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

LpLqL^{p}-L^{q} estimates for solutions to the plate equation with mass term

Alexandre Arias Junior Alexandre A. Junior (alexandre.ariasjunior@usp.br) Department of Computer Science and Mathematics (FFCLRP), University of São Paulo (USP),
Ribeirão Preto, SP, 14040-901, Brazil
Halit Sevki Aslan Halit S. Aslan (halitsevkiaslan@gmail.com) Department of Computer Science and Mathematics (FFCLRP), University of São Paulo (USP),
Ribeirão Preto, SP, 14040-901, Brazil
Antonio Lagioia Antonio Lagioia (antonio.lagioia@uniba.it) Department of Mathematics, University of Bari, Via E. Orabona 4, 70125 Bari, Italy Marcelo Rempel Ebert Marcelo R. Ebert (ebert@ffclrp.usp.br) (corresponding author) Department of Computer Science and Mathematics (FFCLRP), University of São Paulo (USP),
Ribeirão Preto, SP, 14040-901, Brazil
Abstract

In this paper, we study the Cauchy problem for the linear plate equation with mass term and its applications to semilinear models. For the linear problem we obtain LpLqL^{p}-L^{q} estimates for the solutions in the full range 1pq1\leq p\leq q\leq\infty, and we show that such estimates are optimal. In the sequel, we discuss the global in time existence of solutions to the associated semilinear problem with power nonlinearity |u|α|u|^{\alpha}. For low dimension space n4n\leq 4, and assuming L1L^{1} regularity on the second datum, we were able to prove global existence for α>max{αc(n),α~c(n)}\alpha>\max\{\alpha_{c}(n),\tilde{\alpha}_{c}(n)\} where αc=1+4/n\alpha_{c}=1+4/n and α~c=2+2/n\tilde{\alpha}_{c}=2+2/n. However, assuming initial data in H2(n)×L2(n)H^{2}(\mathbb{R}^{n})\times L^{2}(\mathbb{R}^{n}), the presence of the mass term allows us to obtain global in time existence for all 1<α(n+4)/[n4]+1<\alpha\leq(n+4)/[n-4]_{+}. We also show that the latter upper bound is optimal, since we prove that there exist data such that a non-existence result for local weak solutions holds when α>(n+4)/[n4]+\alpha>(n+4)/[n-4]_{+}.

Keywords: plate equation; Cauchy problem; decay estimates; optimal estimates; global existence of solution; local existence of solution.

AMS Classification (2020) 35L75; 35L30; 35A01; 35G25.

1 Introduction

We consider the forward Cauchy problem for the plate equation with mass

{utt+Δ2u+u=0,(t,x)(0,)×n,u(0,x)=u0(x),ut(0,x)=u1(x),xn.\begin{cases}u_{tt}+\Delta^{2}u+u=0,&\,\,\,(t,x)\in(0,\infty)\times\mathbb{R}^{n},\\ u(0,x)=u_{0}(x),\quad u_{t}(0,x)=u_{1}(x),&\,\,\,x\in\mathbb{R}^{n}.\end{cases} (1.1)

This is an important model in the literature, known as beam operator and plate operator in the case of space dimension n=1n=1 and n=2n=2, respectively. Models to analyse the vibrations of thin plates (n=2n=2) given by the full von Kármán system have been studied by several authors, see [6, 21, 28]. We mention that some plate models include also a term Δutt-\Delta u_{tt} called rotational inertia. Energy estimates for solutions, for which a regularity-loss type decay appears, have been investigated in [3, 4, 5, 14, 32]. We address the interested reader to [9, 10, 19, 27] for LpLqL^{p}-L^{q} estimates for solutions to structurally damped plate equation and more general evolution equations. Strichartz estimates and estimates in modulations spaces for the plate equation were obtained in [7, 8].
Equations whose “principal part” is wtt+(Δ)σw=0,w_{tt}+(-\Delta)^{\sigma}w=0, like the plate equation when σ=2\sigma=2, are called σ\sigma-evolution equations in the sense of Petrowsky, since their symbols τ2+|ξ|2σ\tau^{2}+|\xi|^{2\sigma} have only pure imaginary and distinct roots τ=±i|ξ|σ\tau=\pm i|\xi|^{\sigma} for all ξ0\xi\neq 0. The set of 1-evolution operators coincides with the set of strictly hyperbolic operators. However, several properties of the hyperbolic operators are missing when σ1\sigma\neq 1.

