left|I(\rho;t,v)-\left(\frac{-i}{6\rho v}\right)^{\frac{1}{2}}\psi(t)\right|\leq D_{0}(\rho v)^{-1}.

Since ψ(t)=1\psi(t)=1 for |t|c|t|\leq c, from (9),

|ν^(ρe3)(i6ρv)12𝑑μΦ(t,v)|D0(ρv)1𝑑μΦ(t,v).\begin{split}\left|\widehat{\nu}(\rho e_{3})-\int\left(\frac{-i}{6\rho v}\right)^{\frac{1}{2}}d\mu\circ\Phi(t,v)\right|&\leq\int D_{0}(\rho v)^{-1}d\mu\circ\Phi(t,v).\\ \end{split}

As v[a,b]v\in[a,b],

|(i6ρv)12𝑑μΦ(t,v)|\displaystyle\left|\int\left(\frac{-i}{6\rho v}\right)^{\frac{1}{2}}d\mu\circ\Phi(t,v)\right|\geq 612b12ρ12,\displaystyle 6^{-\frac{1}{2}}b^{-\frac{1}{2}}\rho^{-\frac{1}{2}},
|D0(ρv)1𝑑μΦ(t,v)|\displaystyle\left|\int D_{0}(\rho v)^{-1}d\mu\circ\Phi(t,v)\right|\leq D0a1ρ1.\displaystyle D_{0}a^{-1}\rho^{-1}.

Therefore, |ν^(ρe3)||\widehat{\nu}(\rho e_{3})| is bounded below by 41b12ρ124^{-1}b^{-\frac{1}{2}}\rho^{-\frac{1}{2}} for ρ\rho sufficiently large. ∎

2.2. A tangent surface generated by a perturbed helix

In this subsection, we examine the tangent surface generated by the curve α(t)=(t,t2+t4,t3)\alpha(t)=(t,t^{2}+t^{4},t^{3}).

Proposition 2.4.

Let I=[2c,2c]I=[-2c,2c] for c>0c>0, and let α:I3\alpha:I\to\mathbb{R}^{3} be defined above. Let

S:={α(t)+vα(t)|tI,a<v<b}S:=\{\alpha(t)+v\alpha^{\prime}(t)|t\in I,a<v<b\}

be the tangent surface generated by α\alpha, where 0<a<b0<a<b. Then, dimF(S)1\dim_{F}(S)\leq 1.

We denote Φ:I×(a,b)3\Phi:I\times(a,b)\to\mathbb{R}^{3} be Φ(t,v)=α(t)+vα(t)\Phi(t,v)=\alpha(t)+v\alpha^{\prime}(t). Proposition 2.4 holds if the following lemma as an analog of (8) holds.

Lemma 2.5.

For sIs\in I, there exists Ts:S3T_{s}:S\to\mathbb{R}^{3}, such that

(10) Ts(α(t)+vα(t))e3=(α(t)+vα(t)α(s))n(s),T_{s}(\alpha(t)+v\alpha^{\prime}(t))\cdot e_{3}=(\alpha(t)+v\alpha^{\prime}(t)-\alpha(s))\cdot\overrightarrow{n}(s),

where

(11) n(s)=(3s2(2s21),3s,6s2+1)\overrightarrow{n}(s)=(-3s^{2}(2s^{2}-1),-3s,6s^{2}+1)

is a normal vector of SS at Φ(s,v)\Phi(s,v) with n(0)=e3\overrightarrow{n}(0)=e_{3}.

Proof of Proposition 2.4 assuming Lemma 2.5.

Let μ(S)\mu\in\mathcal{M}(S), and suppose that there exists β>0\beta>0, such that |μ^(ξ)|C|ξ|β2|\widehat{\mu}(\xi)|\leq C|\xi|^{-\frac{\beta}{2}} for ξ3\xi\in\mathbb{R}^{3}, C>0C>0.

We construct new measures. Let μs(3)\mu_{s}\in\mathcal{M}(\mathbb{R}^{3}) be

(12) μs=Ts#μ.\mu_{s}=T_{s}^{\#}\mu.

Let ψC0()\psi\in C_{0}^{\infty}(\mathbb{R}) be a non-negative bump function with ψ(s)=1\psi(s)=1 when |s|c|s|\leq c, ψ(s)=0\psi(s)=0 when |s|2c|s|\geq 2c. Then we let the average measure ν\nu be defined as

(13) f𝑑ν=f,μsψ(s)𝑑s\int fd\nu=\int\langle f,\mu_{s}\rangle\psi(s)ds

for non-negative Borel functions ff. Proposition 2.4 is a consequence of the following two lemmas.

Lemma 2.6.

For ρ>0\rho>0, |ν^(ρe3)|ρβ2|\widehat{\nu}(\rho e_{3})|\lesssim\rho^{-\frac{\beta}{2}}.

Lemma 2.7.

There exists ρ0>0\rho_{0}>0, such that for ρ>ρ0\rho>\rho_{0}, |ν^(ρe3)|ρ12|\widehat{\nu}(\rho e_{3})|\gtrsim\rho^{-\frac{1}{2}}. ∎

Compared with previous methods, the computations in Lemmas 2.6 and 2.7 do not require spt νS\text{spt }\nu\subset S.

Proof of Lemma 2.6.

First, we show that for each sIs\in I, |μs^(ρe3)|ρβ2|\widehat{\mu_{s}}(\rho e_{3})|\lesssim\rho^{-\frac{\beta}{2}}. By unwinding the definitions of the push-forward measure, Fourier transform, and integration on the manifold, we have

(14) μs^(ρe3)\displaystyle\widehat{\mu_{s}}(\rho e_{3}) =e2πiyρe3𝑑Ts#μ(y)\displaystyle=\int e^{-2\pi iy\cdot\rho e_{3}}dT_{s}^{\#}\mu(y) by (12)
=e2πiρTs(Φ(t,v))e3𝑑μΦ(t,v)\displaystyle=\int e^{-2\pi i\rho T_{s}(\Phi(t,v))\cdot e_{3}}d\mu\circ\Phi(t,v)
=e2πiρ[(α(t)+vα(t))n(s)α(s)n(s)]𝑑μΦ(t,v)\displaystyle=\int e^{-2\pi i\rho[(\alpha(t)+v\alpha^{\prime}(t))\cdot\overrightarrow{n}(s)-\alpha(s)\cdot\overrightarrow{n}(s)]}d\mu\circ\Phi(t,v) by (10)
=e2πiρα(s)n(s)μ^(ρn(s)).\displaystyle=e^{2\pi i\rho\alpha(s)\cdot\overrightarrow{n}(s)}\widehat{\mu}(\rho\overrightarrow{n}(s)).

Since |n(s)|1|\overrightarrow{n}(s)|\geq 1, |μs^(ρe3)|=|μ^(ρn(s))|C|ρn(s)|β2Cρβ2.|\widehat{\mu_{s}}(\rho e_{3})|=|\widehat{\mu}(\rho\overrightarrow{n}(s))|\leq C|\rho\overrightarrow{n}(s)|^{-\frac{\beta}{2}}\leq C\rho^{-\frac{\beta}{2}}. Then,

|ν^(ρe3)|\displaystyle|\widehat{\nu}(\rho e_{3})|\leq |μs^(ρe3)|ψ(s)𝑑s\displaystyle\int|\widehat{\mu_{s}}(\rho e_{3})|\psi(s)ds by (13)
\displaystyle\leq CψL1ρβ2.\displaystyle C\left\lVert\psi\right\rVert_{L^{1}}\rho^{-\frac{\beta}{2}}.
Proof of Lemma 2.7.

An expression of ν^(ρe3)\widehat{\nu}(\rho e_{3}) is given by

(15) ν^(ρe3)\displaystyle\widehat{\nu}(\rho e_{3}) =e2πiρTs(Φ(t,v))e3𝑑μΦ(t,v)ψ(s)𝑑s\displaystyle=\int\int e^{-2\pi i\rho T_{s}(\Phi(t,v))\cdot e_{3}}d\mu\circ\Phi(t,v)\psi(s)ds by (13)
=e2πiρTs(Φ(t,v))e3ψ(s)𝑑s𝑑μΦ(t,v).\displaystyle=\int\int e^{-2\pi i\rho T_{s}(\Phi(t,v))\cdot e_{3}}\psi(s)dsd\mu\circ\Phi(t,v).

We apply the stationary phase method to study the inner integral

I(ρ;t,v):=e2πiρTs(Φ(t,v))e3ψ(s)𝑑s.I(\rho;t,v):=\int e^{-2\pi i\rho T_{s}(\Phi(t,v))\cdot e_{3}}\psi(s)ds.

Fixing t,vt,v, we define the phase function ϕ\phi of I(ρ;t,v)I(\rho;t,v) as

ϕ(s)\displaystyle\phi(s) :=Ts(Φ(t,v))e3=(α(t)+vα(t))n(s)α(s)n(s)\displaystyle:=T_{s}(\Phi(t,v))\cdot e_{3}=(\alpha(t)+v\alpha^{\prime}(t))\cdot\overrightarrow{n}(s)-\alpha(s)\cdot\overrightarrow{n}(s) by (10)\displaystyle\text{ by }(\ref{p_phelix_ex})
=[(ts,t2+t4s2s4,t3s3)+v(1,2t+4t3,3t2)](6s4+3s2,3s,6s2+1)\displaystyle=[(t-s,t^{2}+t^{4}-s^{2}-s^{4},t^{3}-s^{3})+v(1,2t+4t^{3},3t^{2})]\cdot(-6s^{4}+3s^{2},-3s,6s^{2}+1)

Its first derivative is

ϕ(s)=3(ts)[(ts)3+4v(ts)2+(16s2)(ts)+2v(16s2)].\begin{split}\phi^{\prime}(s)=&-3(t-s)[(t-s)^{3}+4v(t-s)^{2}+(1-6s^{2})(t-s)+2v(1-6s^{2})].\end{split}

One critical point is s=ts=t. If cc is sufficiently small, the only critical point in spt ψ\text{spt }\psi is s=ts=t. Note that

ϕ(t)=0,ϕ(t)=0,ϕ′′(t)=v(636t2).\phi(t)=0,\phi^{\prime}(t)=0,\phi^{\prime\prime}(t)=v(6-36t^{2}).

Then, by the stationary phase (part b of Theorem 4.6),

|I(ρ;t,v)(iρv(636t2))12ψ(t)|D1(ρv)32\left|I(\rho;t,v)-\left(\frac{-i}{\rho v(6-36t^{2})}\right)^{\frac{1}{2}}\psi(t)\right|\leq D_{1}(\rho v)^{-\frac{3}{2}}

for a D1>0D_{1}>0 independent of ρ\rho, tt, and vv. Since ψ(t)=1\psi(t)=1 for |t|c|t|\leq c, continuing with (15),

|ν^(ρe3)(iρv(636t2))12𝑑μΦ(t,v)|D1(ρv)1𝑑μΦ(t,v)\begin{split}\left|\widehat{\nu}(\rho e_{3})-\int\left(\frac{-i}{\rho v(6-36t^{2})}\right)^{\frac{1}{2}}d\mu\circ\Phi(t,v)\right|&\leq\int D_{1}(\rho v)^{-1}d\mu\circ\Phi(t,v)\\ \end{split}

As v[a,b]v\in[a,b],

|(iρv(636t2))12𝑑μΦ(t,v)|\displaystyle\left|\int\left(\frac{-i}{\rho v(6-36t^{2})}\right)^{\frac{1}{2}}d\mu\circ\Phi(t,v)\right|\geq 612b12ρ12,\displaystyle 6^{-\frac{1}{2}}b^{-\frac{1}{2}}\rho^{-\frac{1}{2}},
|D1(ρv)1𝑑μΦ(t,v)|\displaystyle\left|\int D_{1}(\rho v)^{-1}d\mu\circ\Phi(t,v)\right|\leq D1a1ρ1.\displaystyle D_{1}a^{-1}\rho^{-1}.

