On sharp Strichartz estimate for hyperbolic Schrödinger equation on \(\mathbb{T}^3\)


Abstract

We prove the sharp Strichartz estimate for hyperbolic Schrödinger equation on \(\mathbb{T}^3\) via an incidence geometry approach. As application, we obtain optimal local well-posedness of nonlinear hyperbolic Schrödinger equations.

1 2 3

1 Introduction↩︎

The question of Strichartz estimates for Schrödinger equation on tori was first addressed by Bourgain [1]. Later, Bourgain-Demeter [2] proved the full range Strichartz estimates with \(N^\varepsilon\) loss by the Fourier decoupling method: \[\label{Bourgain-Demeter32Strichartz32estimate32by32decoupling} \| P_N \mathrm{e}^{\mathrm{i}t\Delta} \phi \|_{{L_{t,x}^p([0,1]\times \mathbb{T}^d)}} \lesssim_\varepsilon N^{\frac{d}{2} - \frac{d+2}{p}+\varepsilon } \| \phi \|_{L^2_x(\mathbb{T}^d)}, \quad \forall p\geq p_{d},\;\forall \varepsilon>0,\tag{1}\] where \(P_N\) denotes the Littlewood-Paley projection operator to frequency \(N\) and \(p_d = 2(d+2)/d\). Decoupling theorems are powerful and robust tools in Fourier analysis, but the \(N^\varepsilon\) loss is inherent in the proof of decoupling theorems. The loss in 1 was removed by Killip-Vişan [3] for \(p>p_d\). Recently, Herr-Kwak [4] proved the sharp endpoint point \(L^4\) estimate on \(\mathbb{T}^2\) \[\| P_N \mathrm{e}^{\mathrm{i}t\Delta} \phi \|_{L^4_{t,x}([0,1]\times \mathbb{T}^2)} \lesssim (\log N)^{1/4} \| \phi \|_{L^2_x(\mathbb{T}^2)},\] which implies global existence of solutions to the cubic (mass-critical) nonlinear Schrödinger equation in \(H^s(\mathbb{T}^2)\) for any \(s>0\).

For the hyperbolic Schrödinger equation, it shares the same Strichatz estimates as the elliptic one in the Euclidean case, but there is a difference on tori. In [5], Bourgain-Demeter proved that \[\| P_N \mathrm{e}^{\mathrm{i}t\Box} \phi \|_{L^p_{t,x}([0,1]\times \mathbb{T}^d)} \lesssim_\varepsilon N^{\mu_{d,v}(p)+\varepsilon} \| \phi \|_{L^2_x(\mathbb{T}^d)}, \quad \forall p\geq 2,\;\forall \varepsilon>0, \label{est:bourgain}\tag{2}\]

where \(\Box = \partial^2_{x_1}+\dots+\partial^2_{x_v}-\partial^2_{x_{v+1}}-\dots-\partial^2_{x_d}\), \(v\leq d/2,\) and \[\mu_{d,v}(p) = \max\left\{ \frac{d}{2} - \frac{d+2}{p}, \frac{v}{2}-\frac{v}{p}\right\}.\label{powerindex}\tag{3}\] The factor \(N^{v(\frac{1}{2}-\frac{1}{p})}\) is due to that the hyperbolic paraboloid contains a vector subspace of dimension \(v\). It’s a natural question to ask whether the \(N^\varepsilon\) loss can be removed.

In this paper, we consider the case \(d=3\) and prove the sharp Strichartz estimate for hyperbolic Schrödinger equation without \(N^\varepsilon\) loss. With the notations \(\Box = \partial^2_{x_1} - \partial^2_{x_2} - \partial^2_{x_3}\) and \(\mu(p)=\mu_{3,1}(p)=\max\{\frac{3}{2}-\frac{5}{p}, \frac{1}{2}-\frac{1}{p}\}\), our main result reads as follows:

Theorem 1. For \(\phi \in L^2(\mathbb{T}^3)\), we have that \[\| P_N \mathrm{e}^{\mathrm{i}t\Box} \phi \|_{L^p_{t,x}([0,1]\times \mathbb{T}^3)} \lesssim N^{\mu(p)} \| \phi \|_{L^2_x(\mathbb{T}^3)},\quad \forall p\geq 2. \label{main-est}\qquad{(1)}\]

Remark 2. By invoking interpolation with \(L^\infty\) and \(L^2\) estimates, it suffices to prove Theorem 1 for \(p=4\), i.e. \[\label{est:L4} \| P_N \mathrm{e}^{\mathrm{i}t\Box}{\phi}\|_{{L_{t,x}^4([0,1]\times \mathbb{T}^3)}} \lesssim N^{1/4} \| \phi \|_{{L^2_x(\mathbb{T}^3)}}.\qquad{(2)}\]

Figure 1: image.

Remark 3. Due to the Galilean invariance of solutions to the linear hyperbolic Schrödinger equation, estimate ?? can be reformulated as \[\| P_S \mathrm{e}^{\mathrm{i}t\Box} \phi \|_{L^4_{t,x}([0,1]\times \mathbb{T}^3)} \lesssim \operatorname{diam}(S)^{1/4} \| \phi \|_{L^2_x(\mathbb{T}^3)}\] for any bounded set \(S\subset \mathbb{Z}^3\). Naively, we ask if this can be replaced by a bound depending only on \(\# S\), as in the work of Herr-Kwak [4]. We will construct examples in Section 2.7 to show there is no efficient bound except for the trivial one \((\#S)^{1/4}\).

As application of Theorem 1, we consider the Cauchy problem for hyperbolic nonlinear Schrödinger equations (HNLS). HNLS arise in many physics contexts, such as plasma waves [6][9] and gravity water waves [10][13]. In particular, the 3d cubic HNLS appear in the study of optical self-focusing of short light pulses in nonlinear media [9], and it is considered one of the canonical NLS equations in 3d [14]. We refer the readers to the survey paper by Saut-Wang [15] for more details.

The Cauchy problem of two-dimensional periodic HNLS with cubic nonlinearity \[\label{cubic32HNLS32on32T2} (\mathrm{i}\partial_t + \partial_{x_1}^2-\partial_{x_2}^2 ) u = |u|^2u,\quad (t,x) \in \mathbb{R} \times \mathbb{T}^2,\tag{4}\] has been considered by Godet-Tzvetkov [16] and Wang [17]. They both established \(L^4\) Strichartz estimate with \(1/4\)-derivative loss, using different methods. Besides, Wang [17] used the Strichartz estimate to prove that the Cauchy problem of 4 is locally well-posed in \(H^{s}(\mathbb{T}^2)\) for \(s>1/2\) while it’s ill-posed for \(s<1/2\) in the sense that the solution map is not \(C^3\) continuous in \(H^s(\mathbb{T}^2)\) even for small data. The recent work [18] established the sharp unconditional well-posedness in Fourier–Lebesgue spaces (modulo the endpoint case) for 4 and [19] considered HNLS with all odd power nonlinearities on \(\mathbb{R}\times\mathbb{T}\) and proved sharp local well-posedness.

Here we study the three-dimensional periodic HNLS \[\label{HNLS32on32T3} \mathrm{i}\partial_t u + \Box u = \pm |u|^{2k}u,\quad (t,x) \in \mathbb{R} \times \mathbb{T}^3,\tag{5}\] where \(k\) is a positive integer. The Cauchy problem for 5 was posed by Saut-Wang [15]. In the Euclidian case, the equation 5 enjoys the scaling symmetry, which leaves the critical Sobolev norm \(\| \cdot\|_{\dot{H}^{s_c}(\mathbb{R}^3)}\) invariant for \(s_c=\frac{3}{2}-\frac{1}{k}\). Although in the periodic case we don’t have this natural scaling symmetry, the notation of critical Sobolev index provides us heuristics. We have the following results of local well-posedness.

Theorem 4. For \(k\geq2\), the Cauchy problem of 5 is locally well-posed in \(H^{s_c}(\mathbb{T}^3)\). For \(k= 1\), the Cauchy problem of 5 is locally well-posed in \(H^{s }(\mathbb{T}^3)\) for any \(s>s_c=1/2\).

Theorem 5. For \(k=1\) and \(T>0\) be arbitrarily small. Assume the data-to-solution map \(u_0 \mapsto u(\cdot)\) associated with 5 on smooth data extends continuously to a map from \(H^{1/2}(\mathbb{T}^3)\) to \(C([0,T];H^{1/2}(\mathbb{T}^3))\). Then this map will not be \(C^3\) at the origin.

The outline of this paper is as follows. In Section 2 we prove Theorem 1. We take the Fourier transform and reduce the \(L^4\) estimate ?? to a counting argument for parallelograms with vertices in given sets. We distinguish two cases depending on whether the sides of the parallelograms lie on a cone.

In Section 3 we prove Theorem 4 based on a multilinear estimate and contraction mapping argument. Then we construct specific solutions to prove Theorem 5.

2 Strichartz Estimate↩︎

2.1 Notations↩︎

We denote \(A\lesssim B\) or \(A=O(B)\) if \(A\leq CB\) holds for some constant \(C>0\) independent with \(A\) and \(B\). We write \(A\approx B\) if \(A\lesssim B\) and \(B\lesssim A\). We denote \(\#S\) the cardinality of finite set \(S\). For integers \(a,b\), we denote \(a|b\) if \(a^{-1}b\in\mathbb{Z}\). For \(f \in L^2(\mathbb{T}^3)\), the Fourier coefficients of \(f\) are given by \[\hat{f}(k) = \int_{\mathbb{T}^3} f(x) \mathrm{e}^{-2\pi\mathrm{i}k\cdot x} \mathop{}\!\mathrm{d}x, \quad k\in\mathbb{Z}^3,\] and the Fourier series of \(f\) is \[f(x) = \sum_{k\in\mathbb{Z}^3} \hat{f}(k) \mathrm{e}^{2\pi\mathrm{i}k\cdot x}.\] The series converges in \(L^2(\mathbb{T}^3)\) sense. For any subset \(S\subset \mathbb{Z}^3\), we denote \(P_S\) for the Fourier multiplier with symbol \(\chi_S\), i.e. \[P_Sf = \sum_{k\in S} \hat{f}(k) \mathrm{e}^{ 2\pi\mathrm{i}k\cdot x}.\] In this paper, \(N\) will always be a dyadic integer, i.e. \(N=2^n\) for some \(n\in\mathbb{N}\). For \(S = {\{ k \in \mathbb{Z}^3 \mid N \leq |k| < 2N \}}\) we simply write \(P_S\) as \(P_N\), and \[P_{\leq N} f= \sum_{M\leq N,\;M \text{ dyadic}} P_Mf, \quad P_{>N}f = f-P_{\leq N}f.\] For \(s\in\mathbb{R}\), the Sobolev space \(H^s(\mathbb{T}^3)\) is the set of all functions \(f\in L^2(\mathbb{T}^3)\) such that the norm \[\| f \|_{H^s(\mathbb{T}^3) }: = \left( \sum_{k\in\mathbb{Z}^3} \left<k\right>^{2s} |\hat{f}(k)|^2 \right)^{1/2}\] is finite, where \(\left< k \right> = \sqrt{1+|k|^2}\).

2.2 Facts from incidence geometry↩︎

We need the Szemeŕedi-Trotter theorem from incidence geometry. An incidence is defined as a point-curve pair so that the point lies on the curve. The problem is to bound the number of incidence that are possible for certain classes of curves.

Theorem 6 ([20], [21]Points-lines incidences). Let \(\mathcal{P}\) be a set of \(n\) points and \(\mathcal{L}\) be a set of \(m\) lines. Then the number of incidences between \(\mathcal{P}\) and \(\mathcal{L}\) is \(O(n^{2/3}m^{2/3}+m+n)\).

Corollary 7 ([21]). Let \(\mathcal{P}\) be a set of \(n\) points and \(\mathcal{L}\) be a set of lines. Suppose that every line in \(\mathcal{L}\) contains at least \(k\geq2\) points of \(\mathcal{P}\). Then the number of incidences between \(\mathcal{P}\) and \(\mathcal{L}\) is \(O(n^{2}/k^2 + n)\).

We also need the following upper bound on points-circles incidences.

Theorem 8 (Points-circles incidences on sphere). Let \(\mathcal{P}\) be a set of \(n\) points on the unit sphere and \(\mathcal{C}\) be a set of \(m\) great circles on the unit sphere. Then the number of incidences between \(\mathcal{P}\) and \(\mathcal{C}\) is \(O(n^{2/3}m^{2/3}+m+n)\).

Proof. It suffices to consider the incidences on a half sphere, since \(\mathbb{S}^2\) can be covered by eight half spheres. We define the map \(\Psi\colon \{(x_1,x_2,x_3)\in \mathbb{S}^2 \mid x_3>0\} \to \mathbb{R}^2\), \(\Psi(x_1,x_2,x_3)=(x_1/x_3,x_2/x_3)\). It’s easy to see that \(\Psi\) is a bijection, hence it preserves the number of incidences. Besides, \(\Psi\) maps the intersection of great circles and half sphere into lines on the plane, the conclusion follows from Theorem 6. ◻

Remark 9. The same points-circles incidences estimate on the sphere holds true if no three circles intersect in two common points; for example, if all circles are congruent and are not great circles on the sphere, see [22] for more information.