It is well-known that the Cauchy problem (1.1) is well-posed in the energy space, namely, if (u0,u1)H2(n)×L2(n)(u_{0},u_{1})\in H^{2}(\mathbb{R}^{n})\times L^{2}(\mathbb{R}^{n}), then there exists a uniquely determined solution u𝒞([0,T],H2(n))𝒞1([0,T],L2(n))u\in\mathcal{C}([0,T],H^{2}(\mathbb{R}^{n}))\cap\mathcal{C}^{1}([0,T],L^{2}(\mathbb{R}^{n})) to (1.1) for all T>0T>0. Moreover, its energy is conserved (see [12]), that is,

E(u)(0)=E(u)(t):=12n|ut|2+|Δu|2+|u|2dx.E(u)(0)=E(u)(t):=\frac{1}{2}\int_{\mathbb{R}^{n}}|u_{t}|^{2}+|\Delta u|^{2}+|u|^{2}dx.

As it happens in several equations with oscillatory multipliers, to derive dispersive estimates to (1.1), one has to deal with oscillatory integrals. If u00u_{0}\neq 0, without additional regularity on the initial data, LpLqL^{p}-L^{q} estimates for solutions hold only for p=q=2p=q=2.

In [23] the authors considered the Cauchy problem for the Klein-Gordon equation

{uttΔu+u=0,(t,x)(0,)×n,u(0,x)=0,ut(0,x)=u1(x),xn,\begin{cases}u_{tt}-\Delta u+u=0,&\,\,\,(t,x)\in(0,\infty)\times\mathbb{R}^{n},\\ u(0,x)=0,\quad u_{t}(0,x)=u_{1}(x),&\,\,\,x\in\mathbb{R}^{n},\end{cases}

and proved that the operator Tt:u1u(t,)T_{t}:u_{1}\rightarrow u(t,\cdot) is bounded from LpL^{p} into LqL^{q} if, and only if,

n(1p1q)+(n1)max{(121p),(1q12)}1,n\left(\frac{1}{p}-\frac{1}{q}\right)+(n-1)\max\left\{\left(\frac{1}{2}-\frac{1}{p}\right),\left(\frac{1}{q}-\frac{1}{2}\right)\right\}\leq 1, (1.2)

with the exception for (p,q)=(1,2)(p,q)=(1,2) and (p,q)=(2,)(p,q)=(2,\infty) if n=2n=2. We refer to [26, 31] for LpLqL^{p}-L^{q} estimates to the solutions of the Cauchy problem for the free wave equation with the first datum zero.

Our first goal in this paper is to derive LpLqL^{p}-L^{q} estimates for the solutions to the Cauchy problem for the plate equation with mass term

{utt+Δ2u+u=0,(t,x)(0,)×n,u(0,x)=0,ut(0,x)=u1(x),xn.\begin{cases}u_{tt}+\Delta^{2}u+u=0,&\,\,\,(t,x)\in(0,\infty)\times\mathbb{R}^{n},\\ u(0,x)=0,\quad u_{t}(0,x)=u_{1}(x),&\,\,\,x\in\mathbb{R}^{n}.\end{cases} (1.3)

In the massless case, if Kpl(t,)K_{\text{pl}}(t,\cdot) denotes the fundamental solution to the free plate equation, then for 1pq1\leq p\leq q\leq\infty the operator

Kpl(t,):u1u(t,)=Kpl(t,)u1,K_{\text{pl}}(t,\cdot)\ast:u_{1}\mapsto u(t,\cdot)=K_{\text{pl}}(t,\cdot)\ast u_{1},

is bounded from Lp(n)L^{p}(\mathbb{R}^{n}) into Lq(n)L^{q}(\mathbb{R}^{n}) if, and only if, dpl(p,q)1d_{\text{pl}}(p,q)\leq 1 with p>1p>1 and q<q<\infty or if, and only if, dpl(p,q)<1d_{\text{pl}}(p,q)<1 with p=1p=1 and q=q=\infty (see Theorem 4.1 in  [24]), where

dpl(p,q)=n2(1p1q)+nmax{(121p),(1q12)}.d_{\text{pl}}(p,q)=\frac{n}{2}\left(\frac{1}{p}-\frac{1}{q}\right)+n\max\left\{\left(\frac{1}{2}-\frac{1}{p}\right),\left(\frac{1}{q}-\frac{1}{2}\right)\right\}. (1.4)