Therefore, |ν^(ρe3)||\widehat{\nu}(\rho e_{3})| is bounded below by 41b12ρ124^{-1}b^{-\frac{1}{2}}\rho^{-\frac{1}{2}} for ρ\rho sufficiently large. ∎

Now, we show the existence of TsT_{s} for the perturbed helix.

Proof of Lemma 2.5.

To satisfy (10), a possible TsT_{s} fixes the first two coordinates of α(t)+vα(t)\alpha(t)+v\alpha^{\prime}(t) and updates the third coordinate to the right-hand side of (10). ∎

Remark.

The tangent surface generated by the curve γ(t)=(t,t2,t3)\gamma(t)=(t,t^{2},t^{3}) presented in Section 2.1 is one case where TsT_{s} satisfies (10) and Ts(S)ST_{s}(S)\subset S. For the example in this section, it is not possible to ensure that Ts(S)ST_{s}(S)\subset S and μs,ν\mu_{s},\nu are still supported on SS, even if we only require n(s)\overrightarrow{n}(s) to be any normal vector of SS at Φ(s,v)\Phi(s,v) and have a norm comparable to 1.

One interpretation is that TsT_{s} defined in the proof maps SSsS\to S_{s}, where SsS_{s} is the tangent surface generated by the curve

αs(t)=t,t2+t4,6s3(s1)(ts)+(16s2)(ts)33s(ts)4,\alpha_{s}(t)=\langle t,t^{2}+t^{4},6s^{3}(s-1)(t-s)+(1-6s^{2})(t-s)^{3}-3s(t-s)^{4}\rangle,

and

Ts(α(t)+vα(t))=αs(t)+vαs(t).T_{s}(\alpha(t)+v\alpha^{\prime}(t))=\alpha_{s}(t)+v\alpha_{s}^{\prime}(t).
α(t)+vα(t)\alpha(t)+v\alpha^{\prime}(t)SSαs(t)+vαs(t)\alpha_{s}(t)+v\alpha_{s}^{\prime}(t)SsS_{s} TsT_{s}
Figure 3. Illustration of the map Ts:SSsT_{s}:S\to S_{s}.

3. Proof of Proposition 1.2

We use terminologies commonly used in describing hypersurfaces of constant rank.

Definition 3.1.

MM is a (dk)(d-k)-ruled hypersurface if near pMp\in M, a coordinate chart is given by Φ:U×Vk×dkS\Phi:U\times V\subset\mathbb{R}^{k}\times\mathbb{R}^{d-k}\to S as

(16) Φ(u,v)=α(u)+l=1dkvlwl(u),\Phi(u,v)=\alpha(u)+\sum_{l=1}^{d-k}v_{l}w_{l}(u),

with α,w1,,wdk:kd+1\alpha,w_{1},\ldots,w_{d-k}:\mathbb{R}^{k}\to\mathbb{R}^{d+1}. A ruling is {Φ(u,v)|vV}\{\Phi(u,v)|v\in V\} for a fixed uu.

Let N(u,v)N(u,v) be the normal vector of MM at Φ(u,v)\Phi(u,v). In the proof of Proposition 1.2, we rely on the following parametrization on hypersurfaces with constant rank kk.

Lemma B.

[8, Lemma 3.1] [20] The following statements are equivalent.

  1. (1)

    MM is a smooth hypersurface in d+1\mathbb{R}^{d+1} of constant rank kk.

  2. (2)

    MM is (dk)(d-k)-ruled, and the normal vectors along the rulings are constant.

We deduce the analytic properties of constant rank hypersurfaces that will be applied throughout this section.

  1. a.

    The normal vectors along the rulings being constant is equivalent to

    (17) N(u,v)=N(u,v)N(u,v)=N(u,v^{\prime})

    for vvv\neq v^{\prime}. We may denote N(u)=N(u,v)N(u)=N(u,v) for a vVv\in V. Since N(u,v)N(u,v) is the normal vector at p=Φ(u,v)p=\Phi(u,v), it is perpendicular to all vectors in TpMT_{p}M. Therefore,

    (18) Φuj(u,v),N(u,v)=αuj(u)+l=1dkvlwluj(u),N(u,v)=0,Φvl(u,v),N(u,v)=wl(u),N(u,v)=0.\begin{split}\left\langle\frac{\partial\Phi}{\partial u_{j}}(u,v),N(u,v)\right\rangle&=\left\langle\frac{\partial\alpha}{\partial u_{j}}(u)+\sum_{l=1}^{d-k}v_{l}\frac{\partial w_{l}}{\partial u_{j}}(u),N(u,v)\right\rangle=0,\\ \left\langle\frac{\partial\Phi}{\partial v_{l}}(u,v),N(u,v)\right\rangle&=\left\langle w_{l}(u),N(u,v)\right\rangle=0.\end{split}

    From the first equation of (18), for a fixed uu, since N(u,v)=N(u,v)N(u,v)=N(u,v^{\prime}) when vvv\neq v^{\prime}, for all ll and jj,

    wluj(u),N(u,v)=0.\left\langle\frac{\partial w_{l}}{\partial u_{j}}(u),N(u,v)\right\rangle=0.

    By differentiating the second equation in (18) with respect to uju_{j}, we obtain

    (19) wl(u),Nuj(u,v)=0.\left\langle w_{l}(u),\frac{\partial N}{\partial u_{j}}(u,v)\right\rangle=0.
  2. b.

    The second fundamental form of MM at pp has rank kk. Since

    2Φvlvl(u,v)=0,\frac{\partial^{2}\Phi}{\partial v_{l}\partial v_{l^{\prime}}}(u,v)=0,

    and

    2Φujvl(u,v),N(u,v)=wluj(u),N(u,v)=0,\left\langle\frac{\partial^{2}\Phi}{\partial u_{j}\partial v_{l}}(u,v),N(u,v)\right\rangle=\left\langle\frac{\partial w_{l}}{\partial u_{j}}(u),N(u,v)\right\rangle=0,

    the top-left k×kk\times k submatrix of the second fundamental form

    (2Φujuj(u,v)N(u,v))\left(\frac{\partial^{2}\Phi}{\partial u_{j}\partial u_{j^{\prime}}}(u,v)\cdot N(u,v)\right)

    has full rank kk.

By rotation and translation, we may assume that UU is a neighborhood of 0, N(0)=ed+1d+1N(0)=e_{d+1}\in\mathbb{R}^{d+1}, where {e1,,ed+1}\{e_{1},\cdots,e_{d+1}\} is the standard basis of d+1\mathbb{R}^{d+1}.

Remark.
  1. a.

    A (dk)(d-k)-ruled hypersurface does not have a constant rank of kk if the normal vectors are not constant along the rulings, and an mm-ruled hypersurface can also be viewed as an (m1)(m-1)-ruled hypersurface. For example, in 3\mathbb{R}^{3}, the hyperboloid given by x2+y2=z2+1x^{2}+y^{2}=z^{2}+1 is 11-ruled, but it is a hypersurface with non-vanishing Gaussian curvature. [2, Sec. 3-5] The normal vectors of the hyperboloid are not constant along the rulings.

  2. b.

    The proof of Theorem A uses Morse’s Lemma (Lemma 4.3) to parametrize a hypersurface MM. The author’s previous work [21] partly uses Morse’s Lemma to parametrize cones and cylinders. A version of Proposition 1.2 could also be proven via Morse’s Lemma; however, the conclusion would be the same as the Fourier dimension is independent of the parametrizations.

  3. c.

    The hypersurface MM is cylindrical if wl(u1)=wl(u2)w_{l}(u_{1})=w_{l}(u_{2}) for all u1,u2Uu_{1},u_{2}\in U, 1ldk1\leq l\leq d-k. There is a simpler proof in Lemma 4.1 for a cylindrical hypersurface since it can be written as S×dkS\times\mathbb{R}^{d-k} for a hypersurface Sk+1S\subset\mathbb{R}^{k+1} with non-vanishing Gaussian curvature, and it follows that dimF(M)=dimF(S)=k\dim_{F}(M)=\dim_{F}(S)=k.

3.1. The Main Proof

Proposition 1.2 holds if the following lemma holds.

Lemma 3.2.

For sUs\in U, there exists Ts:Md+1T_{s}:M\to\mathbb{R}^{d+1}, such that

(20) Ts(Φ(u,v))ed+1=(Φ(u,v)α(s))N(s)T_{s}(\Phi(u,v))\cdot e_{d+1}=(\Phi(u,v)-\alpha(s))\cdot N(s)

for the unit normal vector N(s)N(s).

Proof of Proposition 1.2 assuming Lemma 3.2.

Let μ(M)\mu\in\mathcal{M}(M), and suppose that there exists β>0\beta>0, such that |μ^(ξ)|C|ξ|β2|\widehat{\mu}(\xi)|\leq C|\xi|^{-\frac{\beta}{2}} for ξd+1\xi\in\mathbb{R}^{d+1}, C>0C>0. By Lemma 4.2, we may assume μ(Φ(12U,V))\mu\in\mathcal{M}(\Phi(\frac{1}{2}U,V)).

We construct new measures. Let μs(d+1)\mu_{s}\in\mathcal{M}(\mathbb{R}^{d+1}) be

(21) μs=Ts#μ.\mu_{s}=T_{s}^{\#}\mu.

Let ψC0(k)\psi\in C_{0}^{\infty}(\mathbb{R}^{k}) be non-negative bump function with ψ(s)=1\psi(s)=1 when s12Us\in\frac{1}{2}U, ψ(s)=0\psi(s)=0 when s(34U)cs\in(\frac{3}{4}U)^{c}. Then, we define the average measure ν(d+1)\nu\in\mathcal{M}(\mathbb{R}^{d+1}) as

(22) f𝑑ν=f,μsψ(s)𝑑s\int fd\nu=\int\langle f,\mu_{s}\rangle\psi(s)ds

for non-negative Borel functions ff via the Riesz representation theorem for positive linear functionals [16]. Proposition 1.2 is a consequence of the following two lemmas.

Lemma 3.3.

For ρ>0\rho>0, |ν^(ρed+1)|ρβ2|\widehat{\nu}(\rho e_{d+1})|\lesssim\rho^{-\frac{\beta}{2}}.

Lemma 3.4.

There exists ρ0>0\rho_{0}>0, such that for ρ>ρ0\rho>\rho_{0}, |ν^(ρed+1)|ρk2|\widehat{\nu}(\rho e_{d+1})|\gtrsim\rho^{-\frac{k}{2}}. ∎

Proof of Lemma 3.3.