2.3 Preparation↩︎

We will focus on the proof of the \(L^4\) Strichartz estimate ?? . We denote \(A\) the diagonal matrix \(\operatorname{diag}\{1,-1,-1\}\), and \(h(\xi) = \xi \cdot A\xi\) denotes the inner product of \(\xi\in \mathbb{Z}^3\) and \(A\xi\). With these notations, we may write \[\mathrm{e}^{\mathrm{i}t\Box}\phi(x) = \sum_{\xi\in\mathbb{Z}^3}\hat{\phi}(\xi) \mathrm{e}^{2\pi\mathrm{i}(x\cdot \xi + th(\xi))}.\] As a result, its \(L^4\) norm is given by \[\begin{align} & \int_{[0,1]\times \mathbb{T}^3} | \mathrm{e}^{\mathrm{i}t\Box}\phi(x)|^4 \mathop{}\!\mathrm{d}t\mathop{}\!\mathrm{d}x \\=&{} \sum_{\xi_1,\xi_2,\xi_3,\xi_4\in\mathbb{Z}^3} \overline{\hat{\phi}(\xi_1)}\hat{\phi}(\xi_2) \overline{\hat{\phi}(\xi_3)} \hat{\phi}(\xi_4) \int_{[0,1]\times \mathbb{T}^3} \mathrm{e}^{2\pi\mathrm{i}(x\cdot\sum_{i=1}^4(-1)^i\xi_i + t\sum_{i=1}^4 (-1)^ih(\xi_i) )} \mathop{}\!\mathrm{d}t\mathop{}\!\mathrm{d}x \\ =& \sum_{(\xi_1,\xi_2,\xi_3,\xi_4)\in\mathcal{Q}} \overline{\hat{\phi}(\xi_1)}\hat{\phi}(\xi_2) \overline{\hat{\phi}(\xi_3)} \hat{\phi}(\xi_4), \label{formula:L4} \end{align}\tag{6}\] where \[\begin{align} \mathcal{Q} &= \Big\{ (\xi_1,\xi_2,\xi_3,\xi_4) \in \mathbb{Z}^{3\times4} \Bigm| \sum_{i=1}^4(-1)^i\xi_i=0,\;\sum_{i=1}^4(-1)^ih(\xi_i)=0 \Big\} \\& = \Big\{ (\xi_1,\xi_2,\xi_3,\xi_4) \in \mathbb{Z}^{3\times4} \Bigm| \sum_{i=1}^4(-1)^i\xi_i=0,\;{(\xi_1-\xi_2)\cdot A(\xi_1-\xi_4)=0} \Big\}. \end{align}\] The first condition indicates that \(\xi_1,\xi_2,\xi_3,\xi_4\) form a parallelogram, while the second condition indicates some relations between the directions of the sides. We denote \[\mathrm{Cone}= \{ \xi \in \mathbb{Z}^3 \mid \xi \cdot A\xi =0 \},\] which will play a role in our arguments. We denote \(\mathcal{H}(S)\) the set of all planes (not necessarily passing through the origin) with normal vector belonging to \(S\subset \mathbb{Z}^3\). The set \(\mathcal{Q}\) can be decomposed as \(\mathcal{Q}_1 \cup \mathcal{Q}_2\), where \[\label{notation:Q1} \mathcal{Q}_1 = \{ (\xi_1,\xi_2,\xi_3,\xi_4) \in \mathcal{Q} \mid \xi_1 - \xi_2 \notin \mathrm{Cone}, \text{ and } \xi_1 - \xi_4 \notin \mathrm{Cone}\},\tag{7}\] \[\label{notation:Q2} \mathcal{Q}_2 = \{ (\xi_1,\xi_2,\xi_3,\xi_4) \in \mathcal{Q} \mid \xi_1 - \xi_2 \in \mathrm{Cone},\;\text{ or }\;\xi_1 - \xi_4 \in \mathrm{Cone}\}.\tag{8}\] We also denote the four-linear operators \[\Omega_1(f_1,f_2,f_3,f_4) = \sum_{{(\xi_1,\xi_2,\xi_3,\xi_4)} \in \mathcal{Q}_1 } f_1(\xi_1)f_2(\xi_2)f_3(\xi_3)f_4(\xi_4),\] \[\Omega_2 {(f_1,f_2,f_3,f_4)} = \sum_{{(\xi_1,\xi_2,\xi_3,\xi_4)} \in \mathcal{Q}_2} f_1(\xi_1)f_2(\xi_2)f_3(\xi_3)f_4(\xi_4).\] For simplicity, we write \(\Omega_1(f)\) and \(\Omega_2(f)\) instead of \(\Omega_1(f,f,f,f)\) and \(\Omega_2(f,f,f,f)\).

We introduce more notations. For any \(M>0\), \[\mathrm{Cone}_M = \{ \xi \in \mathrm{Cone}\setminus\{0\} \mid |\xi| /\gcd(\xi) \leq M \},\] \[\mathrm{Cone}^{\mathrm{irr}}_M = \{ \xi\in\mathrm{Cone}_M \mid \gcd(\xi)=1 \},\] where \(\gcd(\xi)\) denotes the greatest common divisor of coordinates of \(\xi\in\mathbb{Z}^3\).

Lemma 10. We have the size estimate \(\#\mathrm{Cone}^{\mathrm{irr}}_M \lesssim M\).

Proof. Suppose \((x_1,x_2,x_3) \in \mathrm{Cone}_M^{\text{irr}}\), i.e. \(x_1^2=x_2^2+x_3^2\) and \(\gcd(x_1,x_2,x_3)=1\). It’s clear that \(x_2,x_3\) cannot be both even, we may assume \(x_3\) is odd, and \(x_3 = \pm p_1^{\alpha_1}\dots p_r^{\alpha_r}\) is the prime factorization. We note that \[p_i^{2\alpha_i} | x_3^2 = (x_1-x_2)(x_1+x_2),\] so there exists some \(\gamma_i\in\mathbb{N}\) such that \(p^{\gamma_i}_i | (x_1-x_2)\) and \(p_i^{2\alpha_i-\gamma_i} | (x_1+x_2)\). If \(\gamma_i \neq 0,2\alpha_i\), then \(p_i\) divides both \(x_1-x_2\) and \(x_1+x_2\) and hence \(p_i\) divides both \(2x_1\) and \(2x_2\). Consequently \(p_i|\gcd(x_1,x_2,x_3)\), which is a contradiction. Hence we have exactly one of \(p_i^{2\alpha_i}|(x_1-x_2)\) and \(p_i^{2\alpha_i}|(x_1+x_2)\) holds. Denote \(I = \{ 1\leq i\leq r\mid p_i^{2\alpha_i} \text{ divides } x_1-x_2 \}\) and \[m = \prod_{i\in I} p_i^{\alpha_i},\quad n = \prod_{i\notin I} p_i^{\alpha_i} = |x_3|/m,\] where the product is defined to be \(1\) if the index set is empty. Then \(\gcd(m,n)=1\) and we have \((x_1-x_2 ,x_1+x_2 )= \pm (m^2,n^2)\), or equivalently \[(x_1, x_2) = \pm \left( \frac{n^2+m^2}{2}, \frac{n^2-m^2}{2} \right).\] Therefore, each point \((x_1,x_2,x_3) \in \mathrm{Cone}_M^{\text{irr}}\) can be represented by a pair \((m,n) \in \mathbb{Z}^2\) satisfying \(m^2+n^2 \lesssim M\), and hence \(\#\mathrm{Cone}_M^{\text{irr}} \lesssim M\). ◻

Before the start of proofs, we briefly talk about the geometry of parallelograms in \(\mathcal{Q}_2\). Due to symmetry, we may only consider the case \(\xi_1-\xi_2 \in \mathrm{Cone}\). For each \({(\xi_1,\xi_2,\xi_3,\xi_4)} \in \mathcal{Q}_2\) such that the parallelogram is non-degenerate, the four vertices are contained in some plane \(H\). From the definitions we can see that \(A(\xi_1-\xi_2)\) is perpendicular to both \(\xi_1-\xi_2\) and \(\xi_1-\xi_4\). Hence \(A(\xi_1-\xi_2)\) is a normal vector of \(H\), and it belongs to \(\mathrm{Cone}\). When the the parallelogram is degenerate, we can still find a plane \(H\) containing all vertices and its normal vector belongs to \(\mathrm{Cone}\).

On the other hand, let \(H\) be a plane with normal vector \(n\) which belongs to \(\mathrm{Cone}\), and suppose \(H\) contains the four vertices of \({(\xi_1,\xi_2,\xi_3,\xi_4)} \in \mathcal{Q}\). Clearly, \(n\) is perpendicular to \(\xi_1-\xi_2\), \(\xi_1-\xi_4\) and also \(An\). Notice that \[0 = (\xi_1-\xi_2) \cdot n = A(\xi_1-\xi_2) \cdot An,\] and \({(\xi_1,\xi_2,\xi_3,\xi_4)} \in \mathcal{Q}\) indicates that \(A(\xi_1-\xi_2) \cdot (\xi_1-\xi_4) = 0\). Hence we know that \[\operatorname{span}_{\mathbb{R}}\{ n,A(\xi_1-\xi_2) \} \perp \operatorname{span}_{\mathbb{R}} \{ \xi_1-\xi_4,An \}.\] But the sum of their dimensions is no more than \(3\). Therefore, we have that either \(\xi_1-\xi_4\) is a multiple of \(An\) or \(A(\xi_1-\xi_2)\) is a multiple of \(n\), and in both cases \({(\xi_1,\xi_2,\xi_3,\xi_4)}\) must belong to \(\mathcal{Q}_2\). As a result, we may write \[\label{decomposition32of32Q95232as32union32of32H94432cap32Q} \mathcal{Q}_2 = \bigcup_{H \in \mathcal{H}(\mathrm{Cone})} {\{ {(\xi_1,\xi_2,\xi_3,\xi_4)} \in \mathbb{Z}^{3\times 4} \mid \xi_i\in H,\;1\leq i\leq 4 \} } \cap \mathcal{Q}.\tag{9}\]

2.4 The contributions of parallelograms with side on the cone↩︎

Proposition 11. For \(f\colon \mathbb{Z}^3 \to \mathbb{R}_+\) supported on a finite subset \(S\subset \mathbb{Z}^3\), we have \[\Omega_2(f)\lesssim \operatorname{diam}(S) \| f \|_{\ell^2(\mathbb{Z}^3)}^4.\]

Proof. It suffices to consider the case \(\xi_1-\xi_2 \in \mathrm{Cone}\) and \(\xi_1\neq \xi_4\). For given \((\xi_1,\xi_4)\),

from previous discussion we know there exists some plane \(H\) contains both \(\xi_1,\xi_4\) and its normal vector belongs to \(\mathrm{Cone}\). It’s not hard to check that there exist at most two such planes. For each such plane \(H\), \(A(\xi_1-\xi_2)\) is a multiple of its normal vector and hence \(\xi_2\) lies on a line \(\ell\) passing through \(\xi_1\) with direction determined by \(H\).

We may write \(\xi_2 \in \ell \cap S \subset \mathbb{Z}^3\) as \(\xi_1+r\xi\) with \(\xi \in \mathbb{Z}^3\setminus\{0\}\), thus \(r\xi \in \mathbb{Z}^3\) and \(r\) belongs to an interval of length \(|\xi|^{-1} \operatorname{diam}(S)\). From Bézout’s identity, we know \(\gcd(\xi)\) can be written as linear combination of coordinates of \(\xi\) with integer coefficients, we have \(r \gcd(\xi) \in \mathbb{Z}\). Thus \[\label{number32of32points32on32intersection32of32line32and32S} \#(\ell \cap S) = \# \Bigl\{ r\in \frac{1}{\gcd(\xi)} \mathbb{Z} \Bigm| \xi_1+r\xi \in \ell\cap S \Bigr\} \leq \frac{\gcd(\xi)}{|\xi|} \operatorname{diam}(S),\tag{10}\] which implies for each pair \((\xi_1,\xi_4)\), there exists at most \(O(\operatorname{diam}(S))\) many choices of \((\xi_2,\xi_3)\) such that \({(\xi_1,\xi_2,\xi_3,\xi_4)} \in \mathcal{Q}_2\). We denote all the possible choices as \({(\xi_2,\xi_3)\in \mathcal{R}(\xi_1,\xi_4)}\).

On the other hand, for given \((\xi_2,\xi_3)\), we can also apply the same argument to \((\xi_1,\xi_4)\), and hence for each pair \((\xi_2,\xi_3)\), there exists at most \(O(\operatorname{diam}(S))\) many choices of \((\xi_1,\xi_4)\) such that \({(\xi_1,\xi_2,\xi_3,\xi_4)} \in \mathcal{Q}_2\). As a result, \[\begin{align} \Omega_2(f) &= \sum_{\xi_1,\xi_4} \Big( f(\xi_1)f(\xi_4) \sum_{(\xi_2,\xi_3)\in \mathcal{R}(\xi_1,\xi_4)} f(\xi_2)f(\xi_3) \Big) \\& \leq \Big( \sum_{\xi_1,\xi_4} |f(\xi_1)f(\xi_4)|^2 \Big)^{1/2} \Big( \sum_{\xi_1,\xi_4} \Big( \sum_{(\xi_2,\xi_3)\in \mathcal{R}(\xi_1,\xi_4)} f(\xi_2)f(\xi_3) \Big)^2 \Big)^{1/2} \\& \lesssim \operatorname{diam}(S)^{1/2} \|f\|^2_{\ell^2(\mathbb{Z}^3)} \Big( \sum_{\xi_1,\xi_4} \sum_{(\xi_2,\xi_3)\in \mathcal{R}(\xi_1,\xi_4)} | f(\xi_2)f(\xi_3) |^2 \Big)^{1/2} \\& =\operatorname{diam}(S)^{1/2} \|f\|^2_{\ell^2(\mathbb{Z}^3)} \Big( \sum_{\xi_2,\xi_3} \sum_{(\xi_1,\xi_4)\in\mathcal{R}(\xi_2,\xi_3)} | f(\xi_2)f(\xi_3) |^2 \Big)^{1/2} \\& \lesssim \operatorname{diam}(S) \|f\|_{\ell^2(\mathbb{Z}^3)}^4. \end{align}\] ◻

Proposition 12.