In this range, by homogeneity, it follows

u(t,)Lqt1n2(1p1q)u1Lp,t>0.\|u(t,\cdot)\|_{L^{q}}\lesssim t^{1-\frac{n}{2}\left(\frac{1}{p}-\frac{1}{q}\right)}\|u_{1}\|_{L^{p}},\quad t>0.

One of the reasons that nn in dpl(p,q)d_{\text{pl}}(p,q) replaces n1n-1 in (1.2) is due to the fact that the Hessian of |ξ|σ|\xi|^{\sigma}, with σ>0\sigma>0, is singular if, and only if, σ=1\sigma=1 (see, for instance, [29]). Due to this, without difficulties the obtained results in this paper may be extended to the Cauchy problem for the equation utt+(Δ)σu+u=0u_{tt}+(-\Delta)^{\sigma}u+u=0, for σ>0\sigma>0, σ1\sigma\neq 1.

As in the massless case, the condition dpl(p,q)1d_{\text{pl}}(p,q)\leq 1 come into play to derive LpLqL^{p}-L^{q} estimates for solutions at high frequencies in the phase space, but since for (1.3) we no longer have the homogeneity property, a challenging problem is to understand the influence of the mass term on the long-time behaviour on the solutions.

A partial answer to this question was obtained in the dual line, namely, the following estimate holds for solutions to (1.3) for all 2q2:=2n[n4]+2\leq q\leq 2^{**}:=\frac{2n}{[n-4]_{+}} (see [22]):

u(t,)Lqu1Lq{tn4+n2qt1,t1n2(12q)t(0,1),\|u(t,\cdot)\|_{L^{q}}\lesssim\|u_{1}\|_{L^{q^{{}^{\prime}}}}\begin{cases}t^{-\frac{n}{4}+\frac{n}{2q}}&t\geq 1,\\ t^{1-\frac{n}{2}\left(1-\frac{2}{q}\right)}&t\in(0,1),\end{cases}

where qq^{\prime} denotes the conjugate exponent of qq. Besides, when n5n\geq 5 the above estimate was extended for 2q<2\leq q<\infty (see [15]). Using Fourier analysis, stationary phase method and the techniques from WKB analysis (see [10], [12]), in Theorem 2.1 we extended the previous results for 1pq1\leq p\leq q\leq\infty out of the dual line, provided that the condition dpl(p,q)1d_{\text{pl}}(p,q)\leq 1 is satisfied. Our approach is different from [23] and produce sharp estimates to (1.3), as it happens for the free plate equation [13] and to the structurally damped plate equation [10]. However, it is not suitable to get sharp estimates to the free wave equation and to the Klein-Gordon equation. Nevertheless, one may combine the approach used in this paper together with the obtained behaviour for wave type kernels around the unit sphere (see [29], Lemma 5.1) in order to produce sharp estimates.

The second purpose of this paper is to study the long time behaviour of global (in time) solutions for the semi-linear problem

{utt+Δ2u+u=f(u),(t,x)(0,)×n,u(0,x)=u0(x),ut(0,x)=u1(x),xn,\begin{cases}u_{tt}+\Delta^{2}u+u=f(u),&\,\,\,(t,x)\in(0,\infty)\times\mathbb{R}^{n},\\ u(0,x)=u_{0}(x),\quad u_{t}(0,x)=u_{1}(x),&\,\,\,x\in\mathbb{R}^{n},\end{cases} (1.5)

where f(u)=|u|αf(u)=|u|^{\alpha}, with α>1\alpha>1. The local well-posedness for 1<α<n+4[n4]+1<\alpha<\frac{n+4}{[n-4]_{+}} and scattering with small data for 1+8n<α<n+4[n4]+1+\frac{8}{n}<\alpha<\frac{n+4}{{[n-4]}_{+}} to (1.5) were studied in the energy space 𝒞([0,T),H2(n))𝒞1([0,T),L2(n))\mathcal{C}([0,T),H^{2}(\mathbb{R}^{n}))\cap\mathcal{C}^{1}([0,T),L^{2}(\mathbb{R}^{n})) in [22]. We refer to the survey [16] and references therein for further results about well-posedness, blow-up in finite time, long time existence, and the existence of uniform bounds for global solutions to (1.5).