First, we show that for each sUs\in U, |μs^(ρed+1)|ρβ2|\widehat{\mu_{s}}(\rho e_{d+1})|\lesssim\rho^{-\frac{\beta}{2}}. By unwinding the definitions of the push-forward measure, Fourier transform, and integration on the manifold, we have

(23) μs^(ρed+1)\displaystyle\widehat{\mu_{s}}(\rho e_{d+1}) =e2πiyρed+1𝑑Ts#μ(y)\displaystyle=\int e^{-2\pi iy\cdot\rho e_{d+1}}dT_{s}^{\#}\mu(y) by (21)
=e2πiρTs(Φ(u,v))ed+1𝑑μΦ(u,v)\displaystyle=\int e^{-2\pi i\rho T_{s}(\Phi(u,v))\cdot e_{d+1}}d\mu\circ\Phi(u,v)
=e2πiρ[Φ(u,v)N(s)α(s)N(s)]𝑑μΦ(u,v)\displaystyle=\int e^{-2\pi i\rho[\Phi(u,v)\cdot N(s)-\alpha(s)\cdot N(s)]}d\mu\circ\Phi(u,v) by (20)
=e2πiρα(s)N(s)μ^(ρN(s)).\displaystyle=e^{2\pi i\rho\alpha(s)\cdot N(s)}\widehat{\mu}(\rho N(s)).

Since |N(s)|=1|N(s)|=1, |μs^(ρed+1)|=|μ^(ρN(s))|Cρβ2.|\widehat{\mu_{s}}(\rho e_{d+1})|=|\widehat{\mu}(\rho N(s))|\leq C\rho^{-\frac{\beta}{2}}. Then

|ν^(ρed+1)|\displaystyle|\widehat{\nu}(\rho e_{d+1})|\leq |μs^(ρN(s))|ψ(s)𝑑s\displaystyle\int|\widehat{\mu_{s}}(\rho N(s))|\psi(s)ds by (22)
\displaystyle\leq CψL1ρβ2.\displaystyle C\left\lVert\psi\right\rVert_{L^{1}}\rho^{-\frac{\beta}{2}}.
Proof of Lemma 3.4.

An expression of ν^(ρed+1)\widehat{\nu}(\rho e_{d+1}) is given by

(24) ν^(ρed+1)\displaystyle\widehat{\nu}(\rho e_{d+1}) =e2πiρTs(Φ(u,v))ed+1𝑑μΦ(u,v)ψ(s)𝑑s\displaystyle=\int\int e^{-2\pi i\rho T_{s}(\Phi(u,v))\cdot e_{d+1}}d\mu\circ\Phi(u,v)\psi(s)ds by (22)
=e2πiρTs(Φ(u,v))ed+1ψ(s)𝑑s𝑑μΦ(u,v).\displaystyle=\int\int e^{-2\pi i\rho T_{s}(\Phi(u,v))\cdot e_{d+1}}\psi(s)dsd\mu\circ\Phi(u,v).

We apply the stationary phase method to study the inner integral

I(ρ;u,v):=e2πiρTs(Φ(u,v))ed+1ψ(s)𝑑s.I(\rho;u,v):=\int e^{-2\pi i\rho T_{s}(\Phi(u,v))\cdot e_{d+1}}\psi(s)ds.

Fixing u,vu,v, we define the phase function ϕ\phi of I(ρ;t,v)I(\rho;t,v) as

ϕ(s):=Ts(Φ(u,v))ed+1=Φ(u,v)N(s)α(s)N(s) by (20).\phi(s):=T_{s}(\Phi(u,v))\cdot e_{d+1}=\Phi(u,v)\cdot N(s)-\alpha(s)\cdot N(s)\text{ by }(\ref{e621}).

Two claims about the critical point of ϕ\phi are as follows:

  • One critical point is s=us=u: by (19),

    ϕsj(u)=(Φ(u,v)α(u))Nuj(u)=(α(u)+l=1dkvlwl(u)α(u))Nuj(u)=l=1dkvlwl(u)Nuj(u)=0.\begin{split}\frac{\partial\phi}{\partial s_{j}}(u)=&(\Phi(u,v)-\alpha(u))\cdot\frac{\partial N}{\partial u_{j}}(u)\\ =&\left(\alpha(u)+\sum_{l=1}^{d-k}v_{l}w_{l}(u)-\alpha(u)\right)\cdot\frac{\partial N}{\partial u_{j}}(u)\\ =&\sum_{l=1}^{d-k}v_{l}w_{l}(u)\cdot\frac{\partial N}{\partial u_{j}}(u)=0.\end{split}
  • If UU is sufficiently small, no other critical point exists: we will show that the Hessian has full rank, so the critical points of the function ϕ\phi are isolated. By differentiating (19) with respect to uju_{j^{\prime}}, we obtain

    wluj(u),Nuj(u)+wl(u),2Nujuj(u)=0\left\langle\frac{\partial w_{l}}{\partial u_{j^{\prime}}}(u),\frac{\partial N}{\partial u_{j}}(u)\right\rangle+\left\langle w_{l}(u),\frac{\partial^{2}N}{\partial u_{j}\partial u_{j}^{\prime}}(u)\right\rangle=0

    Then,

    2ϕsjsj(u)=(Φ(u,v)α(u))2Nujuj(u)αuj(u)Nuj(u)=Φuj(u,v)Nuj(u).\begin{split}\frac{\partial^{2}\phi}{\partial s_{j}\partial s_{j^{\prime}}}(u)=&(\Phi(u,v)-\alpha(u))\cdot\frac{\partial^{2}N}{\partial u_{j}\partial u_{j^{\prime}}}(u)-\frac{\partial\alpha}{\partial u_{j^{\prime}}}(u)\cdot\frac{\partial N}{\partial u_{j}}(u)\\ =&-\frac{\partial\Phi}{\partial u_{j}^{\prime}}(u,v)\cdot\frac{\partial N}{\partial u_{j}}(u).\end{split}

    By part b) of the discussion following Lemma B, the matrix (2Φujuj(u,v)N(u))\left(\frac{\partial^{2}\Phi}{\partial u_{j}\partial u_{j^{\prime}}}(u,v)\cdot N(u)\right) has full rank kk. Therefore, Δsϕ(u)\Delta_{s}\phi(u) has full rank kk.

Then, by the stationary phase (part b of Theorem 4.6),

|I(ρ;u,v)ρk2(1)kik2(det(2Φujuj(u,v)N(u)))12ψ(u)|Dρk+12\left|I(\rho;u,v)-\rho^{-\frac{k}{2}}(-1)^{k}i^{\frac{k}{2}}\left(\det\left(\frac{\partial^{2}\Phi}{\partial u_{j}\partial u_{j^{\prime}}}(u,v)\cdot N(u)\right)\right)^{-\frac{1}{2}}\psi(u)\right|\leq D^{\prime}\rho^{-\frac{k+1}{2}}

for a D>0D^{\prime}>0 independent of ρ\rho, uu, and vv. Since ψ(u)=1\psi(u)=1 for u12Uu\in\frac{1}{2}U, continuing from (24),

|ν^(ρed+1)ρk2(1)kik2(det(2Φujuj(u,v)N(u)))12𝑑μΦ(u,v)|Dρk+12𝑑μΦ(u,v).\begin{split}&\left|\widehat{\nu}(\rho e_{d+1})-\int\rho^{-\frac{k}{2}}(-1)^{k}i^{\frac{k}{2}}\left(\det\left(\frac{\partial^{2}\Phi}{\partial u_{j}\partial u_{j^{\prime}}}(u,v)\cdot N(u)\right)\right)^{-\frac{1}{2}}d\mu\circ\Phi(u,v)\right|\\ \leq&\int D^{\prime}\rho^{-\frac{k+1}{2}}d\mu\circ\Phi(u,v).\\ \end{split}

Since

det(2Φujuj(u,v)N(u))D\det\left(\frac{\partial^{2}\Phi}{\partial u_{j}\partial u_{j^{\prime}}}(u,v)\cdot N(u)\right)\leq D

for a D>0D>0,

|ρk2(1)kik2(det(2Φujuj(u,v)N(u)))12𝑑μΦ(u,v)|\displaystyle\left|\int\rho^{-\frac{k}{2}}(-1)^{k}i^{\frac{k}{2}}\left(\det\left(\frac{\partial^{2}\Phi}{\partial u_{j}\partial u_{j^{\prime}}}(u,v)\cdot N(u)\right)\right)^{-\frac{1}{2}}d\mu\circ\Phi(u,v)\right|\geq D12ρk2,\displaystyle D^{-\frac{1}{2}}\rho^{-\frac{k}{2}},
|Dρk+12𝑑μΦ(u,v)|\displaystyle\left|\int D^{\prime}\rho^{-\frac{k+1}{2}}d\mu\circ\Phi(u,v)\right|\leq Dρk+12.\displaystyle D^{\prime}\rho^{-\frac{k+1}{2}}.

Therefore, |ν^(ρed+1)||\widehat{\nu}(\rho e_{d+1})| is bounded below by 21D12ρk22^{-1}D^{-\frac{1}{2}}\rho^{-\frac{k}{2}} for ρ\rho sufficiently large. ∎

3.2. Existence of TsT_{s}

Proof of Lemma 3.2.

One choice of TsT_{s} fixes the first nn coordinates of Φ(u,v)\Phi(u,v) and changes the last coordinate to

(Φ(u,v)α(s))N(s)=(α(u)α(s))N(s)+l=1dkvlwl(u)N(s).(\Phi(u,v)-\alpha(s))\cdot N(s)=(\alpha(u)-\alpha(s))\cdot N(s)+\sum_{l=1}^{d-k}v_{l}w_{l}(u)\cdot N(s).

Alternatively,

(25) Ts(Φ(u,v))=αs(u)+l=1dkvlwl,s(u),\begin{split}T_{s}(\Phi(u,v))=\alpha_{s}(u)+\sum_{l=1}^{d-k}v_{l}w_{l,s}(u),\end{split}

where

αs(u)=\displaystyle\alpha_{s}(u)= α(u)+ed+1[α(u)(N(s)ed+1)α(s)N(s)],\displaystyle\alpha(u)+e_{d+1}[\alpha(u)\cdot(N(s)-e_{d+1})-\alpha(s)\cdot N(s)],
wl,s(u)=\displaystyle w_{l,s}(u)= wl(u)+ed+1[wl(u)(N(s)ed+1)].\displaystyle w_{l}(u)+e_{d+1}[w_{l}(u)\cdot(N(s)-e_{d+1})].
Remark.

The new hypersurface Ms:=Ts(M)M_{s}:=T_{s}(M) is also dkd-k ruled with a coordinate chart Φs:=TsΦ\Phi_{s}:=T_{s}\circ\Phi if ss is sufficiently small. Since for all ll and jj, p=Ts(Φ(u,v))p=T_{s}(\Phi(u,v)),

wl,suj(u)TpMs,\frac{\partial w_{l,s}}{\partial u_{j}}(u)\in T_{p}M_{s},

for TsT_{s} defined in (25), the normal vectors of Ts(M)T_{s}(M) are stable along the rulings, so MsM_{s} is also a smooth hypersurface of constant rank kk.

4. Appendix

4.1. Hausdorff and Fourier dimensions

One notion of size often used in fractal geometry is the Hausdorff dimension, which is defined as follows. For a set AnA\subset\mathbb{R}^{n}, s,δ>0s,\delta>0,

δs(A):=inf{jdiam(Ej)s|AjEj,diam(Ej)<δ},\mathcal{H}^{s}_{\delta}(A):=\inf\left\{\left.\sum_{j}\text{diam}(E_{j})^{s}\right|A\subset\bigcup_{j}E_{j},\text{diam}(E_{j})<\delta\right\},

and the ss-dimensional Hausdorff measure s(A):=limδ0δs(A)\mathcal{H}^{s}(A):=\lim_{\delta\to 0}\mathcal{H}^{s}_{\delta}(A). The Hausdorff dimension of AA is dimH(A):=sup{s:s(A)=}=inf{s:s(A)=0}.\dim_{H}(A):=\sup\{s:\mathcal{H}^{s}(A)=\infty\}=\inf\{s:\mathcal{H}^{s}(A)=0\}.