For \(f\colon \mathbb{Z}^3 \to \mathbb{R}_+\) supported on a finite subset \(S\subset \mathbb{Z}^3\) and \(M>0\), there exists at most \(O(M^3)\) planes \(\{H_i\} \subset \mathcal{H}(\mathrm{Cone}^{\mathrm{irr}}_M)\), such that \[\| f\chi_H \|_{\ell^2(\mathbb{Z}^3)}^2 \geq M^{-2} \|f\|_{\ell^2(\mathbb{Z}^3)}^2.\] If we denote \(f^{\mathrm{error}}:=f\chi_{S \setminus \cup_i H_i}\), then we have \[\label{estimate32for32f94error} \Omega_2(f^{\mathrm{error}}) \lesssim M^{-1}\operatorname{diam}(S) \| f\|_{\ell^2(\mathbb{Z}^3)}^4.\qquad{(3)}\]

Proof. We set \(\{H_i\}\) to be the set of all planes \(H\) with normal vector in \(\mathrm{Cone}_M^{\mathrm{irr}}\) and satisfying \(\| f\chi_H \|_{\ell^2(\mathbb{Z}^3)}^2 \geq M^{-2} \|f\|_{\ell^2(\mathbb{Z}^3)}^2\). For each \(n\in \mathrm{Cone}_M^{\mathrm{irr}}\), the planes with normal vector \(n\) are parallel with each other, which implies \[\# \{ H \in \mathcal{H}(\{n\}) \mid \| f\chi_H \|_{\ell^2(\mathbb{Z}^3)}^2 \geq M^{-2} \|f\|_{\ell^2(\mathbb{Z}^3)}^2 \} \leq M^2,\] thus \(\#\{H_i\} \leq M^2 \# \mathrm{Cone}_M^{\mathrm{irr}} \lesssim M^3\).

It remains to verify ?? . By the decomposition 9 and the facts that \(0\leq f^{\mathrm{error}} \leq f\), \(f^{\mathrm{error}} \chi_{H_i}=0\) and \(\mathcal{H}(\mathrm{Cone}_M^{\mathrm{irr}})= \mathcal{H}(\mathrm{Cone}_M)\), we get \[\begin{align} \Omega_2(f^{\mathrm{error}}) &\leq \sum_{ H \in \mathcal{H}(\mathrm{Cone}) } \Omega_2(f^{\mathrm{error}} \chi_H) \\& \leq \sum_{ H \in \mathcal{H}(\mathrm{Cone}\setminus \mathrm{Cone}_M) } \Omega_2(f \chi_H) + \sum_{ H \in \mathcal{H}(\mathrm{Cone}_M^{\mathrm{irr}} ) \setminus \{H_i\} } \Omega_2(f \chi_H). \end{align}\] Recall the estimate 10 , by using arguments similar to that in proof of Proposition 11, we have \[\sum_{ H \in \mathcal{H}(\mathrm{Cone}\setminus \mathrm{Cone}_M) } \Omega_2(f \chi_H) \lesssim M^{-1}\operatorname{diam}(S) \| f\|_{\ell^2(\mathbb{Z}^3)}^4.\] On the other hand, for each \(H \in \mathcal{H}(\mathrm{Cone}_M^{\mathrm{irr}}) \setminus \{H_i\},\) we have \(\| f\chi_H \|_{\ell^2(\mathbb{Z}^3)}^2 \leq M^{-2} \| f \|_{\ell^2(\mathbb{Z}^3)}^2\), and \[\begin{align} \sum_{ H \in \mathcal{H}(\mathrm{Cone}_M^{\mathrm{irr}} ) \setminus \{H_i\} } \Omega_2(f \chi_H) & ={} \sum_{n\in \mathrm{Cone}_M^{\mathrm{irr}}} \sum_{\substack{H \notin \{H_i\} \\ H \perp n}} \Omega_2(f\chi_H) \\& \lesssim{} \sum_{n\in \mathrm{Cone}_M^{\mathrm{irr}}} \sum_{\substack{H \notin \{H_i\} \\ H \perp n}}\operatorname{diam}(S) \| f\chi_H \|_{\ell^2(\mathbb{Z}^3)}^4 \\& \leq{} \frac{\operatorname{diam}(S) \|f \|_{\ell^2(\mathbb{Z}^3)}^2 }{M^2}\sum_{n\in \mathrm{Cone}_M^{\mathrm{irr}}} \sum_{\substack{H \notin \{H_i\} \\ H \perp n}} \|f\chi_H\|_{\ell^2(\mathbb{Z}^3)}^2 \\& \leq{} M^{-2}\#\mathrm{Cone}_M^{\mathrm{irr}} \operatorname{diam}(S) \|f\|_{\ell^2(\mathbb{Z}^3)}^4 \\& \lesssim M^{-1} \operatorname{diam}(S) \| f ||_{\ell^2(\mathbb{Z}^3)}^4 . \end{align}\] Here we used Proposition 11 for the first inequality and Lemma 10 for the last inequality. Hence \[\Omega_2(f^{\mathrm{error}}) \lesssim M^{-1}\operatorname{diam}(S) \| f\|_{\ell^2(\mathbb{Z}^3)}^4.\] ◻

Combining the above two propositions, we see that if \(\Omega_2(f)\) is large, then \(f\) should concentrate on few planes. Thus we get more information about the geometric structure of the distribution of \(f\). This observation is crucial in our proof.

2.5 The contributions of parallelograms without side on the cone↩︎

Proposition 13. For \(f=\chi_S\) with \(S\) a finite subset of \(\mathbb{Z}^3\), we have \[\Omega_1(f) \lesssim (\#S)^{7/3}=(\#S)^{1/3} \| f\|^4_{\ell^2(\mathbb{Z}^3)}.\label{bound:Omega1}\qquad{(4)}\]

This can be proved by the same method in [23]. For \(\xi\in\mathbb{Z}^3\) and \(\ell,\ell'\) two lines in \(\mathbb{R}^3\), we denote \[\mathfrak{c}(\xi,\ell,\ell') = \begin{cases} 1,& \text{if } \xi = \ell \cap \ell' \text{ and }v_\ell \cdot Av_{\ell'}=0,\\0,&\text{otherwise,} \end{cases}\] where \(v_\ell\) denotes the direction vector of the line \(\ell\). We need the following lemma.

Lemma 14. For fixed \(\xi \in \mathbb{Z}^3\), let \(\mathcal{L},\mathcal{L'}\) be two finite families of lines passing through \(\xi\). Then \[\sum_{\ell \in \mathcal{L}} \sum_{\ell'\in\mathcal{L}'} \mathfrak{c}(\xi,\ell,\ell') \lesssim (\# \mathcal{L})^{2/3}(\#\mathcal{L}')^{2/3} + \#\mathcal{L}+\#\mathcal{L}'.\]

Proof. For each line \(\ell\), we denote its direction vector as \(v_\ell \in \mathbb{S}^2\), and \(c_\ell = \{ v\in \mathbb{S}^2 \mid v \cdot Av_{\ell} =0 \}\). Then \(\mathfrak{c}(\xi,\ell,\ell') = 1\) is equivalent to \(v_\ell \in c_{\ell'}\). Set \(\mathcal{P} = \{ v_\ell \mid \ell \in \mathcal{L} \}\) and \(\mathcal{C} = \{ c_{\ell'} \mid \ell'\in \mathcal{L}' \}\), then the summation of \(\mathfrak{c}(\xi,\ell,\ell')\) is bounded by the number of incidences between points in \(\mathcal{P}\) and circles in \(\mathcal{C}\), see Theorem 8. ◻

Figure 2: image.

Proof of Proposition 13. For each dyadic integer \(s\), we put \(\mathcal{L}_s\) to be the set of lines \(\ell\) such that \(s\leq \#(\ell\cap S) < 2s\) and the direction vector of \(\ell\) doesn’t belong to \(\mathrm{Cone}\). Then \[\begin{align} \Omega_1(f) &= \sum_{s,t\text{ dyadic}} \#\{ {(\xi_1,\xi_2,\xi_3,\xi_4)}\in \mathcal{Q}_1 \mid \ell_{\xi_1\xi_2} \in \mathcal{L}_s,\ell_{\xi_1,\xi_4}\in\mathcal{L}_t \} \\& \leq \sum_{s,t\text{ dyadic}} \sum_{\ell \in \mathcal{L}_s} \sum_{\ell'\in \mathcal{L}_t} \sum_{\xi_1\in S} \mathfrak{c}(\xi_1,\ell,\ell')\#(\ell\cap S)\#(\ell'\cap S). \label{bound:Omega1-inci} \end{align}\tag{11}\] For \(s,t\leq (\#S)^{1/3}\), we use Lemma 14 to estimate \[\begin{align} \sum_{s,t\text{ dyadic}} \sum_{\xi_1\in S}st \Big( \sum_{\ell \in \mathcal{L}_s} \sum_{\ell'\in \mathcal{L}_t} \mathfrak{c}(\xi_1,\ell,\ell') \Big) \lesssim \sum_{s,t\text{ dyadic}} st \sum_{\xi_1\in S}( (\mathcal{J}_{\xi_1}^s\mathcal{J}_{\xi_1}^t)^{2/3} + \mathcal{J}_{\xi_1}^s+\mathcal{J}_{\xi_1}^t ), \end{align}\] where \(\mathcal{J}_{\xi_1}^s,\mathcal{J}_{\xi_1}^t\) denote the number of lines in \(\mathcal{L}_s,\mathcal{L}_t\) passing through \(\xi_1\). By Corollary 7, we have \[\sum_{\xi_1\in S} \mathcal{J}_{\xi_1}^s \lesssim \frac{(\#S)^2}{s^2},\quad \sum_{\xi_1\in S} \mathcal{J}_{\xi_1}^t \lesssim \frac{(\#S)^2}{t^2},\] hence \[\sum_{s,t \leq (\#S)^{1/3}\text{ dyadic}} st \sum_{\xi_1\in S}(\mathcal{J}_{\xi_1}^s+\mathcal{J}_{\xi_1}^t) \lesssim (\#S)^{7/3}.\] On the other hand, since all the lines passing through \(\xi_1\) are pairwise disjoint (excluding the common point \(\xi_1\)), we have \(\mathcal{J}_{\xi_1}^s \lesssim \#S/s\) and \(\mathcal{J}_{\xi_1}^t \lesssim \#S/t\). Therefore, \[\begin{align} \sum_{s,t\text{ dyadic}} st \sum_{\xi_1\in S} (\mathcal{J}_{\xi_1}^s\mathcal{J}_{\xi_1}^t)^{2/3} &\lesssim \sum_{s,t\text{ dyadic}} st \sum_{\xi_1 \in S} \left( \frac{(\#S)^2}{st} \right)^{1/6} (\mathcal{J}_{\xi_1}^s\mathcal{J}_{\xi_1}^t)^{1/2} \\&\leq \sum_{s,t\text{ dyadic}} {(\#S)^{1/3}}{(st)^{5/6}} \Big( \sum_{\xi_1\in S} \mathcal{J}_{\xi_1}^s \Big)^{1/2}\Big( \sum_{\xi_1\in S} \mathcal{J}_{\xi_1}^t\Big)^{1/2} \\& \lesssim \sum_{s,t\text{ dyadic}} {(\#S)^{1/3}}{(st)^{5/6}}\frac{\#S}{s}\frac{\#S}{t} \lesssim (\#S)^{7/3}. \end{align}\] This proves ?? when \(s,t\leq (\#S)^{1/3}\).

For \(s\geq(\#S)^{1/3}\) or \(t\geq (\#S)^{1/3}\), we assume \(s\geq (\#S)^{1/3}\) without loss of generality. We notice that for any fixed \(\ell\) and \(\xi_1\in\ell\), the lines \(\ell'\) such that \(\mathfrak{c}(\xi_1,\ell,\ell')\neq 0\) belong to some plane determined by \(\ell\) and \(\xi_1\). Furthermore, for different choices of \(\xi_1\), these planes are pairwise disjoint due to the fact that the direction vector of \(\ell\) doesn’t belong to \(\mathrm{Cone}\). Thus all these lines \(\ell'\) are pairwise disjoint (excluding the possible common points on \(\ell\)), and \[\sum_t \sum_{\ell'\in\mathcal{L}_t} \sum_{\xi_1\in S} \mathfrak{c}(\xi_1,\ell,\ell') \#(\ell'\cap S) \lesssim \#S. \label{est:case2-1}\tag{12}\] On the other hand, by Corollary 7 we have \[\sum_{s\geq (\#S)^{1/3}} \sum_{\ell \in \mathcal{L}_s} \#(\ell \cap S) \lesssim (\#S)^{4/3}.\label{est:case2-2}\tag{13}\] 11 , 12 and 13 together imply the bound ?? . ◻

2.6 Proof of Theorem 1↩︎

Proposition 13 deals with characteristic functions. To extend this result to the general case, we employ an atomic decomposition that reduces an arbitrary function to a sum of characteristic functions. This leads us to the task of estimating the multilinear form \(\Omega_1(f_1, f_2, f_3, f_4)\) with \(\operatorname{supp} f_i\subset S_i\). A similar problem was tackled by Herr-Kwak [4], who performed a very careful analysis to bound the number of parallelograms in terms of the size of the sets \(S_i\). However, the problem becomes more intricate in 3 dimensions. To overcome this, we apply Proposition 12 to further reduce the problem to the case where the support of each function \(f_i\) lies in a plane \(H_i\in \mathcal{H}(\mathrm{Cone}^{\mathrm{irr}}_M)\). The particular geometric structure of these planes becomes crucial for obtaining the desired estimate. Now we turn to the details.