The nonlinearity in (1.5) may have several shapes, for instance, the derived results in this paper also hold if f(u)=|u|α1uf(u)=|u|^{\alpha-1}u or if ff is locally Lipschitz-continuous satisfying

f(0)=0,|f(u)f(v)|C|uv|(|f(u)|α1+|f(v)|α1),f(0)=0,\quad|f(u)-f(v)|\leq C|u-v|(|f(u)|^{\alpha-1}+|f(v)|^{\alpha-1}),

for some α>1\alpha>1. Mainly, we are interested in nonlinearities that can not be included in some energy, so it may create blow-up in finite time. These kind of non-linearities are known in the literature as of source type.Semi-linear plate equations with source nonlinearity, with or without damping, have been investigated by many authors, see for instance [9, 11, 13, 14, 27] and references therein. For the semi-linear model (1.5) with zero mass, one may have blow-up for small power nonlinearities, even assuming small data in the energy space.

The outline of our study of the Cauchy problem (1.5) is the following:

  • We apply the derived estimates in Theorem 2.1 to discuss the local (in time) existence of Sobolev solutions, even for large data (u0,u1)H2(n)×L2(n)(u_{0},u_{1})\in H^{2}(\mathbb{R}^{n})\times L^{2}(\mathbb{R}^{n}) (see Theorem 2.2).

  • In Theorem 2.4, under the assumptions u00u_{0}\equiv 0 and u1Lm(n)u_{1}\in L^{m}(\mathbb{R}^{n}), with m[1,2]m\in[1,2], we prove a non-existence result for local weak solution to (1.5) when n>2mn>2m and α>n+2mn2m\alpha>\frac{n+2m}{n-2m}.

  • In Theorems 2.5 and 2.6 the global (in time) existence of solutions are obtained under the assumption of small initial datum u1L1(n)u_{1}\in L^{1}(\mathbb{R}^{n}) for n4n\leq 4 and α>max{αc(n),α~c(n)}\alpha>\max\{\alpha_{c}(n),\tilde{\alpha}_{c}(n)\}, where such exponents are defined in (2.7).

Notations:

  • fgf\lesssim g means that there exists a positive constant CC fulfilling fCgf\leqslant Cg, which may be changed in different lines, analogously, for fgf\gtrsim g. Furthermore, the asymptotic relation fgf\simeq g holds when gfgg\lesssim f\lesssim g.

  • [a]+[a]_{+} and [a][a]_{-} denote max{a,0}\max\{a,0\} and max{a,0}\max\{-a,0\}, respectively.

  • pp^{\prime} stands for the conjugate exponent of a given p[1,]p\in[1,\infty].

  • Lpq=Lpq(n)L_{p}^{q}=L_{p}^{q}(\mathbb{R}^{n}) denotes the space of tempered distributions T𝒮(n)T\in\mathcal{S}^{\prime}(\mathbb{R}^{n}) such that TfLq(n)T\ast f\in L^{q}(\mathbb{R}^{n}) for any f𝒮(n)f\in\mathcal{S}(\mathbb{R}^{n}), where 𝒮(n)\mathcal{S}(\mathbb{R}^{n}) stands for the space of rapidly decreasing Schwartz functions, and

    TfLqCfLp\|T\ast f\|_{L^{q}}\leq C\|f\|_{L^{p}}

    for all f𝒮(n)f\in\mathcal{S}(\mathbb{R}^{n}) with a constant C>0C>0, which is independent of ff. In this case, the operator TT\ast is extended by density to the space Lp(n)L^{p}(\mathbb{R}^{n}).