Frostman’s lemma [15, Theorem 2.7] offers an alternative characterization of the Hausdorff dimension. For a Borel set AnA\subset\mathbb{R}^{n},

dimH(A)=sup{s[0,n]:μ(A),Is(μ):=|μ^(ξ)|2|ξ|sn𝑑ξ<},\dim_{H}(A)=\sup\{s\in[0,n]:\exists\mu\in\mathcal{M}(A),I_{s}(\mu):=\int|\widehat{\mu}(\xi)|^{2}|\xi|^{s-n}d\xi<\infty\},

where IsI_{s} is the ss-energy of μ\mu.

For μ(A)\mu\in\mathcal{M}(A), if supξn|ξ|s2|μ^(ξ)|<\sup_{\xi\in\mathbb{R}^{n}}|\xi|^{\frac{s}{2}}|\widehat{\mu}(\xi)|<\infty for an s>0s>0, Is0(μ)<I_{s_{0}}(\mu)<\infty when 0<s0<s0<s_{0}<s. From the characterization above, dimH(A)s\dim_{H}(A)\geq s. Consequently, dimF(A)dimH(A)\dim_{F}(A)\leq\dim_{H}(A), and the inequality is strict for some sets AA. The set AA is Salem if dimF(A)=dimH(A)\dim_{F}(A)=\dim_{H}(A).

4.2. Fourier dimension of the Cartesian product

We present a lemma that yields the Fourier dimension of cylindrical hypersurfaces.

Lemma 4.1.

Suppose that SnS\subset\mathbb{R}^{n}, TmT\subset\mathbb{R}^{m} are compact, dimF(S)<n\dim_{F}(S)<n, and dimF(T)<m\dim_{F}(T)<m. Then,

  1. a.
    (26) dimF(S×T)=min{dimF(S),dimF(T)},\dim_{F}(S\times T)=\min\{\dim_{F}(S),\dim_{F}(T)\},
  2. b.
    (27) dimF(S×m)=dimF(S).\dim_{F}(S\times\mathbb{R}^{m})=\dim_{F}(S).

We refer readers to the discussion in [4] on the Hausdorff dimension of the Cartesian product.

Proof.

For part a, let s:=dimF(S)<ns:=\dim_{F}(S)<n, t:=dimF(T)<mt:=\dim_{F}(T)<m. We write x=(x1,x2)n×mx=(x_{1},x_{2})\in\mathbb{R}^{n}\times\mathbb{R}^{m}.

  • First, we show that

    dimF(S×T)min{s,t}.\dim_{F}(S\times T)\geq\min\{s,t\}.

    Let ε>0\varepsilon>0. Let μS(S)\mu_{S}\in\mathcal{M}(S), with |μS^(ξ1)|C1|ξ1|s2+ε|\widehat{\mu_{S}}(\xi_{1})|\leq C_{1}|\xi_{1}|^{-\frac{s}{2}+\varepsilon} for C1>0C_{1}>0, ξ1n\xi_{1}\in\mathbb{R}^{n}, |ξ1|1|\xi_{1}|\geq 1.

    Similarly, let μT(T)\mu_{T}\in\mathcal{M}(T), with |μT^(ξ2)|C2|ξ2|t2+ε|\widehat{\mu_{T}}(\xi_{2})|\leq C_{2}|\xi_{2}|^{-\frac{t}{2}+\varepsilon} for C2>0C_{2}>0, ξ2m\xi_{2}\in\mathbb{R}^{m}, |ξ2|1|\xi_{2}|\geq 1. We consider the measure μS×μT(S×T)\mu_{S}\times\mu_{T}\in\mathcal{M}(S\times T). For ξ=(ξ1,ξ2)n×m\xi=(\xi_{1},\xi_{2})\in\mathbb{R}^{n}\times\mathbb{R}^{m},

    μS×μT^(ξ)=n+me2πixξd(μS×μT)(x)=(ne2πix1ξ1𝑑μS(x1))(me2πix2ξ2𝑑μT(x2))=μS^(ξ1)μT^(ξ2).\begin{split}\widehat{\mu_{S}\times\mu_{T}}(\xi)&=\int_{\mathbb{R}^{n+m}}e^{-2\pi ix\cdot\xi}d(\mu_{S}\times\mu_{T})(x)\\ =&\left(\int_{\mathbb{R}^{n}}e^{-2\pi ix_{1}\cdot\xi_{1}}d\mu_{S}(x_{1})\right)\left(\int_{\mathbb{R}^{m}}e^{-2\pi ix_{2}\cdot\xi_{2}}d\mu_{T}(x_{2})\right)\\ =&\widehat{\mu_{S}}(\xi_{1})\widehat{\mu_{T}}(\xi_{2}).\end{split}

    There are two cases to consider.

    • Case 1: |ξ1|12|ξ||\xi_{1}|\geq\frac{1}{2}|\xi|. Then,

      |μS×μT^(ξ)|=|μS^(ξ1)μT^(ξ2)|C1|ξ1|s2+εC12s2ε|ξ|s2+ε.|\widehat{\mu_{S}\times\mu_{T}}(\xi)|=|\widehat{\mu_{S}}(\xi_{1})\widehat{\mu_{T}}(\xi_{2})|\leq C_{1}|\xi_{1}|^{-\frac{s}{2}+\varepsilon}\leq C_{1}2^{\frac{s}{2}-\varepsilon}|\xi|^{-\frac{s}{2}+\varepsilon}.
    • Case 2: |ξ2|12|ξ||\xi_{2}|\geq\frac{1}{2}|\xi|. Similarly, |μS×μT^(ξ)|C22t2ε|ξ|t2+ε.|\widehat{\mu_{S}\times\mu_{T}}(\xi)|\leq C_{2}2^{\frac{t}{2}-\varepsilon}|\xi|^{-\frac{t}{2}+\varepsilon}.

    This shows that dimF(S×T)min{s2ε,t2ε}\dim_{F}(S\times T)\geq\min\{s-2\varepsilon,t-2\varepsilon\}, then we can let ε0\varepsilon\to 0.

  • Next, we show that

    dimF(S×T)min{s,t}.\dim_{F}(S\times T)\leq\min\{s,t\}.

    Let μ(S×T)\mu\in\mathcal{M}(S\times T), and suppose that |μ^(ξ)|C|ξ|r2|\widehat{\mu}(\xi)|\leq C|\xi|^{-\frac{r}{2}} for ξ=(ξ1,ξ2)n×m\xi=(\xi_{1},\xi_{2})\in\mathbb{R}^{n}\times\mathbb{R}^{m}. We will show that rsr\leq s and rtr\leq t.

    • Let πS:S×TS\pi_{S}:S\times T\to S be the projection map to the first nn coordinates and πT:S×TT\pi_{T}:S\times T\to T be the projection map to the last mm coordinates. Since SS and TT are compact, for μ(S×T)\mu\in\mathcal{M}(S\times T),

      sptπS#μ=πS(sptμ)S,\text{spt}\pi_{S}^{\#}\mu=\pi_{S}(\text{spt}\mu)\subset S,
      sptπT#μ=πT(sptμ)T.\text{spt}\pi_{T}^{\#}\mu=\pi_{T}(\text{spt}\mu)\subset T.
    • First, consider ξ=(ξ1,0)\xi=(\xi_{1},0) (so ξ2=0\xi_{2}=0). Then,

      (28) μ^(ξ)=e2πixξ𝑑μ(x)=e2πix1ξ1𝑑μ(x)=e2πiπS(x)ξ1𝑑μ(x)=e2πiyξ1𝑑πS#μ(y)=πS#μ^(ξ1)\begin{split}\widehat{\mu}(\xi)&=\int e^{-2\pi ix\cdot\xi}d\mu(x)\\ &=\int e^{-2\pi ix_{1}\cdot\xi_{1}}d\mu(x)\\ &=\int e^{-2\pi i\pi_{S}(x)\cdot\xi_{1}}d\mu(x)\\ &=\int e^{-2\pi iy\cdot\xi_{1}}d\pi_{S}^{\#}\mu(y)\\ &=\widehat{\pi_{S}^{\#}\mu}(\xi_{1})\end{split}

      Since πS#μ(S)\pi_{S}^{\#}\mu\in\mathcal{M}(S), dimF(S)=s\dim_{F}(S)=s, |πS#μ^(ξ1)|=|μ^(ξ)|C|ξ|r2=C|ξ1|r2|\widehat{\pi_{S}^{\#}\mu}(\xi_{1})|=|\widehat{\mu}(\xi)|\leq C|\xi|^{-\frac{r}{2}}=C|\xi_{1}|^{-\frac{r}{2}}, then rsr\leq s.

    • Next, consider ξ=(0,ξ2)\xi=(0,\xi_{2}) (so ξ1=0\xi_{1}=0). Similar to (28), |μ^(ξ)|=|πT#μ^(ξ2)||\widehat{\mu}(\xi)|=|\widehat{\pi_{T}^{\#}\mu}(\xi_{2})|. As πT#μ(T)\pi_{T}^{\#}\mu\in\mathcal{M}(T), dimF(T)=t\dim_{F}(T)=t, |πT#μ^(ξ2)|=|μ^(ξ)|C|ξ|r2=C|ξ2|r2|\widehat{\pi_{T}^{\#}\mu}(\xi_{2})|=|\widehat{\mu}(\xi)|\leq C|\xi|^{-\frac{r}{2}}=C|\xi_{2}|^{-\frac{r}{2}}, we have rtr\leq t.

The proof for part b) is similar, where we can take μT𝒮(m)\mu_{T}\in\mathcal{S}(\mathbb{R}^{m}). ∎

4.3. Reduction to compactly supported measure

We present a lemma that allows us to assume that the measure μ(M)\mu\in\mathcal{M}(M) we study has smaller support in a smaller closed set on the manifold MM.

Lemma 4.2.

[3, Theorem 1] Suppose that μ0(n)\mu_{0}\in\mathcal{M}(\mathbb{R}^{n}), supξn|ξ|α|μ0^(ξ)|<\sup_{\xi\in\mathbb{R}^{n}}|\xi|^{\alpha}|\widehat{\mu_{0}}(\xi)|<\infty for α>0\alpha>0. Let f𝒮(n)f\in\mathcal{S}(\mathbb{R}^{n}) with f0f\geq 0, and μ(n)\mu\in\mathcal{M}(\mathbb{R}^{n}) such that dμ=fdμ0d\mu=fd\mu_{0}. Then, |μ^(ξ)|cμ|ξ|α|\widehat{\mu}(\xi)|\leq c_{\mu}|\xi|^{-\alpha} for a cμ>0c_{\mu}>0.

Proof.