Proof of Theorem 1. We set \(f = |\hat{\phi}| \chi_{[-N,N]^3}\) and enumerate \(\mathbb{Z}^3 \cap [-N,N]^3\) as \(\xi_1,\xi_2,\dots\) such that \[f(\xi_1) \geq f(\xi_2) \geq \dots.\] Let \(S_j = \{ \xi_{2^{j}},\dots,\xi_{2^{j+1}-1} \}\), \(f_j = f(\xi_{2^j}) \chi_{S_j}\) and \(\lambda_j = 2^{j/2} f(\xi_{2^j})\) for \(0\leq j \leq {j_{\text{max}}}\) with \(2^{j_{\text{max}}}\lesssim N^3\). We have \(\# S_j \leq 2^j\), \(f \leq \sum_{j} f_j\), \(|\lambda_j|=\|f_j\|_{\ell^2(\mathbb{Z}^3)}\) and \[\| \lambda_j\|_{\ell^2_{j\leq {j_{\text{max}}}}}^2 = \sum_{j=0}^{{j_{\text{max}}}} 2^j|f(\xi_{2^j})|^2 \leq |f(\xi_1)|^2 + \sum_j \| f\chi_{S_{j-1} }\|_{\ell^2(\mathbb{Z}^3)}^2 \lesssim \|f\|_{\ell^2(\mathbb{Z}^3)}^2. \label{est:lambda-f}\tag{14}\] Let \(\delta>0\) sufficiently small. Given such a decomposition of \(f\), we say \(j\) is good if \(\Omega_2(f_j) \lesssim N^{1-\delta} \| f_j \|_{\ell^2(\mathbb{Z}^3)}^4\), otherwise we say \(j\) is bad. For each bad \(j\), by taking \(M=N^\delta\) in Proposition 12, we can find at most \(O(N^{3\delta})\) planes \(\{H_{j}^{i_j}\} \subset \mathcal{H}(\mathrm{Cone}_M^{\mathrm{irr}})\), such that \(f_j^{\mathrm{error}}:=f\chi_{S_j \setminus \cup_{i_j} H_{j}^{i_j}}\) satisfies \[\Omega_2(f_j^{\mathrm{error}}) \lesssim N^{1-\delta} \| f_j\|_{\ell^2(\mathbb{Z}^3)}^4. \label{def:fgood}\tag{15}\] We denote \(f_j^{\mathrm{good}}=f_j^{}\) if \(j\) is good and \(f_j^{\mathrm{good}}=f_j^{\mathrm{error}}\) if \(j\) is bad, and \[f_{\text{bad}} = \sum_{j \text{ bad} } (f_j^{}-f_j^{\mathrm{good}}).\] Then from Proposition 13 and 15 , we get bounds for each \(f_j^{\mathrm{good}}\) \[\begin{align} \| \mathrm{e}^{\mathrm{i}t\Box} \check{f}_j^\text{good} \|^4_{{L_{t,x}^4([0,1]\times \mathbb{T}^3)}} &= \Omega_1(f_j^\text{good}) + \Omega_2(f_j^\text{good})\\ & \lesssim (2^{{j}/{3}}+N^{1-\delta})\|f_j^\text{good}\|^4_{\ell^2(\mathbb{Z}^3)}. \label{est:f-good} \end{align}\tag{16}\] Now we control the contribution of \(f_{\text{bad}}\), which can be written as \[\begin{align} &\| \mathrm{e}^{\mathrm{i}t\Box} \check f_{\text{bad}}\|_{{L_{t,x}^4([0,1]\times \mathbb{T}^3)}}^4 = \Omega_1(f_{\text{bad}}) +\Omega_2(f_{\text{bad}}) . \end{align}\] Proposition 11 indicates that \(\Omega_2(f_{\text{bad}}) \lesssim N \| f_{\text{bad}} \|_{{\ell^2(\mathbb{Z}^3)}}^4\). Meanwhile, we have the estimate \[\begin{align} \Omega_1(f_{\text{bad}}) &\lesssim \sum_{\substack{j_1,j_2,j_3,j_4 \text{ bad} \\ i_{j_1},i_{j_2},i_{j_3},i_{j_4} \lesssim N^{3\delta}}} \Omega_1(f_{j_1}\chi_{H_{j_1}^{i_{j_1}}},f_{j_2}\chi_{H_{j_2}^{i_{j_2}}}, f_{j_3}\chi_{H_{j_3}^{i_{j_3}}}, f_{j_4}\chi_{H_{j_4}^{i_{j_4}}}) \\ &\lesssim \sum_{j_1,j_2,j_3,j_4} N^{3/4+O(\delta)} \prod_{k=1}^4 \| f_{j_k} \|_{\ell^2(\mathbb{Z}^3)} \lesssim N \| f_{\text{bad}}\|_{\ell^2(\mathbb{Z}^3)}^4, \end{align}\] which is a consequence of applying the following Proposition 15 to each quadruple \((i_{j_1},i_{j_2},i_{j_3},i_{j_4})\) with \(M=N^\delta\).

Combining all the estimates, we get \[\begin{align} \| \mathrm{e}^{\mathrm{i}t\Box}\check{f} \|_{{L_{t,x}^4([0,1]\times \mathbb{T}^3)}} &\lesssim \sum_{j} \| \mathrm{e}^{\mathrm{i}t\Box} \check{f}_j^\text{good} \|_{{L_{t,x}^4([0,1]\times \mathbb{T}^3)}} + \| \mathrm{e}^{\mathrm{i}t\Box} \check f_{\text{bad}}\|_{{L_{t,x}^4([0,1]\times \mathbb{T}^3)}}\\& \lesssim \sum_{j \leq {j_{\text{max}}}} (2^{j/12} + N^{(1-\delta)/4})\lambda_j + \| \mathrm{e}^{\mathrm{i}t\Box} \check f_{\text{bad}}\|_{{L_{t,x}^4([0,1]\times \mathbb{T}^3)}}\\& \lesssim N^{1/4} \| \lambda_j \|_{\ell_{j\leq {j_{\text{max}}}}^2} + \| \mathrm{e}^{\mathrm{i}t\Box} \check f_{\text{bad}}\|_{{L_{t,x}^4([0,1]\times \mathbb{T}^3)}}\\ &\lesssim N^{1/4} \| f\|_{\ell^2(\mathbb{Z}^3)}. \end{align}\] From 6 , we see \(\| \mathrm{e}^{\mathrm{i}t\Box}\phi \|_{{L_{t,x}^4([0,1]\times \mathbb{T}^3)}}\leq \| \mathrm{e}^{\mathrm{i}t\Box}\check{f} \|_{{L_{t,x}^4([0,1]\times \mathbb{T}^3)}}\). Thus we finish the proof for the \(L^4\) estimate ?? , and the general case ?? follows from interpolation of \(L^4\) with \(L^2\) and \(L^\infty\) estimates. ◻

Now we are left to prove the following:

Proposition 15. Suppose \(n_1,n_2,n_3,n_4 \in \mathrm{Cone}^{\mathrm{irr}}_M\) are vectors and \(H_1,H_2,H_3,H_4\) are planes such that \(n_j\perp H_j\) for each \(j\), and functions \(f_j \colon \mathbb{Z}^3 \to \mathbb{R}_+\) supported on \({ H_j \cap [-N,N]^3}\). Then \[\Omega_1{(f_1,f_2,f_3,f_4)} \lesssim M^{2}N^{1/2}(M^2+N^{1/4}) \prod_{j=1}^4 \| f_j\|_{\ell^2(\mathbb{Z}^3)}. \label{est:prop-4f}\qquad{(5)}\]

Proof. We decompose \(\mathcal{Q}_1\) into \[\bigcup_{(a,b) \in \mathbb{Z}^3\times \mathbb{Z} } \Gamma_{a,b} :=\bigcup \Bigg\{ {(\xi_1,\xi_2,\xi_3,\xi_4)} \in \mathcal{Q}_1 \Biggm| \begin{array}{cc} \xi_1+\xi_3=a=\xi_2+\xi_4 \\ h(\xi_1)+h(\xi_3) = b = h(\xi_2)+h(\xi_4)\end{array}\Bigg\}.\] Also we denote \(\mathcal{A}_{a,b} = \{ \xi \in \mathbb{Z}^3 \mid h(\xi)+h(a-\xi) = b \}\). Then \({(\xi_1,\xi_2,\xi_3,\xi_4)} \in \Gamma_{a,b}\) implies \(\xi_j \in \mathcal{A}_{a,b}\) for \(1\leq j\leq 4\). It’s easy to see that \[\begin{align} \Omega_1{(f_1,f_2,f_3,f_4)} &= \sum_{a,b} \sum_{\Gamma_{a,b}} f_1(\xi_1)f_2(\xi_2)f_3(\xi_3)f_4(\xi_4) \tag{17} \\& = \sum_{a,b} \Big( \sum_{\substack{\xi_1,\xi_3\in \mathcal{A}_{a,b}\\\xi_1+\xi_3=a}} f_1(\xi_1)f_3(\xi_3) \Big)\Big( \sum_{\substack{\xi_2,\xi_4\in \mathcal{A}_{a,b}\\\xi_2+\xi_4=a}} f_2(\xi_2)f_4(\xi_4) \Big) . \\\tag{18} \end{align}\] By Cauchy-Schwarz, it’s further bounded by \[\bigg( \sum_{a,b} \Big( \sum_{\substack{\xi_1 \in \mathcal{A}_{a,b} }} f_1(\xi_1)f_3(a-\xi_1) \Big)^2 \bigg)^{1/2} \bigg( \sum_{a,b} \Big( \sum_{\substack{\xi_2 \in \mathcal{A}_{a,b} }} f_2(\xi_2)f_4(a-\xi_2) \Big)^2 \bigg)^{1/2}.\label{est:4f-Aab-CS}\tag{19}\]

Next we estimate the size of \(\mathcal{A}_{a,b}\cap H_1\cap[-N,N]^3\), the same argument holds for \(j=2,3,4\). Since \(n_1\in \mathrm{Cone}_M^{\mathrm{irr}}\) is the normal vector of \(H_1\) and \(n_1 \perp An_1\), hence \(\{n_1,An_1,n_1\times An_1\}\) is an orthogonal basis of \(\mathbb{R}^3\). We decompose \(\xi\) with respect to this orthogonal basis and denote \[\xi^\perp = \xi - \frac{ \xi\cdot n_1 }{|n_1|^2} n_1 - \frac{\xi\cdot An_1}{|n_1|^2}An_1 \in \frac{1}{|n_1|^2} \mathbb{Z}^3. \label{definition32of32perp32part}\tag{20}\] Clearly \(\xi^\perp\) belongs to the one-dimensional linear subspace spanned by \(n_1\times An_1\). For \(\xi\in H_1\), the inner product \(\xi\cdot n_1 = c_1\) is constant, and we compute \(h(\xi)\) by using the decomposition 20 \[\begin{align} h(\xi) &= \xi \cdot A\xi = \xi^\perp \cdot A\xi^\perp + (\xi - \xi^\perp)\cdot A(\xi-\xi^\perp) + 2\xi^\perp \cdot A(\xi-\xi^\perp) \\ &= h(\xi^\perp) + 2c_1 \frac{\xi\cdot An_1}{|n_1|^2} . \end{align}\] As a result, it holds that for \(\xi \in \mathcal{A}_{a,b} \cap H_1\) \[\begin{align} b &= h(\xi) + h(a-\xi) = 2h(\xi)+h(a)-2\xi\cdot Aa \\& = 2h(\xi^\perp) + 4c_1 \frac{ \xi \cdot An_1 }{|n_1|^2} + h(a) -2\xi^\perp\cdot Aa - \frac{2( c_1 a\cdot An_1 + (\xi_1\cdot An_1)(a\cdot n_1) )}{|n_1|^2} \end{align}\] i.e. \[\frac{4c_1 - 2a\cdot n_1}{|n_1|^2} \xi\cdot An_1 = -2h(\xi^\perp)+2\xi^\perp\cdot Aa + \Big[ b - h(a) + \frac{2c_1}{|n_1|^2} a\cdot An_1 \Big]. \label{equation32of32xi44a44b}\tag{21}\]

Figure 3: image.