  • Mpq=Mpq(n)M_{p}^{q}=M_{p}^{q}(\mathbb{R}^{n}) stands for the set of Fourier transforms T^\hat{T} of distributions TLpqT\in L_{p}^{q}, equipped by the norm

    mMpq:=sup{1(m(f))Lq:f𝒮(n),fLp=1},\|m\|_{M_{p}^{q}}:=\sup\big{\{}\big{\|}\mathcal{F}^{-1}\big{(}m\mathcal{F}(f)\big{)}\big{\|}_{L^{q}}:f\in\mathcal{S}(\mathbb{R}^{n}),\|f\|_{L^{p}}=1\big{\}},

    and we set Mp=MppM_{p}=M_{p}^{p}. The elements in MpqM_{p}^{q} are called multipliers of type (p,q)(p,q).

  • We fix a nonnegative function ϕ𝒞(n)\phi\in\mathcal{C}^{\infty}(\mathbb{R}^{n}) with compact support in {ξn:21|ξ|2}\{\xi\in\mathbb{R}^{n}:2^{-1}\leq|\xi|\leq 2\} such that

    k=+ϕk(ξ)=1,whereϕk(ξ):=ϕ(2kξ).\sum_{k=-\infty}^{+\infty}\phi_{k}(\xi)=1,\qquad\text{where}\,\,\,\phi_{k}(\xi):=\phi(2^{-k}\xi).

    For any p[1,]p\in[1,\infty] and q[1,)q\in[1,\infty) the Besov space is defined as follows:

    Bp,q0(n)={f𝒮(n):k,1(ϕkf^)Lp(n),fBp,q0<},B_{p,q}^{0}(\mathbb{R}^{n})=\big{\{}f\in\mathcal{S}^{\prime}(\mathbb{R}^{n}):\forall k\in\mathbb{Z},\mathcal{F}^{-1}(\phi_{k}\hat{f})\in L^{p}(\mathbb{R}^{n}),\,\,\,\|f\|_{B_{p,q}^{0}}<\infty\big{\}},

    where

    fBp,q0=1(ϕkf^)q(Lp)=(k=+1(ϕkf^)Lpq)1q.\|f\|_{B_{p,q}^{0}}=\|\mathcal{F}^{-1}(\phi_{k}\hat{f})\|_{\ell^{q}(L^{p})}=\left(\sum_{k=-\infty}^{+\infty}\|\mathcal{F}^{-1}(\phi_{k}\hat{f})\|_{L^{p}}^{q}\right)^{\frac{1}{q}}.

2 Main results

The main results of this paper read as follows.

2.1 Estimates for the linear Cauchy problem

Theorem 2.1.

Let u1Lp(n)u_{1}\in L^{p}(\mathbb{R}^{n}). Then, the solution uu to the Cauchy problem (1.3) satisfies the following estimates:

u(t,)Lqu1Lp{tn4(1p1qβ(p,q)),t1,t1n2(1p1q),t(0,1),\|u(t,\cdot)\|_{L^{q}}\lesssim\|u_{1}\|_{L^{p}}\begin{cases}t^{-\frac{n}{4}\left(\frac{1}{p}-\frac{1}{q}-\beta(p,q)\right)},&t\geq 1,\\ t^{1-\frac{n}{2}\left(\frac{1}{p}-\frac{1}{q}\right)},&t\in(0,1),\end{cases} (2.1)

for all 1pq1\leq p\leq q\leq\infty such that dpl(p,q)<1d_{\text{pl}}(p,q)<1 and for all 1<p2q<1<p\leq 2\leq q<\infty such that dpl(p,q)=1d_{\text{pl}}(p,q)=1, where dpl(p,q)d_{\text{pl}}(p,q) is given by (1.4) and