Note that

|μ^(ξ)|=|μ0^f^(ξ)|=|μ0^(ξη)f^(η)𝑑η|||η||ξ|2μ0^(ξη)f^(η)𝑑η|+||η||ξ|2μ0^(ξη)f^(η)𝑑η|.\begin{split}|\widehat{\mu}(\xi)|&=|\widehat{\mu_{0}}*\widehat{f}(\xi)|\\ &=\left|\int\widehat{\mu_{0}}(\xi-\eta)\widehat{f}(\eta)d\eta\right|\\ &\leq\left|\int_{|\eta|\leq\frac{|\xi|}{2}}\widehat{\mu_{0}}(\xi-\eta)\widehat{f}(\eta)d\eta\right|+\left|\int_{|\eta|\geq\frac{|\xi|}{2}}\widehat{\mu_{0}}(\xi-\eta)\widehat{f}(\eta)d\eta\right|.\end{split}

For the first integral, since |η||ξ|2|\eta|\leq\frac{|\xi|}{2}, |ξη||ξ|2|\xi-\eta|\geq\frac{|\xi|}{2}, and f^\widehat{f} is integrable, the integral is bounded above by a constant multiple of |ξ|α|\xi|^{-\alpha}. For the second integral, we apply the bounds μ0^L=1\left\lVert\widehat{\mu_{0}}\right\rVert_{L^{\infty}}=1 and |η||ξ|2|f^(η)|𝑑ηcm|ξ|m\int_{|\eta|\geq\frac{|\xi|}{2}}|\widehat{f}(\eta)|d\eta\leq c_{m}|\xi|^{-m} for mm\in\mathbb{N} with a cm>0c_{m}>0 since f^𝒮(n)\widehat{f}\in\mathcal{S}(\mathbb{R}^{n}). If we choose m>αm>\alpha, the sum of two bounds is bounded by a constant multiple of |ξ|α|\xi|^{-\alpha} for large |ξ||\xi|. ∎

4.4. Morse’s lemma

The version stated below generalizes the one shown in [18, Section 8.2].

Lemma 4.3.

Suppose that fC(d×d)f\in C^{\infty}(\mathbb{R}^{d}\times\mathbb{R}^{d}), f(0,t)=0f(0,t)=0, xf(0,t)=0\nabla_{x}f(0,t)=0, and that Δxf(x,t)\Delta_{x}f(x,t) has a constant rank of nn. Then, there exist neighborhoods V,WV,W of 0 and a smooth τ:V×Wd\tau:V\times W\to\mathbb{R}^{d} such that τ(0,t)=0\tau(0,t)=0, det𝐉xτ(x,t)0\det{\bf J}_{x}\tau(x,t)\neq 0, and

(29) f(x,t)=Qm,n(τ(x,t)),f(x,t)=Q_{m,n}(\tau(x,t)),

where

(30) Qm,n(y):=j=1myj2j=m+1nyj2Q_{m,n}(y):=\sum_{j=1}^{m}y_{j}^{2}-\sum_{j=m+1}^{n}y_{j}^{2}

with 0mnd0\leq m\leq n\leq d.

Remark.

The number of positive eigenvalues of the matrix Δxf(0,0)\Delta_{x}f(0,0) is mm. By differentiating (29) twice,

(31) [𝐉xτ(0,t)]TΔQm,n(0)𝐉xτ(0,t)=Δxf(0,t).[{\bf J}_{x}\tau(0,t)]^{T}\Delta Q_{m,n}(0){\bf J}_{x}\tau(0,t)=\Delta_{x}f(0,t).
Proof.

We claim that the function τ\tau can be expressed as

(32) τ=Lτdτ1,\tau=L\circ\tau_{d}\circ\cdots\circ\tau_{1},

where each τr\tau_{r} is a change of variables in the first dd coordinates and where LL is a permutation of the same coordinates. τr\tau_{r} is constructed inductively as follows: suppose that at step rr, we have

gr~(x,t):=f(τ11τr11(x,t))=±[x1]2±±[xr1]2+j,krdxjxkfj,k~(x,t),\widetilde{g_{r}}(x,t):=f(\tau_{1}^{-1}\circ\cdots\circ\tau_{r-1}^{-1}(x,t))=\pm[x_{1}]^{2}\pm\cdots\pm[x_{r-1}]^{2}+\sum_{j,k\geq r}^{d}x_{j}x_{k}\widetilde{f_{j,k}}(x,t),

where

fj,k~(x,t)=01(1s)2gr~xjxk(sx,t)𝑑s,fj,k~(0,t)=122gr~xjxk(0,t).\widetilde{f_{j,k}}(x,t)=\int_{0}^{1}(1-s)\frac{\partial^{2}\widetilde{g_{r}}}{\partial x_{j}\partial x_{k}}(sx,t)ds,\widetilde{f_{j,k}}(0,t)=\frac{1}{2}\frac{\partial^{2}\widetilde{g_{r}}}{\partial x_{j}\partial x_{k}}(0,t).

When r>nr>n, since Δxgr~(x,t)\Delta_{x}\widetilde{g_{r}}(x,t) only has rank nn, all fj,k~(x,t)=0\widetilde{f_{j,k}}(x,t)=0, and we are done. Otherwise, there exists an orthonormal matrix OrO_{r}, which is a linear change in the variables xr,,xdx_{r},\cdots,x_{d}, such that

gr(y,t):=f(τ11τr11(Ory,t))=±[y1]2±±[yr1]2+j,krdyjykfj,k(y,t),g_{r}(y,t):=f(\tau_{1}^{-1}\circ\cdots\circ\tau_{r-1}^{-1}(O_{r}y,t))=\pm[y_{1}]^{2}\pm\cdots\pm[y_{r-1}]^{2}+\sum_{j,k\geq r}^{d}y_{j}y_{k}f_{j,k}(y,t),

where

fj,k(y,t)=01(1s)2gryjyk(sy,t)𝑑s,fj,k(0,t)=122gryjyk(0,t)f_{j,k}(y,t)=\int_{0}^{1}(1-s)\frac{\partial^{2}g_{r}}{\partial y_{j}\partial y_{k}}(sy,t)ds,f_{j,k}(0,t)=\frac{1}{2}\frac{\partial^{2}g_{r}}{\partial y_{j}\partial y_{k}}(0,t)

and the additional condition that fr,r(0,t)0f_{r,r}(0,t)\neq 0. Then, for each tt, we can perform a change of variables from yy to zz such that zj=yjz_{j}=y_{j} for jrj\neq r, and

(33) zr=[±fr,r(y,t)]12[yr+j>ryjfj,r(y,t)±fr,r(y,t)],z_{r}=[\pm f_{r,r}(y,t)]^{\frac{1}{2}}\left[y_{r}+\sum_{j>r}\frac{y_{j}f_{j,r}(y,t)}{\pm f_{r,r}(y,t)}\right],

where ±\pm is the sign of fr,r(0,t)f_{r,r}(0,t). This change of variables can be expressed as z=σr(y,t)z=\sigma_{r}(y,t) for σr:Vr×Wrd\sigma_{r}:V_{r}\times W_{r}\to\mathbb{R}^{d}, where Vr,WrV_{r},W_{r} are neighborhoods of 0, fr,rf_{r,r} does not change sign on Vr×WrV_{r}\times W_{r}, and

det𝐉yσr(y,t)0.\det{\bf J}_{y}\sigma_{r}(y,t)\neq 0.

Then, we let

(34) τr(x,t)=(σr(Or1x,t),t),\tau_{r}(x,t)=(\sigma_{r}(O_{r}^{-1}x,t),t),

and proceed to the next step by noting that τr1\tau_{r}^{-1} exists in a neighborhood of 0.

From the construction above, there exist V,WV,W neighborhoods of 0, such that for (x,t)=(x,t)V×W(x,t)=(x,t)\in V\times W, for all rr from 11 to dd, fr,r(y,t)f_{r,r}(y,t) does not change sign, and

det𝐉yσr(y,t)0.\det{\bf J}_{y}\sigma_{r}(y,t)\neq 0.

Therefore, τ\tau is defined on V×WV\times W. ∎

Let τt(x)=τ(x,t)\tau_{t}(x)=\tau(x,t). For k0k\geq 0, tWt\in W, it is possible to bound 𝐉τtCk(V)\left\lVert{\bf J}\tau_{t}\right\rVert_{C^{k}(V)} and inf{|det𝐉τt(x)|xV}\inf\{|\det{\bf J}\tau_{t}(x)|x\in V\} via ftCk+2\left\lVert f_{t}\right\rVert_{C^{k+2}}, |det𝐉τt(0)||\det{\bf J}\tau_{t}(0)|, and the size of VV via (32), (33), and (34).

4.5. Oscillatory Integrals

We refer readers to the complete proof of oscillatory integrals results in [18]. In this section, we outline the key steps and bounds of the error terms.

For fC(n)f\in C^{\infty}(\mathbb{R}^{n}), UU open in n\mathbb{R}^{n}, and k0k\geq 0, we denote fCk(U)\left\lVert f\right\rVert_{C^{k}(U)}, or fCk\left\lVert f\right\rVert_{C^{k}} if the implication of the open set UU is clear, as the quantity

|β|k|β|yβfL(U).\sum_{|\beta|\leq k}\left\lVert\frac{\partial^{|\beta|}}{\partial y^{\beta}}f\right\rVert_{L^{\infty}(U)}.
Proposition 4.4.

[18, Chapter 8, Propositions 1 and 4] (Localization)

  1. a.

    (Single variable edition) For NN\in\mathbb{N}, there exist CN>0C_{N}>0, pN,qNp_{N},q_{N}\in\mathbb{N} such that, for ϕC()\phi\in C^{\infty}(\mathbb{R}), ψC0()\psi\in C^{\infty}_{0}(\mathbb{R}) with |ϕ(x)|1|\phi^{\prime}(x)|\geq 1 in spt ψ\text{spt }\psi, λ1\lambda\geq 1,

    |eiλϕ(x)ψ(x)𝑑x|CNλN|spt ψ|ψCNpNϕ′′CN1qN\left|\int e^{i\lambda\phi(x)}\psi(x)dx\right|\leq C_{N}\lambda^{-N}|\text{spt }\psi|\left\lVert\psi\right\rVert_{C^{N}}^{p_{N}}\left\lVert\phi^{\prime\prime}\right\rVert_{C^{N-1}}^{q_{N}}
  2. b.

    (Multivariate edition) For NN\in\mathbb{N}, there exist qN,rNq_{N},r_{N}\in\mathbb{N}, such that for ϕC(n)\phi\in C^{\infty}(\mathbb{R}^{n}), ψC0(n)\psi\in C^{\infty}_{0}(\mathbb{R}^{n}) with |ϕ(x)|1|\nabla\phi(x)|\geq 1 in spt ψ\text{spt }\psi, Λ\Lambda being the maximum absolute values of the eigenvalues of Δϕ(x)\Delta\phi(x^{\prime}) for all xspt ψx^{\prime}\in\text{spt }\psi, λ1\lambda\geq 1, there exists CN,ψ>0C_{N,\psi}>0 with

    |eiλϕ(x)ψ(x)dx|CN,ψλN(1+Λ)rNmaxξ𝕊n12ϕξ2CN1qN.\left|\int e^{i\lambda\phi(x)}\psi(x)dx\right|\leq C_{N,\psi}\lambda^{-N}(1+\Lambda)^{r_{N}}\max_{\xi\in\mathbb{S}^{n-1}}\left\lVert\frac{\partial^{2}\phi}{\partial\xi^{2}}\right\rVert_{C^{N-1}}^{q_{N}}.
Proof.
  1. a.

    Using integration by parts NN times, the integral can be written as

    eiλϕ(x)(tD)N{ψ(x)}dx\int e^{i\lambda\phi(x)}(^{t}D)^{N}\{\psi(x)\}dx

    with Dtg=iλ1ddx(gϕ){}^{t}Dg=i\lambda^{-1}\frac{d}{dx}\left(\frac{g}{\phi^{\prime}}\right). Note that

    λN(tD)Ng=g(N)(ϕ)N+r=1N(ϕ)(N+r)FN,r(g,g,,g(N1),ϕ′′,,ϕ(N+1))\lambda^{N}(^{t}D)^{N}g=\frac{g^{(N)}}{(\phi^{\prime})^{N}}+\sum_{r=1}^{N}(\phi^{\prime})^{-(N+r)}F_{N,r}(g,g^{\prime},\cdots,g^{(N-1)},\phi^{\prime\prime},\cdots,\phi^{(N+1)})

    for some polynomials FN,rF_{N,r}. Therefore, the integral is bounded above by

    CNλN|spt ψ|ψCNpNϕ′′CN1qNC_{N}\lambda^{-N}|\text{spt }\psi|\left\lVert\psi\right\rVert_{C^{N}}^{p_{N}}\left\lVert\phi^{\prime\prime}\right\rVert_{C^{N-1}}^{q_{N}}

    for CN>0C_{N}>0, pN,qNp_{N},q_{N}\in\mathbb{N}.