The following lines are dedicated to further simplify 21 , which gives us the desired estimate of the size of \(\mathcal{A}_{a,b}\cap H_1\cap[-N,N]^3\). As in 20 , we denote \[a^\perp = a - \frac{ a \cdot n_1 }{|n_1|^2} n_1 - \frac{a \cdot An_1}{|n_1|^2}An_1 \in \frac{1}{|n_1|^2} \mathbb{Z}^3. \label{a32perp32part}\tag{22}\] There exists some \(\eta \in \mathbb{Z}^3\) such that \(|\eta| \leq |n_1\times An_1| \leq M^2\) and \(2\xi^\perp - a^\perp = \frac{\lambda}{|n_1|^2} \eta\) for some \(\lambda \in\mathbb{Z}\). Since the matrix \(A\) is non-singular, \(h(\eta)\) must be non-zero. Now 21 can be written as \[\begin{align} \frac{\lambda^2}{|n_1|^4} h(\eta) &= h(2\xi^\perp-a^\perp) \\ &= 2\left[ b+h(a^\perp)-h(a)+\frac{2c_1}{|n_1|^2}a\cdot n_1 \right] - \frac{8c_1-4a\cdot n_1}{|n_1|^2}\xi\cdot An_1. \end{align}\] Set \(z = \lambda h(\eta)\), \(y = |n_1|^2 \xi \cdot An_1\) and \[\begin{align} q&= h(\eta )( 8c_1-4a_1\cdot n_1) , \\ \omega&=2|n_1|^4h(\eta )\left[ b+h(a^\perp)-h(a)+\frac{2c_1}{|n_1|^2}a\cdot n_1 \right], \end{align}\] then \((y,z,q,\omega)\) is an integer solution of \[z^2=qy+\omega, \quad |q|,|y|,|z| \lesssim M^8N. \label{eq:claim}\tag{23}\]

For given \(a,b\), if \(a \cdot n_1 \neq 2c_1\), which is equivalent to \(a/2 \notin H_1\), then from 21 , \(\xi\cdot An_1\) is determined by \(\xi^\perp\). From 20 , \(\xi\in \mathcal{A}_{a,b} \cap H_1\) is determined by \(\xi^\perp\) since \(\xi\cdot n_1=c_1\). Let us state the following claim, postponing its proof for later.

Claim: For given \(N,q\in\mathbb{Z}_+\) and \(\omega \in \mathbb{Z}\), \[\#\Big( \{ (y,z)\in \mathbb{Z}^2 \mid z^2= qy+\omega \} \cap [-N,N]^2\Big)\lesssim \sqrt{N}+\sqrt{q}.\] Applying the claim to 23 , we see that there are \(O(M^{4}N^{1/2})\) many choices of \(\xi^\perp\), hence \[\#(\mathcal{A}_{a,b}\cap H_1\cap[-N,N]^3) \lesssim M^{4} N^{1/2}.\label{Aab-bound:curve}\tag{24}\]

On the other hand, if \(a\cdot n_1 = 2c_1\), i.e. \(a/2\in H_1\), the left hand side of 21 is \(0\) and we have at most two choices of \(\xi^\perp\), hence from 20 we know \(\xi\) belongs to the union of two lines contained in \(H_1\), \[\#(\mathcal{A}_{a,b} \cap H_1\cap[-N,N]^3) \lesssim N. \label{Aab-bound:line}\tag{25}\]

Now we prove the main estimate ?? .

Case 1: \(a\cdot n_{j_0} \neq 2c_{j_0}\), i.e. \(a/2\notin H_{j_0}\), for some \(1\leq j_0\leq 4\). Without loss of generality, we assume \(j_0=1\). By Cauchy-Schwarz and 24 we have \[\begin{align} \Big( \sum_{\xi \in \mathcal{A}_{a,b} } f_{1}(\xi) f_{3}(a-\xi) \Big)^2 \lesssim M^{4} N^{1/2} \sum_{\xi\in\mathcal{A}_{a,b}} |f_{1}(\xi)|^2 | f_{3}(a-\xi)|^2, \end{align}\] summing over \(a,b\) gives that \[\sum_{\frac{a}{2} \notin H_1} \sum_b \Big( \sum_{\xi \in \mathcal{A}_{a,b} } f_{1}(\xi) f_{3}(a-\xi) \Big)^2 \lesssim M^{4}N^{1/2} \| f_1 \|_{\ell^2(\mathbb{Z}^3)}^2 \| f_3 \|_{\ell^2(\mathbb{Z}^3)}^2.\] On the other hand \[\begin{align} \sum_{a,b} \Big( \sum_{\xi \in \mathcal{A}_{a,b} } f_{2}(\xi) f_{4}(a-\xi) \Big)^2 &\lesssim \max\{ M^4N^{1/2}, N\} \sum_{a,b} \sum_{\xi\in\mathcal{A}_{a,b}} |f_{2}(\xi)|^2 | f_{4}(a-\xi)|^2 \\& \lesssim \max\{M^4N^{1/2},N\} \| f_2 \|_{\ell^2(\mathbb{Z}^3)}^2 \| f_4 \|_{\ell^2(\mathbb{Z}^3)}^2, \end{align}\] as a result \[\begin{align} \sum_{\frac{a}{2} \notin H_1} \sum_b \sum_{\Gamma_{a,b}} f_1(\xi_1)f_2(\xi_2)f_3(\xi_3)f_4(\xi_4) \lesssim M^{2}N^{1/2}(M^2+N^{1/4}) \prod_{j=1}^4 \| f_j \|_{\ell^2(\mathbb{Z}^3)}. \end{align}\]

Case 2: \(a/2 \in H_j\) for all \(j\).

•If \(\dim \operatorname{span}_\mathbb{R} \{n_1,n_2,n_3,n_4\} = 3\). In this case there are at most one such \(a\), and hence from 18 \[\begin{align} & \sum_{\frac{a}{2} \in \cap_j H_j} \sum_b \sum_{\Gamma_{a,b}} f_1(\xi_1)f_2(\xi_2)f_3(\xi_3)f_4(\xi_4)\\ \lesssim & \Big(\sum_{\xi_1} f_1(\xi_1)f_3(a-\xi_1)\Big)\Big( \sum_{\xi_2} f_2(\xi_2)f_4(a-\xi_2) \Big) \leq \prod_{j=1}^4 \| f_j \|_{\ell^2(\mathbb{Z}^3)}. \end{align}\]

•If \(\dim \operatorname{span}_\mathbb{R} \{n_1,n_2,n_3,n_4\} = 2\). Let \(2\leq j \leq 4\) be such that \[\operatorname{span}_\mathbb{R} \{ n_1,n_j\} = \operatorname{span}_{\mathbb{R}} \{ n_1,n_2,n_3,n_4 \},\] we will prove \[\sum_{\frac{a}{2}\in H_1\cap H_j} \sum_{b} \Big( \sum_{\substack{\xi \in \mathcal{A}_{a,b} }} f_1(\xi)f_3(a-\xi) \Big)^2 \lesssim \| f_1\|_{\ell^2(\mathbb{Z}^3)}^2 \| f_3 \|_{\ell^2(\mathbb{Z}^3)}^2.\] By symmetry, the same estimate holds for \(f_2,f_4\).

Recall the definition of \(a^\perp\) in 22 , then we can write that \[2c_j = a\cdot n_j = \left[ a^\perp \cdot n_j + \frac{2c_1}{|n_1|^2}{ n_1\cdot n_j} \right] + \frac{a\cdot An_1}{|n_1|^2} An_1\cdot n_j.\] We claim that \(An_1 \cdot n_j = n_1\cdot An_j\neq 0\). Otherwise, since \(n_1,n_j \in \mathrm{Cone}\), it holds that \[\operatorname{span}_{\mathbb{R}}\{An_1,An_j\} \perp \operatorname{span}_{\mathbb{R}}\{n_1,n_j\},\] but their dimensions are both \(2\) , which is a contradiction. As a result, we can solve \(a\cdot An_1\) in terms of \(a^\perp\), and hence \(a\) is determined by \(a^\perp\).

We set \(\mathcal{A}_{a,b}^\perp = \{ \xi^\perp \mid \xi \in \mathcal{A}_{a,b} \}\), where \(\xi^\perp\) is defined as 20 . Then \[\mathcal{A}_{a,b} =\bigcup_{\beta\in \mathcal{A}_{a,b}^\perp}\{\xi \in \mathcal{A}_{a,b} \mid \xi^\perp=\beta\} .\] From 21 we see that \(\# \mathcal{A}_{a,b}^\perp \leq 2\) for each \(a,b\).

By Cauchy-Schwarz inequality, \[\begin{align} \Big( \sum_{\substack{\xi \in \mathcal{A}_{a,b} }} f_1(\xi)f_3(a-\xi) \Big)^2 &\lesssim \sum_{{\beta} \in \mathcal{A}_{a,b}^\perp} \Big(\sum_{\substack{\xi\in [-N,N]^3, \xi^\perp = {\beta}}} f_1(\xi)f_3(a-\xi) \Big)^2 \\&\leq \sum_{{\beta} \in \mathcal{A}_{a,b}^\perp } {\Big( \sum_{\xi^\perp={\beta}} |f_1(\xi)|^2 \Big)} {\Big( \sum_{\xi^\perp = a^\perp - {\beta}} |f_3(\xi)|^2\Big)}. \end{align}\] Therefore, \[\begin{align} & \mathrel{\phantom{\lesssim}} \sum_{\frac{a}{2}\in H_1\cap H_j} \sum_{b} \Big( \sum_{\substack{\xi \in \mathcal{A}_{a,b} }} f_1(\xi)f_3(a-\xi) \Big)^2 \\ &\lesssim \sum_{\frac{a}{2}\in H_1\cap H_j} \sum_{b} \sum_{{\beta} \in \mathcal{A}_{a,b}^\perp } {\Big( \sum_{\xi^\perp={\beta}} |f_1(\xi)|^2 \Big)} {\Big( \sum_{\xi^\perp = a^\perp - {\beta}} |f_3(\xi)|^2\Big)}\\& = \sum_{\frac{a}{2}\in H_1\cap H_j} \sum_{{\beta}} {\Big( \sum_{\xi^\perp={\beta}} |f_1(\xi)|^2 \Big)} {\Big( \sum_{\xi^\perp = a^\perp - {\beta}} |f_3(\xi)|^2\Big)} \\& = \| f_1\|_{\ell^2(\mathbb{Z}^3)}^2 \| f_3 \|_{\ell^2(\mathbb{Z}^3)}^2. \end{align}\]

• If \(\dim\operatorname{span}_\mathbb{R}\{ n_1,n_2,n_3,n_4\}=1\). In this case \(n_1=n_2=n_3=n_4\) and \(H_1=H_2=H_3=H_4\), otherwise the intersection is empty. Suppose \({(\xi_1,\xi_2,\xi_3,\xi_4)} \in \mathcal{Q}\) with \(\xi_1-\xi_2\) is not a multiple of \(An_1\), we have \({(\xi_1-\xi_2)\cdot A(\xi_1-\xi_4)=0}\) and \(An_1\cdot A(\xi_1-\xi_4)=0\) since \(n_1\) is the normal vector of \(H_1\), which implies \(A(\xi_1-\xi_4)\) is also a normal vector of \(H_1\) and hence \(A(\xi_1-\xi_4)\) is a multiple of \(n_1\) and \(\xi_1-\xi_4\in \mathrm{Cone}\). As a result, we must have \({(\xi_1,\xi_2,\xi_3,\xi_4)}\in\mathcal{Q}_2\) and \(\Omega_1{(f_1,f_2,f_3,f_4)}=0\). ◻

Proof of the claim. Denote \(\wp_{q,\omega} = \{ (y,z)\in \mathbb{Z}^2 \mid z^2= qy+\omega \} \cap [-N,N]^2\) for fixed \(N,q\in\mathbb{Z}_+\) and \(\omega \in \mathbb{Z}\). Suppose \(q = p_1^{\alpha_1}\dots p_r^{\alpha_r}\) is the prime factorization, we denote \(\theta_i(z)\) the minimal residue of \(z\pmod {p_i^{\alpha_i}}\) for \(1\leq i\leq r\) and \(\theta_0(z) = \lfloor z/q\rfloor\). Then the map \[\mathbb{Z} \to \mathbb{Z}^{1+r},\quad z\mapsto (\theta_0(z),\theta_1(z),\dots,\theta_r(z))\] is an injection. In fact, if \(\theta_i(z) = \theta_i(\tilde{z} )\) for all \(1\leq i\leq r\), then \(z-\tilde{z}\) is divided by all \(p_i^{\alpha_i}\) and hence \(z -\tilde{z}\) is divided by \(q\). But \(\theta_0(z)=\theta_0(\tilde{z})\) implies \(0\leq |z-\tilde{z}|\leq q-1\), which forces that \(z=\tilde{z}\). As a result, \[\# \wp_{q,\omega} \leq \prod_{i=0}^r \#\{ \theta_i(z) \mid (y,z) \in \wp_{q,\omega} \}.\] Without loss of generality, we only consider the case \(z>0\) and fix some \((y_0,z_0) \in \wp_{q,\omega}\).

We note that \(\omega-qN \leq z^2 \leq \omega+qN\) for all \((y,z) \in \wp_{q,\omega}\). If \(\omega>2qN\), then \(|z|,|z_0|>\sqrt{\omega/2}\) and hence \[\left|\theta_0(z) - \theta_0(z_0) \right| \leq 1+\frac{ |z^2- z_0^2| }{ q |z+z_0| }= 1+\frac{|y-y_0|}{|z+z_0|} \lesssim 1+ \frac{N}{\sqrt \omega} \lesssim 1+ \sqrt {N/q}.\] If \(\omega \leq 2qN\), then \(|z| \lesssim \sqrt{qN}\) and hence \(|\theta_0(z)| \lesssim \sqrt{ N/q}\). Therefore, we know \(\theta_0(z)\) belongs to some interval of length \(O(\sqrt{N/q}+1)\) for all \((y,z) \in \wp_{q,\omega}\).