β(p,q):={[1p+3q2]+, if 1p+1q1,[3p+1q2], if 1p+1q1,\beta(p,q):=\begin{cases}\quad\left[\frac{1}{p}+\frac{3}{q}-2\right]_{+},&\text{ if }\frac{1}{p}+\frac{1}{q}\geq 1,\\ -\,\left[\frac{3}{p}+\frac{1}{q}-2\right]_{-},&\text{ if }\frac{1}{p}+\frac{1}{q}\leq 1,\end{cases}

provided that in the case (p,q)=(1,3)(p,q)=(1,3) or (p,q)=(3/2,)(p,q)=(3/2,\infty) with dpl(p,q)<1d_{\text{pl}}(p,q)<1, the estimate (2.1) is replaced by

u(t,)LqCtn4(1p1q)(logt)12u1Lpfor anyt1.\|u(t,\cdot)\|_{L^{q}}\leq Ct^{-\frac{n}{4}\left(\frac{1}{p}-\frac{1}{q}\right)}(\log t)^{\frac{1}{2}}\|u_{1}\|_{L^{p}}\quad\text{for any}\quad t\geq 1.
1/p1/p01/q1/q11dpl(p,q)=1d_{\text{pl}}(p,q)=11p+3q2=0\frac{1}{p}+\frac{3}{q}-2=01123\frac{2}{3}3p+1q2=0\frac{3}{p}+\frac{1}{q}-2=012n1-\frac{2}{n}dpl(p,q)=1d_{\text{pl}}(p,q)=112\frac{1}{2}2n\frac{2}{n}13\frac{1}{3}tn2+nqt^{-\frac{n}{2}+\frac{n}{q}}tn4(1p1q)t^{-\frac{n}{4}\left(\frac{1}{p}-\frac{1}{q}\right)}tn2npt^{\frac{n}{2}-\frac{n}{p}}
Figure 1: Description of the results in the 1/p1/q1/p-1/q plane with the optimal long-time decay estimates
Remark 2.1.

We notice that the estimates in Theorem 2.1 are consistent with the obtained ones in [15, 22] on the dual line, namely, for p[1,2]p\in[1,2],

u(t,)Lpu1Lp{tn4n2p,t1,t1n2(12p),t(0,1).\|u(t,\cdot)\|_{L^{p^{\prime}}}\lesssim\|u_{1}\|_{L^{p}}\begin{cases}t^{\frac{n}{4}-\frac{n}{2p}},&t\geq 1,\\ t^{1-\frac{n}{2}\left(1-\frac{2}{p^{\prime}}\right)},&t\in(0,1).\end{cases}
Remark 2.2.

Assuming the initial datum u1L1u_{1}\in L^{1}, from Theorem 2.1 we get the following estimates for solutions to (1.3):

u(t,)Lqu1L1{tn4(11q),q>3,t1,tn2+nq(logt)γ,q3,t1,t1n2(11q),t(0,1),{}\|u(t,\cdot)\|_{L^{q}}\lesssim\|u_{1}\|_{L^{1}}\begin{cases}t^{-\frac{n}{4}\left(1-\frac{1}{q}\right)},&q>3,\,t\geq 1,\\ t^{-\frac{n}{2}+\frac{n}{q}}\,(\log t)^{\gamma},&q\leq 3,\,t\geq 1,\\ t^{1-\frac{n}{2}\left(1-\frac{1}{q}\right)}\,,&t\in(0,1),\end{cases} (2.2)

provided that dpl(1,q)=n2q<1d_{\text{pl}}(1,q)=\frac{n}{2q}<1, where γ=γ(p,q)\gamma=\gamma(p,q) is equal to 1/21/2 if (p,q)=(1,3)(p,q)=(1,3) and zero otherwise.

Remark 2.3.

Let us consider u1L1u_{1}\in L^{1}. The condition to avoid the singularity when t0t\to 0 in (2.2) is

1n2(11q)0,1-\frac{n}{2}\left(1-\frac{1}{q}\right)\geq 0,

for q[2,]q\in[2,\infty]. The latter condition is satisfied if, and only if,

1qn2n.\frac{1}{q}\geq\frac{n-2}{n}. (2.3)

Hence, if n2n\leq 2 and q[1,]q\in[1,\infty] or

n3andqnn2,n\geq 3\quad\text{and}\quad q\leq\frac{n}{n-2},

we get non-singular estimates. In particular, this means that we may work without asking additional regularity for n2n\leq 2 (that is the case of αc\alpha_{c}), and for n=3n=3 when q3q\leq 3. However, if (2.3) is not satisfied, we have to ask additional regularity u1L1Lru_{1}\in L^{1}\cap L^{r}, where rr is such that dpl(r,q)1d_{\text{pl}}(r,q)\leq 1 and 1n2(1r1q)01-\frac{n}{2}\left(\frac{1}{r}-\frac{1}{q}\right)\geq 0 for any q[2,].q\in[2,\infty]. Choosing r=2r=2, we get that dpl(2,q)1d_{\text{pl}}(2,q)\leq 1 if and only if 2q2n[n4]+2\leq q\leq\frac{2n}{[n-4]_{+}}, with the exception qq\neq\infty if n=4n=4, and the estimates are non-singular for any q2q\geq 2.