  2. b.

    For each x0x_{0} in the support of ψ\psi, there is a unit vector ξ0=(ϕ)(x0)/(ϕ)(x0)\xi_{0}=(\nabla\phi)(x_{0})/\left\lVert(\nabla\phi)(x_{0})\right\rVert and a ball centered B(x0)B(x_{0}) at x0x_{0} with

    (35) ξ0(ϕ)(x)21\xi_{0}\cdot(\nabla\phi)(x)\geq 2^{-1}

    if xB(x0)x\in B(x_{0}). We claim that B(x0)B(x_{0}) can be chosen with the radius bounded below by a uniform constant depending on ϕ\phi. By the mean-value theorem, there exists xx^{\prime} on the line segment joining x0x_{0} and xx, such that

    |ξ0(ϕ)(x)ξ0(ϕ)(x0)|=|(xx0)Δϕ(x)ξ0|.|\xi_{0}\cdot(\nabla\phi)(x)-\xi_{0}\cdot(\nabla\phi)(x_{0})|=|(x-x_{0})\Delta\phi(x^{\prime})\xi_{0}|.

    If

    |xx0|(2(1+Λ))1,|x-x_{0}|\leq(2(1+\Lambda))^{-1},

    then (35) holds.

    We choose a finite cover {Bk}{B(x0)}\{B_{k}\}\subset\{B(x_{0})\}, and write ψ=kψk\psi=\sum_{k}\psi_{k} for spt ψkBk\text{spt }\psi_{k}\subset B_{k}. For each BkB_{k}, we choose a coordinate system where x1x_{1} is parallel to ξk\xi_{k}. Then,

    eiλϕ(x)ψk(x)𝑑x=(eiλϕ(x1,,xn)ψk(x1,,xn)𝑑x1)𝑑x2𝑑xn,\int e^{i\lambda\phi(x)}\psi_{k}(x)dx=\int\left(\int e^{i\lambda\phi(x_{1},\cdots,x_{n})}\psi_{k}(x_{1},\cdots,x_{n})dx_{1}\right)dx_{2}\cdots dx_{n},

    and we can apply part a) to the inner integral and bound it above by

    CNλN|spt ψk|ψkCNpN2ϕx12CN1qN.C_{N}\lambda^{-N}|\text{spt }\psi_{k}|\left\lVert\psi_{k}\right\rVert_{C^{N}}^{p_{N}}\left\lVert\frac{\partial^{2}\phi}{\partial x_{1}^{2}}\right\rVert_{C^{N-1}}^{q_{N}}.

    When all kk’s are summed, a bound on the original integral is

    CNλN|spt ψ|ψCNpN(1+Λ)rNmaxξ𝕊n12ϕξ2CN1qNC_{N}^{\prime}\lambda^{-N}|\text{spt }\psi|\left\lVert\psi\right\rVert_{C^{N}}^{p_{N}}(1+\Lambda)^{r_{N}}\max_{\xi\in\mathbb{S}^{n-1}}\left\lVert\frac{\partial^{2}\phi}{\partial\xi^{2}}\right\rVert_{C^{N-1}}^{q_{N}}

    for CN>0C_{N}^{\prime}>0, and rNr_{N}\in\mathbb{N}. ∎

Lemma 4.5.

[18, Chapter 8, Proposition 6] Let Qm,n(y)Q_{m,n}(y) be defined in (30).

  1. a.

    If ηC0(n)\eta\in C^{\infty}_{0}(\mathbb{R}^{n}),

    (36) |neiλQm,n(y)ylη(y)𝑑y|Alλn+|l|2,\left|\int_{\mathbb{R}^{n}}e^{i\lambda Q_{m,n}(y)}y^{l}\eta(y)dy\right|\leq A_{l}\lambda^{-\frac{n+|l|}{2}},

    where Al>0A_{l}>0, lnl\in\mathbb{Z}^{n}, lj0l_{j}\geq 0, and |l|=j=1nlj|l|=\sum_{j=1}^{n}l_{j}.

  2. b.

    If g𝒮(n)g\in\mathcal{S}(\mathbb{R}^{n}) and there exists δ>0\delta>0, such that g(y)=0g(y)=0 for yB(0,δ)y\in B(0,\delta), then for NN\in\mathbb{N}, there exists BN>0B_{N}>0 such that

    (37) |neiλQm,n(y)g(y)𝑑y|BNλN.\left|\int_{\mathbb{R}^{n}}e^{i\lambda Q_{m,n}(y)}g(y)dy\right|\leq B_{N}\lambda^{-N}.
Proof.
  1. a.

    Consider the cones

    Γk={yn||yk|2|y|22n},\Gamma_{k}=\left\{y\in\mathbb{R}^{n}||y_{k}|^{2}\geq\frac{|y|^{2}}{2n}\right\},

    and

    Γk0={yn||yk|2|y|2n}.\Gamma_{k}^{0}=\left\{y\in\mathbb{R}^{n}||y_{k}|^{2}\geq\frac{|y|^{2}}{n}\right\}.

    Since k=1nΓk0=n\cup_{k=1}^{n}\Gamma_{k}^{0}=\mathbb{R}^{n}, there are functions {Ωk}1kn\{\Omega_{k}\}_{1\leq k\leq n} such that each Ωk\Omega_{k} is homogeneous of degree 0, smooth away from the origin, 0Ωk10\leq\Omega_{k}\leq 1 with

    k=1nΩk(x)=1\sum_{k=1}^{n}\Omega_{k}(x)=1

    for x0x\neq 0, and each Ωk\Omega_{k} is supported in Γk\Gamma_{k}. Then,

    neiλQm,n(y)ylη(y)𝑑y=k=1nΓkeiλQm,n(y)ylη(y)Ωk(y)𝑑y.\int_{\mathbb{R}^{n}}e^{i\lambda Q_{m,n}(y)}y^{l}\eta(y)dy=\sum_{k=1}^{n}\int_{\Gamma_{k}}e^{i\lambda Q_{m,n}(y)}y^{l}\eta(y)\Omega_{k}(y)dy.

    In the cone Γk\Gamma_{k}, we will show that there exists Al,k>0A_{l,k}>0 such that

    (38) |ΓkeiλQm,n(y)ylη(y)Ωk(y)𝑑y|Al,kλn+|l|2.\left|\int_{\Gamma_{k}}e^{i\lambda Q_{m,n}(y)}y^{l}\eta(y)\Omega_{k}(y)dy\right|\leq A_{l,k}\lambda^{-\frac{n+|l|}{2}}.

    Summing over all kk yields (36). Let αC(n)\alpha\in C^{\infty}(\mathbb{R}^{n}), such that α(y)=1\alpha(y)=1 for |y|1|y|\leq 1, and α(y)=0\alpha(y)=0 for |y|2|y|\geq 2. Then, for ϵ>0\epsilon>0,

    ΓkeiλQm,n(y)ylη(y)Ωk(y)𝑑y=ΓkeiλQm,n(y)ylη(y)Ωk(y)α(ϵ1y)𝑑y+ΓkeiλQm,n(y)ylη(y)Ωk(y)[1α(ϵ1y)]𝑑y.\begin{split}&\int_{\Gamma_{k}}e^{i\lambda Q_{m,n}(y)}y^{l}\eta(y)\Omega_{k}(y)dy\\ =&\int_{\Gamma_{k}}e^{i\lambda Q_{m,n}(y)}y^{l}\eta(y)\Omega_{k}(y)\alpha(\epsilon^{-1}y)dy+\int_{\Gamma_{k}}e^{i\lambda Q_{m,n}(y)}y^{l}\eta(y)\Omega_{k}(y)[1-\alpha(\epsilon^{-1}y)]dy.\end{split}

    For the first integral,

    |ΓkeiλQm,n(y)ylη(y)Ωk(y)α(ϵ1y)𝑑y|ηLϵ|l|+1.\left|\int_{\Gamma_{k}}e^{i\lambda Q_{m,n}(y)}y^{l}\eta(y)\Omega_{k}(y)\alpha(\epsilon^{-1}y)dy\right|\lesssim\left\lVert\eta\right\rVert_{L^{\infty}}\epsilon^{|l|+1}.

    Let NN\in\mathbb{N}. Using integration by parts NN times, the second integral can be written as

    (39) ΓkeiλQm,n(y)(tDk)N{ylη(y)Ωk(y)[1α(ϵ1y)]}dy\int_{\Gamma_{k}}e^{i\lambda Q_{m,n}(y)}(^{t}D_{k})^{N}\{y^{l}\eta(y)\Omega_{k}(y)[1-\alpha(\epsilon^{-1}y)]\}dy

    with Dktg=sk(2iλ)1yk(gyk){}^{t}D_{k}g=s_{k}(2i\lambda)^{-1}\frac{\partial}{\partial y_{k}}\left(\frac{g}{y_{k}}\right) for a differentiable function gg, and sk=1s_{k}=-1 if kmk\leq m, sk=1s_{k}=1 if km+1k\geq m+1. Note that

    (40) (tDk)Ng=λNr=0NaN,r(m,k)ykr2Nrgykr(^{t}D_{k})^{N}g=\lambda^{-N}\sum_{r=0}^{N}a^{(m,k)}_{N,r}y_{k}^{r-2N}\frac{\partial^{r}g}{\partial y_{k}^{r}}

    for aN,r(m,k)a^{(m,k)}_{N,r}\in\mathbb{C}. When we apply (40) and the product rule of the derivative to expand (39), we obtain a summation of terms where a term, ignoring the constant, is

    λNΓkB(0,ϵ)ceiλQm,n(y)ykr2N[r1ykr1yl][r2ykr2η(y)][r3ykr3Ωk(y)][r4ykr4[1α(ϵ1y)]]𝑑y\lambda^{-N}\int_{\Gamma_{k}\cap B(0,\epsilon)^{c}}e^{i\lambda Q_{m,n}(y)}y_{k}^{r-2N}\left[\frac{\partial^{r_{1}}}{\partial y_{k}^{r_{1}}}y^{l}\right]\left[\frac{\partial^{r_{2}}}{\partial y_{k}^{r_{2}}}\eta(y)\right]\left[\frac{\partial^{r_{3}}}{\partial y_{k}^{r_{3}}}\Omega_{k}(y)\right]\left[\frac{\partial^{r_{4}}}{\partial y_{k}^{r_{4}}}[1-\alpha(\epsilon^{-1}y)]\right]dy

    for r1,r2,r3,r40r_{1},r_{2},r_{3},r_{4}\geq 0, r1+r2+r3+r4=rNr_{1}+r_{2}+r_{3}+r_{4}=r\leq N. We note that r3Ωkykr3\frac{\partial^{r_{3}}\Omega_{k}}{\partial y_{k}^{r_{3}}} is a homogeneous function of degree r3-r_{3}. When |l|N<n|l|-N<-n, the term above is bounded by a constant multiple of

    λNϵr4ΓkB(0,ϵ)c|y|r2N+|l|r1r3r2ηykr2Lr4αykr4L𝑑yλNϵr4ϵr2N+|l|r1r3+nηCr2αCr4λNϵ|l|2N+rr1r3r4+nηCr2αCr4.\begin{split}&\lambda^{-N}\epsilon^{-r_{4}}\int_{\Gamma_{k}\cap B(0,\epsilon)^{c}}|y|^{r-2N+|l|-r_{1}-r_{3}}\left\lVert\frac{\partial^{r_{2}}\eta}{\partial y_{k}^{r_{2}}}\right\rVert_{L^{\infty}}\left\lVert\frac{\partial^{r_{4}}\alpha}{\partial y_{k}^{r_{4}}}\right\rVert_{L^{\infty}}dy\\ \lesssim&\lambda^{-N}\epsilon^{-r_{4}}\epsilon^{r-2N+|l|-r_{1}-r_{3}+n}\left\lVert\eta\right\rVert_{C^{r_{2}}}\left\lVert\alpha\right\rVert_{C^{r_{4}}}\\ \lesssim&\lambda^{-N}\epsilon^{|l|-2N+r-r_{1}-r_{3}-r_{4}+n}\left\lVert\eta\right\rVert_{C^{r_{2}}}\left\lVert\alpha\right\rVert_{C^{r_{4}}}.\end{split}

    Then we obtain (38) by setting ϵ=λ12\epsilon=\lambda^{-\frac{1}{2}}. AkA_{k} depends on ηCN\left\lVert\eta\right\rVert_{C^{N}}.