On the other hand, for any \((y,z) \in \wp_{q,\omega}\) we have \(q|(z^2-z_0^2)\), and hence \(p_i^{\alpha_i} | \theta_i(z-z_0)\theta_i(z+z_0)\) for \(1\leq i\leq r\). Consequently, we can find some \(\gamma_i\in\mathbb{N}\) which depends on \(z\), such that \(p^{\gamma_i}_i | \theta_i (z-z_0)\) and \(p_i^{\alpha_i-\gamma_i} | \theta_i(z+z_0)\). This further implies \[\theta_i(z) \in \big( \theta_i(z_0) + p_i^{\gamma_i} \mathbb{Z}\big) \cap \big(-\theta_i(z_0)+p_i^{\alpha_i-\gamma_i}\mathbb{Z}\big).\] Let us put \(\tilde{\gamma}_i =\max_{z} \min\{\gamma_i, \alpha_i-\gamma_i\}\). Notice \(p_i^{\tilde{\gamma}_i} | 2\theta_i(z_0)\) and \(p_i^{\max\{\gamma_i, \alpha_i-\gamma_i\}} \mathbb{Z}\subset p_i^{\alpha_i-\tilde{\gamma}_i}\mathbb{Z}\) , we see that at least one of \[\theta_i(z)\in \big( \theta_i(z_0) + p_i^{\alpha_i-\tilde{\gamma}_i} \mathbb{Z}\big) ,\qquad \theta_i(z)\in \big(-\theta_i(z_0)+p_i^{\alpha_i-\tilde{\gamma}_i}\mathbb{Z}\big)\] holds true. Thus \[\{ \theta_i(z) \mid (y,z) \in \wp_{q,\omega} \} \subset \big( \theta_i(z_0) + p_i^{\alpha_i-\tilde{\gamma}_i} \mathbb{Z}\big) \cup \big(-\theta_i(z_0)+p_i^{\alpha_i-\tilde{\gamma}_i}\mathbb{Z}\big),\] and \[\begin{align} \#\{ \theta_i(z) \mid (y,z) \in \wp_{q,\omega} \} &\leq \begin{cases} 2 p_i^{\tilde{\gamma}_i} , & \tilde{\gamma}_i\neq \alpha_i/2, \\ p_i^{\alpha_i/2}, & \tilde{\gamma}_i=\alpha_i/2, \end{cases} \\ &\leq p_i^{\alpha_i/2} \max \{1, 2p_i^{-1/2}\} . \end{align}\] As a result \(\# \wp_{q,\omega} \lesssim \sqrt N+\sqrt{\vphantom{N} q}\). ◻

2.7 Sharpness of Strichartz estimate↩︎

We now present several examples showing the sharpness of ?? .

Example 1: We take \[\phi_0(x) = N^{-3/2} \sum_{\xi\in \mathbb{Z}^3 \cap [-N,N]^3} \mathrm{e}^{2\pi\mathrm{i}\xi\cdot x}.\] It’s easy to calculate that \(\| \phi_0\|_{{L^2_x(\mathbb{T}^3)} } \approx 1\), while \(| \mathrm{e}^{\mathrm{i}t\Box} \phi_0(x) | \gtrsim N^{3/2}\) for \(|x| < 1/{(100N)}\) and \(0<t<1/{(100 N^2)}\). As a consequence, \[\| \mathrm{e}^{\mathrm{i}t \Box} \phi_0 \|_{{L_{t,x}^p([0,1]\times \mathbb{T}^3)}} \geq \left| \int_{\substack{|x|<\frac{1}{100 N},0<t<\frac{1}{100 N^2}}} | \mathrm{e}^{\mathrm{i}t\Box} \phi_0|^p \mathop{}\!\mathrm{d}t\mathop{}\!\mathrm{d}x \right|^{1/p} \gtrsim N^{\frac{3}{2} - \frac{5}{p} }.\] This example shows estimate ?? is sharp for \(p\geq 4.\)

In particular for \(p=4\), we consider \(\Omega_2(\hat{\phi}_0)\), for each \(\xi_1\), the number of choices of \(\xi_2\) such that \(\xi_1-\xi_2\in\mathrm{Cone}\) is bounded by \[\begin{align} \#(\mathrm{Cone}\cap[-N,N]^3) &\leq \sum_{M \leq N\text{ dyadic}} \#(\mathrm{Cone}_M\setminus \mathrm{Cone}_{M/2}) \\& \lesssim \sum_{M\leq N\text{ dyadic}} \frac{N}{M} \#(\mathrm{Cone}_{M}^{\mathrm{irr}} \setminus \mathrm{Cone}_{M/2}^{\mathrm{irr}}) \lesssim N\log N, \end{align}\] and \((\xi_3,\xi_4)\) lies on a plane passing through \(\xi_1\) with normal vector \(A(\xi_1-\xi_2)\), which gives \(O(N^2)\) choices, hence \[\Omega_2(\hat{\phi}_0) \lesssim \log N .\] Thus \(\Omega_1(\hat{\phi}_0)\) will give the major contribution in the \(L^4\) estimate.

Example 2: We take \[\phi_0(x) = N^{-1/2}\sum_{\xi =1}^N \mathrm{e}^{2\pi\mathrm{i}\xi x\cdot(1,1,0)}\] it’s not hard to see \(\| \phi_0\|_{{L^2_x(\mathbb{T}^3)}} \approx 1\). Note that \(\phi_0\) is invariant under the group \(\{ \mathrm{e}^{\mathrm{i}t\Box}\}_{t\in\mathbb{R}}\), hence \[|\mathrm{e}^{\mathrm{i}t\Box} \phi_0(x)| = |\phi_0(x)|\gtrsim N^{1/2} \text{ for } |x \cdot (1,1,0)|<\frac{1}{100N},\] which implies \[\| \mathrm{e}^{\mathrm{i}t\Box} \phi_0 \|_{{L_{t,x}^p([0,1]\times \mathbb{T}^3)} } \gtrsim N^{\frac{1}{2}-\frac{1}{p}}.\] This example shows the estimate ?? is sharp for \(p\in [2,4]\). Also for \(p=4\), we notice \(\Omega_1(\hat{\phi}_0)=0\), so \(\Omega_2(\hat{\phi}_0)\) gives the major contribution.

Now if we set \(S=\operatorname{supp} \hat{\phi}_0\), which is of size \(N\), then we get \[\| \mathrm{e}^{\mathrm{i}t\Box} \phi_0 \|_{{L_{t,x}^4([0,1]\times \mathbb{T}^3)} }/\|\phi_0\|_{{{L^2_x(\mathbb{T}^3)}}}\gtrsim (\# S)^\frac{1}{4}.\] Consequently, for the \(L^4\) estimate, we cannot obtain a non-trivial bound involving only \(\#S\) without resorting to the trivial relation \(\operatorname{diam}(S)\leq \#S\) (see Remark 3).

Example 3: We take \[\phi_{0,1}(y) = N^{-1/2} \sum_{\xi=1}^N \mathrm{e}^{2\pi\mathrm{i}\xi y},\quad y\in \mathbb{R}\] and \[\phi_0(x) = N^{-1} \sum_{\xi,\eta=1}^N \mathrm{e}^{2\pi\mathrm{i}(\xi,\xi,\eta)\cdot x} = \phi_{0,1}(x_1+x_2) \phi_{0,1}(x_3).\] Thus we see \(\| \phi_0 \|_{{L^2_x(\mathbb{T}^3)}} \approx 1\).

On the other hand, we have \[\begin{align} | \mathrm{e}^{\mathrm{i}t \Box}\phi_0(x)| &= |\phi_{0,1}(x_1+x_2) | |\mathrm{e}^{\mathrm{i}t \partial^2}\phi_{0,1}(x_3)| \\& \gtrsim N^{1/2} |\mathrm{e}^{\mathrm{i}{t}\partial^2} \phi_{0,1}(x_3)| \end{align}\] for \(|x_1+x_2|<1/(100N)\), therefore \[\begin{align} \| \mathrm{e}^{\mathrm{i}t \Box} \phi_0 \|_{{L_{t,x}^p([0,1]\times \mathbb{T}^3)}} &\gtrsim N^{\frac{1}{2}-\frac{1}{p}} \| \mathrm{e}^{\mathrm{i}t\partial^2} \phi_{0,1} \|_{{L_{t,x}^p([0,1]\times \mathbb{T}^{})}}\\& \gtrsim N^{\frac{1}{2}-\frac{1}{p}} \| \mathrm{e}^{\mathrm{i}t \partial^2}{ \phi_{0,1}} \|_{{L_{t,x}^2([0,1]\times \mathbb{T}^{})}}\\& \approx N^{\frac{1}{2}-\frac{1}{p}}. \end{align}\] This example also shows ?? is sharp for \(p\in[2,4].\) We also observe for \(p=4\), \(\Omega_1(\hat{\phi}_0)=0\).

Remark 16. We notice that the examples should easily generalize to the higher dimensional case. Example 1 also extends to irrational tori, while the validity of examples 2 and 3 on irrational tori depends on the equation.

3 Local Well-posedness↩︎

3.1 Function spaces↩︎

We use the adapted function spaces \(X^s,Y^s\), whose definitions are based on the \(U^p,V^p\) spaces. We will give their definitions and state the basic properties. We refer the readers to [24], [25] for detailed proofs of the following propositions, where a general theory can also be found.

Let \(\mathcal{H}\) be a separable Hilbert space over \(\mathbb{C}\); in this paper, this will be \(\mathbb{C}\) or \(H^s(\mathbb{T}^3)\). Let \(\mathcal{Z}\) be the set of finite partitions \(-\infty <t_0 <t_1 < \cdots < t_K \leq \infty\) of the real line.

Definition 17. Let \(1 \leq p < \infty\). For \(\{ t_k \}_{k=0}^K \in\mathcal{Z}\) and \(\{ \phi_k \}_{k=0}^{K-1} \subset \mathcal{H}\) with \(\sum_{k=1}^{K-1} \| \phi_k \|_{\mathcal{H}}^p =1\), we call a piecewise defined function \(a:\mathbb{R} \to \mathcal{H}\), \[a(t) = \sum_{k=1}^{K-1} \chi_{[t_k,t_{k+1})} \phi_k\] a \(U^p\)-atom, and we define the atomic space \(U^p(\mathbb{R}, \mathcal{H})\) of all functions \(u \colon \mathbb{R}\to\mathcal{H}\) such that \[u = \sum_j \lambda_j a_j ,\quadwitha_jareU^p-atoms, and\{\lambda_j\} \in \ell^1,\] with norm \[\| u\|_{U^p(\mathbb{R},\mathcal{H})} := \inf \left\{ \sum_j |\lambda_j| \Biggm| u = \sum_j \lambda_j a_j, \; a_j \text{ are } U^p\text{-atoms}\right\}.\]

Definition 18. Let \(1\leq p<\infty\), we define the space \(V^p(\mathbb{R},\mathcal{H})\) of functions \(v \colon \mathbb{R} \to \mathcal{H}\) such that \(\lim_{t\to-\infty} v(t)=0\) and the norm \[\| v \|_{V^p(\mathbb{R}, \mathcal{H})} := \sup_{\{t_k\}_{k=0}^K \in\mathcal{Z}} \left( \sum_{k=0}^{K-1} \| v(t_{k+1}) - v(t_k) \|_{\mathcal{H}}^p \right)^{1/p}\] is finite.

Corresponding to the linear flow generated by the group \(\{\mathrm{e}^{\mathrm{i}t\Box}\}_{t\in\mathbb{R}}\), we define the following.

Definition 19. For \(s\in \mathbb{R}\), we define the space \(U^p_\Box H^s\) \((resp., V^p_\Box H^s)\) of functions \(u \colon \mathbb{R} \to H^s(\mathbb{T}^3)\) such that \(t \mapsto \mathrm{e}^{-\mathrm{i}t\Box} u(t)\) is in \(U^p(\mathbb{R}, H^s(\mathbb{T}^3))\) \((resp., V^p(\mathbb{R}, H^s(\mathbb{T}^3)))\) with the norms \[\| u \|_{U^p_\Box H^s} := \| \mathrm{e}^{-\mathrm{i}t\Box} u \|_{U^p(\mathbb{R}, H^s(\mathbb{T}^3))} ,\quad \| u\|_{V^p_\Box H^s} := \| \mathrm{e}^{-\mathrm{i}t\Box} u \|_{V^p(\mathbb{R}, H^s(\mathbb{T}^3))} .\]

Due to the atomic structure of \(U^p\), we can extend bounded operators on \(L^2(\mathbb{T}^3)\) to \(U^p_\Box L^2\).

Proposition 20 ([24]). Let \(1\leq p< \infty\) and \(T_0 \colon L^2(\mathbb{T}^3) \times \cdots \times L^2(\mathbb{T}^3) \to L_{\mathrm{loc}}^1 ( \mathbb{R} \times \mathbb{T}^3)\) be a \(n\)-linear operator. If \[\| T_0( \mathrm{e}^{\mathrm{i}t\Box} \phi_1, \cdots , \mathrm{e}^{\mathrm{i}t\Box} \phi_n) \|_{L^p_{t,x}} \leq C_{T_0} \prod_{i=1}^n \| \phi_i \|_{{L^2_x(\mathbb{T}^3)}},\] then \(T_0\) extends to a \(n\)-linear operator \(T\) on \(U^p_\Box L^2 \times \cdots \times U_\Box^pL^2\), satisfying \[\| T ( u_1, \cdots , u_n) \|_{L^p_{t,x}} \lesssim C_{T_0} \prod_{i=1}^n \| u_i \|_{U^p_\Box L^2}.\]

The following corollary is a direct application of this proposition to our main result Theorem 1 and Remark 3.