  2. b.

    The proof for b) is similar to that for a). We use the same cone Γk\Gamma_{k} and functions Ωk\Omega_{k}, α\alpha. It suffices to show

    (41) |ΓkeiλQm,n(y)g(y)Ωk(y)𝑑y|BN,kλN\left|\int_{\Gamma_{k}}e^{i\lambda Q_{m,n}(y)}g(y)\Omega_{k}(y)dy\right|\leq B_{N,k}\lambda^{-N}\

    for a BN,k>0B_{N,k}>0. If 2ϵ<δ2\epsilon<\delta,

    ΓkeiλQm,n(y)g(y)Ωk(y)𝑑y=ΓkB(0,ϵ)ceiλQm,n(y)g(y)Ωk(y)[1α(ϵ1y)]𝑑y.\int_{\Gamma_{k}}e^{i\lambda Q_{m,n}(y)}g(y)\Omega_{k}(y)dy=\int_{\Gamma_{k}\cap B(0,\epsilon)^{c}}e^{i\lambda Q_{m,n}(y)}g(y)\Omega_{k}(y)[1-\alpha(\epsilon^{-1}y)]dy.

    Then, we apply integration by parts NN times to obtain

    (42) ΓkB(0,ϵ)ceiλQm,n(y)(tD)N{g(y)Ωk(y)[1α(ϵ1y)]}dy.\int_{\Gamma_{k}\cap B(0,\epsilon)^{c}}e^{i\lambda Q_{m,n}(y)}(^{t}D)^{N}\left\{g(y)\Omega_{k}(y)[1-\alpha(\epsilon^{-1}y)]\right\}dy.

    After we apply (40) and the product rule of the derivative to expand (42), we obtain a summation of terms where a term, ignoring the constant, is

    λNΓkeiλQm,n(y)ykr2N[r1ykr1g(y)][r2ykr2Ωk(y)][r3ykr3[1α(ϵ1y)]]𝑑y\lambda^{-N}\int_{\Gamma_{k}}e^{i\lambda Q_{m,n}(y)}y_{k}^{r-2N}\left[\frac{\partial^{r_{1}}}{\partial y_{k}^{r_{1}}}g(y)\right]\left[\frac{\partial^{r_{2}}}{\partial y_{k}^{r_{2}}}\Omega_{k}(y)\right]\left[\frac{\partial^{r_{3}}}{\partial y_{k}^{r_{3}}}[1-\alpha(\epsilon^{-1}y)]\right]dy

    for r1,r2,r30r_{1},r_{2},r_{3}\geq 0, r1+r2+r3=rNr_{1}+r_{2}+r_{3}=r\leq N. We note that r2Ωkykr2\frac{\partial^{r_{2}}\Omega_{k}}{\partial y_{k}^{r_{2}}} is a homogeneous function of degree r2-r_{2}. When N<n-N<-n, the term above is bounded by a constant multiple of

    λNϵr3ΓkB(0,ϵ)c|y|r2Nr2r1gykr1Lr3αykr3L𝑑yλNϵr3ϵr2Nr2+ngCr1αCr3λNϵr2Nr2r3+ngCr1αCr3.\begin{split}&\lambda^{-N}\epsilon^{-r_{3}}\int_{\Gamma_{k}\cap B(0,\epsilon)^{c}}|y|^{r-2N-r_{2}}\left\lVert\frac{\partial^{r_{1}}g}{\partial y_{k}^{r_{1}}}\right\rVert_{L^{\infty}}\left\lVert\frac{\partial^{r_{3}}\alpha}{\partial y_{k}^{r_{3}}}\right\rVert_{L^{\infty}}dy\\ \lesssim&\lambda^{-N}\epsilon^{-r_{3}}\epsilon^{r-2N-r_{2}+n}\left\lVert g\right\rVert_{C^{r_{1}}}\left\lVert\alpha\right\rVert_{C^{r_{3}}}\\ \lesssim&\lambda^{-N}\epsilon^{r-2N-r_{2}-r_{3}+n}\left\lVert g\right\rVert_{C^{r_{1}}}\left\lVert\alpha\right\rVert_{C^{r_{3}}}.\end{split}

    Then, we obtain (41) by setting ϵ=δ3\epsilon=\frac{\delta}{3}. ∎

Theorem 4.6.

[18, Chapter 8, Proposition 6]

  1. a.

    Let Qm,n(y)Q_{m,n}(y) be defined as in (30). Let

    Im(λ;ψ):=neiλQm,n(y)ψ(y)𝑑y,I_{m}(\lambda;\psi):=\int_{\mathbb{R}^{n}}e^{i\lambda Q_{m,n}(y)}\psi(y)dy,

    where ψ\psi is supported in a small neighborhood of 0, and λ0>1\lambda_{0}>1. For λλ0\lambda\geq\lambda_{0},

    (43) |Im(λ;ψ)(1)nm2(πi)n2ψ(0)λn2|Dλn+12,\left|I_{m}(\lambda;\psi)-(-1)^{\frac{n-m}{2}}(\pi i)^{\frac{n}{2}}\psi(0)\lambda^{-\frac{n}{2}}\right|\leq D\lambda^{-\frac{n+1}{2}},

    where DD depends on λ0\lambda_{0}, the size of spt ψ\text{spt }\psi, and ψCn+3\left\lVert\psi\right\rVert_{C^{n+3}}.

  2. b.

    Let ϕ,ψC(n)\phi,\psi\in C^{\infty}(\mathbb{R}^{n}). Suppose that ϕ\phi has only one non-degnerate critical point z0z_{0} in spt ψ\text{spt }\psi and ϕ(z0)=0\phi(z_{0})=0. Let

    I(λ;ϕ,ψ):=neiλϕ(z)ψ(z)𝑑z,I(\lambda;\phi,\psi):=\int_{\mathbb{R}^{n}}e^{i\lambda\phi(z)}\psi(z)dz,

    λ0>1\lambda_{0}>1, and c=detΔϕ(z0)c=\det\Delta\phi(z_{0}). For λλ0\lambda\geq\lambda_{0}, there exist 0mn0\leq m\leq n and D=D(λ0,ϕ,ψ)>0D=D(\lambda_{0},\phi,\psi)>0, such that

    |I(λ;ϕ,ψ)(1)nm2(2πi)n2c12ψ(z0)λn2|D(λ0,ϕ,ψ)|λ|n+12.\left|I(\lambda;\phi,\psi)-(-1)^{\frac{n-m}{2}}(2\pi i)^{\frac{n}{2}}c^{-\frac{1}{2}}\psi(z_{0})\lambda^{-\frac{n}{2}}\right|\leq D(\lambda_{0},\phi,\psi)|\lambda|^{-\frac{n+1}{2}}.

    The error D(λ0,ϕ,ψ)D(\lambda_{0},\phi,\psi) depends on λ0\lambda_{0}, cc, the size of spt ψ\text{spt }\psi, and ϕCn+6\left\lVert\phi\right\rVert_{C^{n+6}}, ψCn+3\left\lVert\psi\right\rVert_{C^{n+3}}.

Proof.
  1. a.

    Step 1: We have

    eiλx2ex2𝑑x=(1iλ)12ex2𝑑x,\int_{-\infty}^{\infty}e^{i\lambda x^{2}}e^{-x^{2}}dx=(1-i\lambda)^{-\frac{1}{2}}\int_{-\infty}^{\infty}e^{-x^{2}}dx,

    and ex2𝑑x=π\int_{-\infty}^{\infty}e^{-x^{2}}dx=\sqrt{\pi}. We can fix the principal branch of z12z^{-\frac{1}{2}} in the complex plane slit along the negative real-axis. Therefore,

    (44) neiλQm,n(y)e|y|2𝑑y=(j=1meiλyj2eyj2𝑑yj)(j=m+1neiλyj2eyj2𝑑yj)=πn2(1iλ)m2(1+iλ)nm2=πn2λn2(λ1i)m2(λ1+i)nm2.\begin{split}\int_{\mathbb{R}^{n}}e^{i\lambda Q_{m,n}(y)}e^{-|y|^{2}}dy=&\left(\prod_{j=1}^{m}\int_{\mathbb{R}}e^{i\lambda y_{j}^{2}}e^{-y_{j}^{2}}dy_{j}\right)\left(\prod_{j=m+1}^{n}\int_{\mathbb{R}}e^{-i\lambda y_{j}^{2}}e^{-y_{j}^{2}}dy_{j}\right)\\ =&\pi^{\frac{n}{2}}(1-i\lambda)^{-\frac{m}{2}}(1+i\lambda)^{-\frac{n-m}{2}}\\ =&\pi^{\frac{n}{2}}\lambda^{-\frac{n}{2}}(\lambda^{-1}-i)^{-\frac{m}{2}}(\lambda^{-1}+i)^{-\frac{n-m}{2}}.\end{split}

    We write the power series expansion of fm(w)=(wi)m2(w+i)nm2f_{m}(w)=(w-i)^{-\frac{m}{2}}(w+i)^{-\frac{n-m}{2}} at 0 as k=0akwk,\sum_{k=0}^{\infty}a_{k}w^{k}, where a0=(1)nm2in2a_{0}=(-1)^{\frac{n-m}{2}}i^{\frac{n}{2}}. Let γ\gamma be a line segment from 0 to λ1\lambda^{-1} and bm=sup|w|=λ01|fm(w)|b_{m}=\sup_{|w|=\lambda_{0}^{-1}}|f_{m}^{\prime}(w)|. Then for λλ0>1\lambda\geq\lambda_{0}>1, the error of approximation by the constant term is bounded by

    (45) |fm(λ1)a0||γfm(w)𝑑w||γ|sup|w|=λ1|fm(w)|bmλ1.\begin{split}|f_{m}(\lambda^{-1})-a_{0}|\leq&\left|\int_{\gamma}f_{m}^{\prime}(w)dw\right|\\ \leq&|\gamma|\sup_{|w|=\lambda^{-1}}|f_{m}^{\prime}(w)|\\ \leq&b_{m}\lambda^{-1}.\end{split}

    The integral form of the Taylor remainder and the maximum modulus principle are used. Putting everything together, for λλ0\lambda\geq\lambda_{0},

    (46) |neiλQm,n(y)e|y|2𝑑ya0πn2λn2|\displaystyle\left|\int_{\mathbb{R}^{n}}e^{i\lambda Q_{m,n}(y)}e^{-|y|^{2}}dy-a_{0}\pi^{\frac{n}{2}}\lambda^{-\frac{n}{2}}\right|
    =\displaystyle= |πn2λn2(λ1i)m2(λ1+i)nm2a0πn2λn2|\displaystyle\left|\pi^{\frac{n}{2}}\lambda^{-\frac{n}{2}}(\lambda^{-1}-i)^{-\frac{m}{2}}(\lambda^{-1}+i)^{-\frac{n-m}{2}}-a_{0}\pi^{\frac{n}{2}}\lambda^{-\frac{n}{2}}\right| by (44)
    =\displaystyle= πn2λn2|(λ1i)m2(λ1+i)nm2a0|\displaystyle\pi^{\frac{n}{2}}\lambda^{-\frac{n}{2}}\left|(\lambda^{-1}-i)^{-\frac{m}{2}}(\lambda^{-1}+i)^{-\frac{n-m}{2}}-a_{0}\right|
    \displaystyle\leq bmπn2λn+22\displaystyle b_{m}\pi^{\frac{n}{2}}\lambda^{-\frac{n+2}{2}} by (45).\displaystyle\text{ by (\ref{eq_fm_a0_d})}.