Corollary 21. For \(u\in U^4_\Box L^2\), and any cube \(C\) of side length \(N\), we have \[\| P_{C} u \|_{L^4_{t,x}([0,1]\times \mathbb{T}^3)} \lesssim N^{1/4} \| u \|_{U^4_\Box L^2}. \label{Strichartz32estimate32in32U94495Box}\qquad{(6)}\]

Definition 22. For \(s\in\mathbb{R}\), we define the space \(X^s\) of functions \(u\colon \mathbb{R} \to H^s(\mathbb{T}^3)\) such that for every \(\xi \in \mathbb{Z}^3\) the mapping \(t \mapsto \mathrm{e}^{ -\mathrm{i}t h(\xi)} \widehat{u(t)}(\xi)\) is in \(U^2(\mathbb{R}, \mathbb{C} )\), with the norm \[\|u\|_{X^s} := \left( \sum_{\xi \in \mathbb{Z}^3} \left<\xi\right>^{2s} \| \mathrm{e}^{ -\mathrm{i}t h(\xi)} \widehat{u(t)}(\xi) \|_{U^2(\mathbb{R}, \mathbb{C})}^2 \right)^{1/2} .\]

Definition 23. For \(s\in\mathbb{R}\), we define the space \(Y^s\) of functions \(u\colon \mathbb{R} \to H^s(\mathbb{T}^3)\) such that for every \(\xi\in \mathbb{Z}^3\) the mapping \(t \mapsto \mathrm{e}^{ -\mathrm{i}t h(\xi)} \widehat{u(t)}(\xi)\) is in \(V^2(\mathbb{R}, \mathbb{C} )\), with the norm \[\|u\|_{Y^s} := \left( \sum_{\xi\in \mathbb{Z}^3} \left<\xi\right>^{2s} \| \mathrm{e}^{ -\mathrm{i}t h(\xi)} \widehat{u(t)}(\xi) \|_{V^2(\mathbb{R}, \mathbb{C})}^2 \right)^{1/2} .\]

Remark 24. We have the embeddings\[U^2_\Box H^s \hookrightarrow X^s \hookrightarrow Y^s \hookrightarrow V^2_\Box H^s \hookrightarrow U^q_\Box H^s\hookrightarrow L^\infty H^s,\quad \forall q\in(2,\infty).\]

Remark 25. For \(s\in \mathbb{R}\), and \(S_1,S_2\) are disjoint subsets of \(\mathbb{Z}^3\), we have \[\| P_{S_1\cup S_2} u\|_{Y^s} ^2 = \| P_{S_1} u\|_{Y^s}^2 + \|P_{S_2} u\|_{Y^s}^2.\]

For time interval \(I\subset \mathbb{R}\), we also consider the restriction spaces \(X^s(I),Y^s(I)\) with norms \[\| u \|_{X^s(I)} = \inf \{ \| \tilde{u} \|_{X^s} \mid \tilde{u}|_{I}=u \},\quad \| u \|_{Y^s(I)} = \inf \{ \| \tilde{u} \|_{Y^s} \mid \tilde{u}|_{I}=u \}.\]

Proposition 26 ([25]). Let \(s\in \mathbb{R}\) and \(T>0\). For \(\phi \in H^s(\mathbb{T}^3)\), we have \(\mathrm{e}^{\mathrm{i}t\Box} \phi \in X^s([0,T))\) and \[\| \mathrm{e}^{\mathrm{i}t\Box} \phi \|_{X^s([0,T))} \leq \| \phi \|_{H^s(\mathbb{T}^3)} .\] For \(f\in L^1 ([0,T); H^s(\mathbb{T}^3))\), we have the estimate for the Duhamel term. \[\begin{gather} \left\| \int_0^t \mathrm{e}^{\mathrm{i}(t-t')\Box} f(t') \mathop{}\!\mathrm{d}t' \right\|_{X^s([0,T))} \leq \sup_{\substack{v\in Y^{-s}([0,T)) \\ \|v\|_{Y^{-s}([0,T))}\leq 1} } \left| \iint_{[0,T)\times \mathbb{T}^3} f(t,x) \overline{v(t,x)} \mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}t \right| \end{gather}\]

Remark 27. The \(X^s([0,T))\) norm of the Duhamel term is also controlled by \(\| f \|_{L^1([0,T);H^s(\mathbb{T}^3))}\).

3.2 Multilinear estimates↩︎

We start from a bilinear estimate for frequency localized functions on \(\mathbb{T}^3\).

Proposition 28. For \(u_1,u_2 \in Y^0([0,1])\) with \(u_i = P_{N_i} u_i\), we have \[\| u_1 u_2 \|_{{L_{t,x}^2([0,1]\times \mathbb{T}^3)}} \lesssim \min\{ N_1,N_2 \}^{1/2} \| u_1\|_{Y^0([0,1])} \|u_2\|_{Y^0([0,1])}.\]

Proof. We may assume that \(N_1\leq N_2\). We decompose \(\mathbb{Z}^3 = \bigcup_j C_j\) into almost disjoint cubes with side length \(N_1\) and write \[u_1u_2 = \sum_{C_j} u_1 P_{C_j} u_2.\] Their Fourier supports are finitely overlapped, hence we have the almost orthogonality \[\begin{align} \| u_1u_2\|_{{L_{t,x}^2([0,1]\times \mathbb{T}^3)}}^2 &\approx \sum_j \| u_1 P_{C_j} u_2\|_{{L_{t,x}^2([0,1]\times \mathbb{T}^3)}}^2 \\& \leq \sum_j \| u_1 \|_{{L_{t,x}^4([0,1]\times \mathbb{T}^3)}}^2 \| P_{C_j} u_2 \|_{{L_{t,x}^4([0,1]\times \mathbb{T}^3)}}^2. \end{align}\] By Corollary ?? , Remark 25 and the embedding properties in Remark 24, \[\begin{align} \| u_1u_2\|_{{L_{t,x}^2([0,1]\times \mathbb{T}^3)}}^2 &\lesssim{} \sum_j N_1 \| u_1\|_{Y^0([0,1])}^2 \| P_{C_j} u_2 \|_{Y^0([0,1])}^2 \\& ={} N_1\| u_1\|_{Y^0([0,1])}^2 \| u_2 \|_{Y^0([0,1])}^2. \end{align}\] ◻

Now we are ready to show the key estimate on the nonlinear term by using duality argument combined with frequency decomposition, which helps to treat the nonlinearity in the fixed point argument.

Proposition 29. Let \(k\in \mathbb{N}_+\), \(s=\frac{3}{2} - \frac{1}{k}\) if \(k\geq 2\) and \(s>\frac{1}{2}\) if \(k=1\). Then for any \(0<T<1\), for \(u_1,\dots,u_{2k+1} \in {X^{s}([0,T))}\), we have \[\left\| \int_0^t \mathrm{e}^{\mathrm{i}(t-t')\Box} \prod_{i=1}^{2k+1} u_i \mathop{}\!\mathrm{d}t' \right\|_{{X^{s}([0,T))}} \lesssim_{s,k} \prod_{i=1}^{2k+1} \| u_i\|_{{X^{s}([0,T))}}.\] Here the implicit constant does not depend on T.

Proof. It suffices to show that for any \(u_0 \in Y^{-s}([0,T))\), we have \[\left| \int_{[0,T)\times \mathbb{T}^3} u_0\prod_{i=1}^{2k+1}u_i \mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}t \right| \lesssim \| u_0\|_{Y^{-s}([0,T))} \prod_{i=1}^{2k+1} \| u_i \|_{X^{s}([0,T))}.\] We apply Littlewood-Paley decomposition to each \(u_i\) to write \[u_i = \sum_{N_i\text{ dyadic}} P_{N_i}u_i = \sum_{N_i\text{ dyadic}} u_{N_{i}}^{(i)},\] hence it suffices to estimate \[\sum_{N_0,\dots,N_{2k+1}} \left| \int_{[0,T)\times \mathbb{T}^3} u_{N_{0}}^{(0)}\prod_{i=1}^{2k+1} u_{N_{i}}^{(i)} \mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}t \right|.\] In order to make the integral non-zero, we must have that the two highest frequencies are comparable. Due to symmetry, it’s harmless to assume \(N_1\geq N_2\geq \dots\geq N_{2k+1}\). Following Proposition 28 we have that \[\begin{align} &\left| \int_{[0,T)\times \mathbb{T}^3} u_{N_{0}}^{(0)}\prod_{i=1}^{2k+1} u_{N_{i}}^{(i)} \mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}t \right| \\\leq&{} \| u_{N_{0}}^{(0)}u_{N_{2}}^{(2)}\|_{{L_{t,x}^2([0,1]\times \mathbb{T}^3)}} \| u_{N_{1}}^{(1)} u_{N_{3}}^{(3)} \|_{{L_{t,x}^2([0,1]\times \mathbb{T}^3)}} \prod_{i\geq 4} \| u_{N_{i}}^{(i)} \|_{{L_{t,x}^\infty([0,1]\times \mathbb{T}^3)}} \\ \lesssim&{} \min\{ N_0,N_2\}^{\frac{1}{2}}_{\vphantom{3}} N_3^{\frac{1}{2}} \| u_{N_{0}}^{(0)}\|_{Y^0} \| u_{N_{1}}^{(1)}\|_{Y^0} \| u_{N_{2}}^{(2)}\|_{Y^0}\| u_{N_{3}}^{(3)}\|_{Y^0} \prod_{i\geq 4} N_i^{3/2} \| u_{N_{i}}^{(i)} \|_{Y^0} \\ \approx&{} \min\{ N_0,N_2\}^{\frac{1}{2}}_{\vphantom{3}} \frac{N_0^sN_3^{\frac{1}{2}-s}}{N_1^{s}N_2^s} \| u_{N_{0}}^{(0)}\|_{Y^{-s}} \| u_{N_{1}}^{(1)}\|_{Y^s} \| u_{N_{2}}^{(2)}\|_{Y^s}\| u_{N_{3}}^{(3)}\|_{Y^s} \prod_{i\geq 4} N_i^{\frac{3}{2}-s} \| u_{N_{i}}^{(i)} \|_{Y^s}. \end{align}\] For \(k=1\), since \(s>1/2\), we directly apply Cauchy-Schwarz to the summation over the two lower frequencies and the two highest frequencies respectively to the desired conclusion. For \(k\geq 2\), applying Cauchy-Schwarz to summation over \(N_i\) for \(i\geq 4\), we get that \[\begin{align} &\min\{ N_0,N_2\}^{\frac{1}{2}}_{\vphantom{3}} \frac{N_0^sN_3^{s-\frac{1}{2}}}{N_1^{s}N_2^s} \| u_{N_{0}}^{(0)}\|_{Y^{-s}} \| u_{N_{1}}^{(1)}\|_{Y^s} \| u_{N_{2}}^{(2)}\|_{Y^s}\| u_{N_{3}}^{(3)}\|_{Y^s} \prod_{i\geq 4} \| u_i\|_{{X^{s}([0,T))}} \\ \leq{}& \left( \frac{N_0}{N_1} \right)^{s} \left( \frac{ N_3 }{N_2} \right)^{s-\frac{1}{2}}\| u_{N_{0}}^{(0)}\|_{Y^{-s}} \| u_{N_{1}}^{(1)}\|_{Y^s} \| u_{N_{2}}^{(2)}\|_{Y^s}\| u_{N_{3}}^{(3)}\|_{Y^s} \prod_{i\geq 4} \| u_i\|_{{X^{s}([0,T))}}. \end{align}\] Then apply Cauchy-Schwarz to the summation over the two lower frequencies and the two highest frequencies respectively, we get the desired conclusion. ◻

3.3 Proof of Theorem 4↩︎

The proof is a standard contraction argument as in [3], [25]. Given initial data \(\phi \in H^s(\mathbb{T}^3)\), with \(\| \phi \|_{H^s(\mathbb{T}^3)} \leq A\), suppose \(\delta\) is a small constant depending on \(A\), and \(N\) is a large number depending on \(\phi\) and \(\delta\) such that \(\|P_{>N}\phi \|_{H^s(\mathbb{T}^3)} \leq \delta\), we will show the Picard iteration mapping given by \[\mathcal{I}(u)(t) := \mathrm{e}^{\mathrm{i}t\Box}\phi \mp \mathrm{i}\int_0^t \mathrm{e}^{\mathrm{i}(t-t') \Box} |u|^{2k}u \mathop{}\!\mathrm{d}t'.\] is a contraction on the set \[\begin{gather} D : = \{ u \in C([0,T);H^s(\mathbb{T}^3)) \cap X^s([0,T)) \mid \\ u(0)=\phi,\;\|u\|_{X^s([0,T))} \leq 2A,\;\| P_{>N} u \|_{X^s([0,T))} \leq 2\delta \}, \end{gather}\] under the metric \[d(u,v) := \| u-v \|_{X^s([0,T))}\] provided \(T\) is chosen sufficiently small (depending on \(A\), \(\delta\), \(N\) and \(k\)).