    Step 2: To obtain (43), we write e|y|2ψ(y)=ψ(0)+y=1nyjRj(y)e^{|y|^{2}}\psi(y)=\psi(0)+\sum_{y=1}^{n}y_{j}R_{j}(y), where RjC0(n)R_{j}\in C_{0}^{\infty}(\mathbb{R}^{n}). Let ψ¯C0(n)\bar{\psi}\in C^{\infty}_{0}(\mathbb{R}^{n}), with ψ¯(y)=1\bar{\psi}(y)=1 on the support of ψ\psi. To apply the results from the previous steps, we write

    neiλQm,n(y)ψ(y)𝑑y=eiλQm,n(y)e|y|2[e|y|2ψ(y)]ψ¯(y)𝑑y=neiλQm,n(y)e|y|2[ψ(0)+j=1nyjRj(y)]ψ¯(y)𝑑y=eiλQm,n(y)e|y|2ψ(0)ψ¯(y)𝑑y+j=1neiλQm,n(y)yje|y|2Rj(y)ψ¯(y)𝑑y=ψ(0)eiλQm,n(y)e|y|2𝑑y+ψ(0)eiλQm,n(y)e|y|2[1ψ¯(y)]𝑑y+j=1neiλQm,n(y)yje|y|2Rj(y)ψ¯(y)𝑑y.\begin{split}&\int_{\mathbb{R}^{n}}e^{i\lambda Q_{m,n}(y)}\psi(y)dy\\ =&\int e^{i\lambda Q_{m,n}(y)}e^{-|y|^{2}}[e^{|y|^{2}}\psi(y)]\bar{\psi}(y)dy\\ =&\int_{\mathbb{R}^{n}}e^{i\lambda Q_{m,n}(y)}e^{-|y|^{2}}\left[\psi(0)+\sum_{j=1}^{n}y_{j}R_{j}(y)\right]\bar{\psi}(y)dy\\ =&\int e^{i\lambda Q_{m,n}(y)}e^{-|y|^{2}}\psi(0)\bar{\psi}(y)dy+\sum_{j=1}^{n}\int e^{i\lambda Q_{m,n}(y)}y_{j}e^{-|y|^{2}}R_{j}(y)\bar{\psi}(y)dy\\ =&\psi(0)\int e^{i\lambda Q_{m,n}(y)}e^{-|y|^{2}}dy+\psi(0)\int e^{i\lambda Q_{m,n}(y)}e^{-|y|^{2}}[1-\bar{\psi}(y)]dy\\ &+\sum_{j=1}^{n}\int e^{i\lambda Q_{m,n}(y)}y_{j}e^{-|y|^{2}}R_{j}(y)\bar{\psi}(y)dy.\end{split}

    Then we obtain (43) by applying (46) to the first integral, (37) to the second integral, and (36) to each term in the summation. We note that the error DD depends on λ0\lambda_{0}, the size of the support of ψ\psi, and ψCn+3\left\lVert\psi\right\rVert_{C^{n+3}}.

  2. b.

    Let VV be a neighborhood of z0z_{0} obtained from applying Morse’s lemma (Lemma 4.3) to ϕ\phi, and we write ψ=ψ1+ψ2\psi=\psi_{1}+\psi_{2}, where spt ψ1Vspt ψ\text{spt }\psi_{1}\subset V\cap\text{spt }\psi, and spt ψ2\text{spt }\psi_{2} does not contain z0z_{0}. Then, we write

    neiλϕ(z)ψ(z)𝑑z=neiλϕ(z)ψ1(z)𝑑z+neiλϕ(z)ψ2(z)𝑑z.\int_{\mathbb{R}^{n}}e^{i\lambda\phi(z)}\psi(z)dz=\int_{\mathbb{R}^{n}}e^{i\lambda\phi(z)}\psi_{1}(z)dz+\int_{\mathbb{R}^{n}}e^{i\lambda\phi(z)}\psi_{2}(z)dz.
    • By Morse’s lemma (Lemma 4.3), there exists a diffeomorphism τ:VU\tau:V\to U, where VV is a neighborhood of z0z_{0} in the zz-space, and UU is a neighborhood of 0 in the yy-space, such that τ(z0)=0\tau(z_{0})=0 and

      ϕ(z)=Qm,n(τ(z)).\phi(z)=Q_{m,n}(\tau(z)).

      By a change of variables z=τ1(y)z=\tau^{-1}(y),

      neiλϕ(z)ψ1(z)𝑑z=eiλQm,n(y)ψ1(τ1(y))|det𝐉τ1(y)|𝑑y=Im(λ;(ψ1τ1)|det𝐉τ1|).\begin{split}\int_{\mathbb{R}^{n}}e^{i\lambda\phi(z)}\psi_{1}(z)dz=&\int e^{i\lambda Q_{m,n}(y)}\psi_{1}(\tau^{-1}(y))|\det{\bf J}\tau^{-1}(y)|dy\\ =&I_{m}(\lambda;(\psi_{1}\circ\tau^{-1})\cdot|\det{\bf J}\tau^{-1}|).\end{split}

      We can apply part a) to the integral above. DD depends on λ0\lambda_{0}, the size of support of ψ1τ1\psi_{1}\circ\tau^{-1}, and the Cn+3C^{n+3} norms of (ψ1τ1)|det𝐉τ1|(\psi_{1}\circ\tau^{-1})\cdot|\det{\bf J}\tau^{-1}|. We note that the LL^{\infty} norm of the kthk^{\text{th}} partial derivative (0kn+30\leq k\leq n+3) of (ψ1τ1)|det𝐉τ1|(\psi_{1}\circ\tau^{-1})\cdot|\det{\bf J}\tau^{-1}| can be bounded by Cn+3C^{n+3} norms of ψ\psi, |det𝐉τ||\det{\bf J}\tau|, and inf{|det𝐉τ(z)|zsptψ}\inf\{|\det{\bf J}\tau(z)|z\in\text{spt}\psi\}. We note that |det𝐉τ(0)|2=c2n|\det{\bf J}\tau(0)|^{2}=\frac{c}{2^{n}} by (31), and we can bound the error in terms of ϕ\phi with the fact that det𝐉τCn+3(V)\left\lVert\det{\bf J}\tau\right\rVert_{C^{n+3}(V)} and inf{|det𝐉τ(z)|zsptψ}\inf\{|\det{\bf J}\tau(z)|z\in\text{spt}\psi\} depend on ϕCn+5(V)\left\lVert\phi\right\rVert_{C^{n+5}(V)}, cc, and the size of spt ψ1\text{spt }\psi_{1} from the discussion of Morse’s lemma in Section 4.4.

    • For the second integral, we apply Proposition 4.4. ∎

References

  • [1] Christian Bluhm. Random recursive construction of Salem sets. Ark. Mat., 34(1):51–63, 1996.
  • [2] Manfredo P. d. Carmo. Differential geometry of curves and surfaces. Prentice-Hall, Englewood Cliffs, NJ, 1976.
  • [3] Fredrik Ekström, Tomas Persson, and Jörg Schmeling. On the Fourier dimension and a modification. J. Fractal Geom., 2(3):309–337, 2015.
  • [4] Kenneth Falconer. Fractal geometry : Mathematical foundations and applications. John Wiley and Sons, Incorporated, Hoboken, NJ, third edition, 2014.
  • [5] Jonathan M. Fraser, Terence L. J. Harris, and Nicholas G. Kroon. On the Fourier dimension of (d,k)(d,k)-sets and Kakeya sets with restricted directions. Math. Z., 301(3):2497–2508, 2022.
  • [6] Robert Fraser and Kyle Hambrook. Explicit Salem sets in n\mathbb{R}^{n}. Adv. Math., 416:Paper No. 108901, 23, 2023.
  • [7] Kanghui Guo. On the pp-thin problem for hypersurfaces of n\mathbb{R}^{n} with zero Gaussian curvature. Canad. Math. Bull., 36(1):64–73, 1993.
  • [8] Philip Hartman. On isometric immersions in Euclidean space of manifolds with non-negative sectional curvatures. Trans. Amer. Math. Soc., 115:94–109, 1965.
  • [9] Edmund Hlawka. über Integrale auf konvexen Körpern. I. Monatsh. Math., 54:1–36, 1950.
  • [10] Jean-Pierre Kahane. Some random series of functions, volume 5 of Cambridge Studies in Advanced Mathematics. Cambridge University Press, Cambridge, UK, second edition, 1985.
  • [11] Robert Kaufman. On the theorem of Jarník and Besicovitch. Acta Arith., 39(3):265–267, 1981.
  • [12] Izabella Laba and Malabika Pramanik. Arithmetic progressions in sets of fractional dimension. Geometric and Functional Analysis, 19(2):429–456, 2009.
  • [13] Yiyu Liang and Malabika Pramanik. Fourier dimension and avoidance of linear patterns. Advances in Mathematics, 399:108252, 2022.
  • [14] Walter Littman. Fourier transforms of surface-carried measures and differentiability of surface averages. Bull. Amer. Math. Soc., 69:766–770, 1963.
  • [15] Pertti Mattila. Fourier analysis and Hausdorff dimension, volume 150. Cambridge University Press, Cambridge, UK, 1 edition, 2015.
  • [16] Walter Rudin. Real and complex analysis. McGraw-Hill, New York, NY, 3rd edition, 1987.
  • [17] Raphaël Salem. On singular monotonic functions whose spectrum has a given Hausdorff dimension. Ark. Mat., 1:353–365, 1951.
  • [18] Elias Stein. Harmonic Analysis Real-Variable Methods, Orthogonality, and Oscillatory Integrals, volume 43 of Princeton Mathematical Series. Princeton University Press, Princeton, NJ, 1993.
  • [19] John Thorpe. Elementary Topics in Differential Geometry. Undergraduate Texts in Mathematics. Springer, New York, NY, 1979.
  • [20] Vitaly Ushakov. Developable surfaces in Euclidean space. J. Austral. Math. Soc. Ser. A, 66(3):388–402, 1999.
  • [21] Junjie Zhu. Fourier dimension of conical and cylindrical hypersurfaces, 2024.

Department of Mathematics, 1984 Mathematics Road, University of British Columbia, Vancouver, BC Canada V6T 1Z2

E-mail address: [email protected]