For \(u,v\in D\), we can decompose \[|u|^{2k}u-|v|^{2k}v = F_1(u,v) + F_2(u,v),\] where \(F_1(u,v)\) is a combination of \(u-v,P_{\leq N}u,P_{\leq N}v\), and all terms involving \(P_{>N}u,P_{>N}v\) appear in \(F_2(u,v)\). Employing Sobolev embeddings and [26], we estimate that \[\begin{align} & \left\| \int_0^t \mathrm{e}^{\mathrm{i}(t-t')\Box} F_1(u,v) \mathop{}\!\mathrm{d}t' \right\|_{X^{s}([0,T))} \leq CT \| F_1(u,v)\|_{L^\infty H^s} \\ \leq{}& CT \left( \| u-v \|_{L^\infty H^s} \left(\| P_{\leq N} u\|_{L^\infty_{t,x}}^{2k} + \| P_{\leq N} v \|_{L^\infty_{t,x}}^{2k}\right) \right.\\ &\qquad + N^s \| u-v\|_{L_t^{\infty\vphantom{/}} L_{x\vphantom{t}}^{6/(3-2s)}} \left.\left (\| P_{\leq N} u\|_{L_t^{\infty\vphantom{/}} L_{x\vphantom{t}}^{6k/s}}^{2k}+ \| P_{\leq N} v \|_{L_t^{\infty\vphantom{/}} L_{x\vphantom{t}}^{6k/s}}^{2k} \right)\right) \\ \leq{}& CTN^{k(3-2s)} (2A)^{2k}\| u-v \|_{{X^{s}([0,T))}} .\label{est:contraction-low} \end{align}\tag{26}\] While by Proposition 29, it holds that \[\begin{align} & \left\| \int_0^t \mathrm{e}^{\mathrm{i}(t-t')\Box} F_2(u,v) \mathop{}\!\mathrm{d}t' \right\|_{X^{s}([0,T))} \\ \leq {}& C \| u-v\|_{X^s} ( \| P_{>N} u \|_{X^s} + \| P_{>N}v \|_{X^s} )( \| u \|_{X^s} + \| v \|_{X^s} )^{2k-1} \\ \leq{}& C(2A)^{2k-1}(2\delta) \| u-v\|_{X^{s}([0,T))}. \label{est:contraction-high} \end{align}\tag{27}\] Hence we get that \[\left\| \mathcal{I}(u) - \mathcal{I}(v) \right\|_{X^s} \leq \frac{1}{10} \| u-v\|_{X^s}, \label{contraction32on32T943}\tag{28}\] provided \(\delta\) is chosen sufficiently small depending on \(A, k\), and \(T\) is chosen sufficiently small depending on \(A,N\) and \(k\).

Next we verify that \(\mathcal{I}\) maps \(D\) into itself. For constant \(C\) large enough, we have \[\begin{align} \left\| \int_0^t \mathrm{e}^{\mathrm{i}(t-t') \Box } |P_{\leq N} u |^{2k} P_{\leq N} u \mathop{}\!\mathrm{d}t' \right\|_{X^s} & \leq CT \left\| |P_{\leq N} u |^{2k} P_{\leq N} u \right\| _{L^\infty H^s} \\& \leq C T \| P_{\leq N} u \|_{L^\infty_{t,x}}^{2k} \| P_{\leq N }u\|_{X^s} \\& \leq C T N^{k(3-2s)}(2A)^{2k+1} ,\label{est:Duhamel32low} \end{align}\tag{29}\] and apply 28 for \(v=P_{\leq N}u\) to get that \[\begin{align} \| \mathcal{I}(u) - \mathcal{I}(P_{\leq N}u) \|_{X^s} \leq \frac{1}{10} \| P_{>N} u \|_{X^s} \leq \frac{\delta}{5}.\label{est:diffwithlow} \end{align}\tag{30}\] To control \(P_{>N} \mathcal{I}(u)\), notice at least one input in the nonlinear term should have high frequency \(\frac{N}{2k+1}\), thus applying 29 30 we get \[\begin{align} \| P_{>N} \mathcal{I}(u)\|_{X^s}\lesssim \|P_{>N}\mathrm{e}^{\mathrm{i}t\Box}\phi\|_{X^s}+ \|P_{>N}\Big(\mathcal{I}(u) - \mathcal{I}(P_{\leq \frac{N}{2k+1}}u)\Big) \|_{X^s}\leq 2\delta. \end{align}\]

To summarize, provided \(\delta\) is chosen sufficiently small depending on \(A, k\), and \(T\) is chosen sufficiently small depending on \(A, N\) and \(k\), we have \[\| \mathcal{I}(u)\|_{X^s} \leq \| \mathrm{e}^{\mathrm{i}t\Box}\phi \|_{X^s} +A\leq 2A,\quad \| P_{>N} \mathcal{I}(u)\|_{X^s} \leq \|P_{>N} \mathrm{e}^{\mathrm{i}t\Box}\phi \|_{X^s} +\delta \leq 2\delta.\]

As for the uniqueness in the whole space \(C([0,T);H^s(\mathbb{T}^3)) \cap X^s([0,T))\), supposing that we have two functions \(u,v\) which both solve the equations 5 with the same initial data \(\phi\), we can choose \(A'\) sufficiently large, \(\delta'\) sufficiently small and \(N'\) sufficiently large such that \(u,v\) are both contained in some \(D=D_{A',N',\delta'}\). By the iteration, we know that there exists some \(T'\) (maybe much smaller than \(T\) given above) such that \(u(t)=v(t)\) for \(t\in[0,T')\). Uniqueness in the whole space \(C([0,T);H^s(\mathbb{T}^3)) \cap X^s([0,T))\) follows from a continuity argument.

3.4 Proof of Theorem 5↩︎

We prove the ill-posedness of the cubic HNLS on \(H^{1/2}(\mathbb{T}^3)\) by showing the first Picard iteration is unbounded. Let us pick \[\phi_N(x) = \sum_{k=1}^N \frac{ \mathrm{e}^{2\pi\mathrm{i}(k,k,0)\cdot x}}{k}.\] It is easy to see \(\| \phi_N \|_{H^{1/2}(\mathbb{T}^3)} \approx (\log N)^{1/2}\). Notice that \(\Box \phi_N = 0\), so \(\phi_N\) (also \(|\phi_N|^{2}\phi_N\)) is invariant under the group \(\{ \mathrm{e}^{\mathrm{i}t\Box} \}_{t\in\mathbb{R}}\), thus \[\mathcal{I}(\mathrm{e}^{\mathrm{i}t\Box} \phi_N) (t) = \mathcal{I}(\phi_N)(t) = \phi_N \pm \mathrm{i}t |\phi_N|^{2}\phi_N. \label{1st-iteration}\tag{31}\] It suffices to show that \(\| |\phi_N|^2 \phi_N \|_{H^{1/2}(\mathbb{T}^3)} \gtrsim \log N \| \phi_N \|^3_{H^{1/2}(\mathbb{T}^3)}\). Since \[| \phi_N|^2\phi_N(x) = \sum_k \mathrm{e}^{2\pi\mathrm{i}(k,k,0)\cdot x} \sum_{k_1-k_2+k_3=k} \frac{1}{k_1k_2k_3},\] we consider the set \[\Gamma(k) = \{ (k_1,k_2,k_3) \in \mathbb{Z}^3 \mid k_3= k-k_1+k_2,\;1\leq k_1,k_2 \leq k/4 \}\] for \(k\) positive and sufficiently large. Then \(k/2 \leq k_3 \leq 3k/2\) for \((k_1,k_2,k_3)\in \Gamma(k)\). Hence \[\sum_{\substack{{k_1-k_2+k_3=k}\\ 1\leq k_i \leq N}} \frac{1}{k_1k_2k_3} \geq \sum_{\Gamma(k)}\frac{1}{k_1k_2k_3} \approx \frac{1}{k} \sum_{k_1=1}^{k/4} \frac{1}{k_1} \sum_{k_2=1}^{k/4}\frac{1}{k_2} \approx \frac{(\log k)^2}{k},\] for \(1\lesssim k\lesssim N\), and \[\Big\| \phi_N |\phi_N|^2 \Big\|_{H^{1/2}(\mathbb{T}^2)} \gtrsim \left( \sum_{1\lesssim k\lesssim N} k\cdot \frac{(\log k)^4}{k^2}\right)^{1/2} \approx (\log N)^{5/2}\approx \log N \| \phi_N \|^3_{H^{1/2}(\mathbb{T}^3)},\] this finishes the proof. 4

References↩︎

[1]
Jean Bourgain. Fourier transform restriction phenomena for certain lattice subsets and applications to nonlinear evolution equations. I. Schrödinger equations. Geometric and Functional Analysis, 3(2):107–156, 1993.
[2]
Jean Bourgain and Ciprian Demeter. The proof of the \(l^2\) decoupling conjecture. Annals of Mathematics. Second Series, 182(1):351–389, 2015.
[3]
Rowan Killip and Monica Vişan. Scale invariant Strichartz estimates on tori and applications. Mathematical Research Letters, 23(2):445–472, 2016.
[4]
Sebastian Herr and Beomjong Kwak. Strichartz estimates and global well-posedness of the cubic NLS on \(\mathbb{T}^2\). Forum of Mathematics, Pi, 12:e14, 2024.
[5]
Jean Bourgain and Ciprian Demeter. Decouplings for curves and hypersurfaces with nonzero Gaussian curvature. Journal d’Analyse Mathématique, 133:279–311, 2017.
[6]
M. Ablowitz and S. Butler. Nonlinear temporal-spatial surface plasmon polaritons. Optics Communications, 330:49–55, 2014.
[7]
J. Myra and C. Liu. Self-modulation of ion bernstein waves. Phys. Fluids, 23:2258–2264, 1980.
[8]
N. Pereira, A. Sen, and A. Bers. Nonlinear development of lower hybrid cones. Phys. Fluids, 21:117–120, 1978.
[9]
L. Bergé. Wave collapse in physics: principles and applications to light and plasma waves. Phys.Rep., 303:259–370, 1998.
[10]
Walter Craig, Ulrich Schanz, and Catherine Sulem. The modulational regime of three-dimensional water waves and the Davey-Stewartson system. Annales de l’Institut Henri Poincaré C. Analyse Non Linéaire, 14(5):615–667, 1997.
[11]
A. Davey and K. Stewartson. On three-dimensional packets of surface waves. Proc. Roy. Soc. London Ser. A, 338:101–110, 1974.
[12]
Jean-Michel Ghidaglia and Jean-Claude Saut. On the initial value problem for the Davey-Stewartson systems. Nonlinearity, 3(2):475–506, 1990.
[13]
Nathan Totz. A justification of the modulation approximation to the 3d full water wave problem. Communications in Mathematical Physics, 335(1):369–443, 2015.
[14]
V.E. Zakharov and E.A. Kuznetsov. Hamiltonian formalism for nonlinear waves. Phys.Usp., 40(11):1087–1116, 1997.
[15]
Jean-Claude Saut and Yuexun Wang. On the hyperbolic nonlinear Schrödinger equations. Adv. Contin. Discrete Models, pages Paper No. 15, 12, 2024.
[16]
Nicolas Godet and Nikolay Tzvetkov. Strichartz estimates for the periodic non-elliptic Schrödinger equation. Comptes Rendus Mathématique. Académie des Sciences. Paris, 350(21-22):955–958, 2012.
[17]
Yuzhao Wang. Periodic cubic Hyperbolic Schrödinger equation on \(\mathbb T^ 2\). Journal of Functional Analysis, 265(3):424–434, 2013.
[18]
Engin Başakoğlu, Tadahiro Oh, and Yuzhao Wang. Sharp unconditional well-posedness of the 2-\(d\) periodic cubic hyperbolic nonlinear schrödinger equation, 2025.
[19]
Engin Başakoğlu, Chenmin Sun, Nikolay Tzvetkov, and Yuzhao Wang. Hyperbolic nonlinear schrödinger equations on \(\mathbb{R}\times \mathbb{T}\), 2025.
[20]
Endre Szemerédi and William T. Trotter, Jr. Extremal problems in discrete geometry. Combinatorica, 3(3-4):381–392, 1983.
[21]
János Pach and Micha Sharir. Geometric incidences. In Towards a theory of geometric graphs, volume 342 of Contemp. Math., pages 185–223. Amer. Math. Soc., Providence, RI, 2004.
[22]
Kenneth L. Clarkson, Herbert Edelsbrunner, Leonidas J. Guibas, Micha Sharir, and Emo Welzl. Combinatorial complexity bounds for arrangements of curves and spheres. Discrete Comput. Geom., 5(2):99–160, 1990.
[23]
Roel Apfelbaum and Micha Sharir. Repeated angles in three and four dimensions. SIAM Journal on Discrete Mathematics, 19(2):294–300, 2005.
[24]
Martin Hadac, Sebastian Herr, and Herbert Koch. Well-posedness and scattering for the KP-II equation in a critical space. Annales de l’Institut Henri Poincaré C, Analyse Non Linéaire, 26(3):917–941, 2009.
[25]
Sebastian Herr, Daniel Tataru, and Nikolay Tzvetkov. Global well-posedness of the energy-critical nonlinear Schrödinger equation with small initial data in \(H^1 (\mathbb T^3)\) . Duke Mathematical Journal, 159(2):329–349, 2011.
[26]
Carlos E. Kenig, Gustavo Ponce, and Luis Vega. Well-posedness and scattering results for the generalized Korteweg-de Vries equation via the contraction principle. Communications on Pure and Applied Mathematics, 46(4):527–620, 1993.

  1. 2020 AMS Mathematics Subject Classification. 35Q55.↩︎

  2. Keywords: Hyperbolic Schrödinger equation, Strichartz estimate, local well-posedness.↩︎

  3. The authors are supported by the NSF of China (No. 12571254, 12341102).↩︎

  4. The construction presented here is essentially two-dimensional. Therefore, it can also be used to prove the ill-posedness of the 2D cubic HNLS for initial data in \(H^{1/2}(\mathbb{T}^2)\), in the same sense as Theorem 5.↩︎