We close this subsection discussing the optimality of the derived estimates in Theorem 2.1. In [22] (see Lemma 2.10), the author proved that

K(t,)MpqCt1n2(1p1q),t(0,1),\|K(t,\cdot)\|_{M_{p}^{q}}\geq Ct^{1-\frac{n}{2}\left(\frac{1}{p}-\frac{1}{q}\right)},\quad t\in(0,1),

and

K(tk,)MpqCtkn4(1p1q),\|K(t_{k},\cdot)\|_{M_{p}^{q}}\geq Ct_{k}^{-\frac{n}{4}\left(\frac{1}{p}-\frac{1}{q}\right)},

for some sequence tk+t_{k}\to+\infty, where KK is defined in (3.2). So, it remains only to check the optimality of

K(t,)Mpqtn2+nq,1p+1q1and1p+3q2,\|K(t,\cdot)\|_{M_{p}^{q}}\leq t^{-\frac{n}{2}+\frac{n}{q}},\quad\frac{1}{p}+\frac{1}{q}\geq 1\,\,\,\,\,\text{and}\,\,\,\,\,\frac{1}{p}+\frac{3}{q}\geq 2, (2.4)

and

K(t,)Mpqtn2np,1p+1q1and3p+1q2.\|K(t,\cdot)\|_{M_{p}^{q}}\leq t^{\frac{n}{2}-\frac{n}{p}},\quad\frac{1}{p}+\frac{1}{q}\leq 1\,\,\,\,\,\text{and}\,\,\,\,\,\frac{3}{p}+\frac{1}{q}\geq 2. (2.5)

We are going to study only (2.4), because (2.5) follows from (2.4) by duality. To prove that (2.4) is optimal we need the following result, which was inspired by [23], Lemma 8.

Lemma 2.1.

Let 1p<1\leq p<\infty. There exist fLpf\in L^{p}, b>a>0b>a>0 and C>0C>0 such that there exists tkt_{k}\to\infty satisfying

|K(tk,x)f(x)|Ctkn2,|K(t_{k},x)\ast f(x)|\geq C\,t^{-\frac{n}{2}}_{k}, (2.6)

for all xx such that atk<|x|<btkat_{k}<|x|<bt_{k}.

The proof of Lemma 2.1 is quite technical and long, for this reason, we decide to postpone it to an appendix at the end of the paper (see Appendix B).

Now, take fLpf\in L^{p}, 0<a<b0<a<b and tkt_{k} as in Lemma 2.1. Assume in addition that fLp=1\|f\|_{L^{p}}=1. Then we have

K(tk,x)f(x)Lq(atk<|x|<btk|K(tk,x)f(x)|q𝑑x)1qtkn2+nq,\displaystyle\|K(t_{k},x)\ast f(x)\|_{L^{q}}\geq\left(\int_{at_{k}<|x|<bt_{k}}|K(t_{k},x)\ast f(x)|^{q}\,dx\right)^{\frac{1}{q}}\gtrsim t^{-\frac{n}{2}+\frac{n}{q}}_{k},

and the optimality of (2.4) is concluded.

2.2 Existence results to the semilinear Cauchy problem

Based on the Duhamel’s principle and the estimates of the linear Cauchy problem from Theorem 2.1, we have the following local existence result.

Theorem 2.2.

(Local Existence) Let n1n\geq 1, (u0,u1)H2(n)×L2(n)(u_{0},u_{1})\in H^{2}(\mathbb{R}^{n})\times L^{2}(\mathbb{R}^{n}), and suppose that 1<α<n+4[n4]+1<\alpha<\frac{n+4}{[n-4]_{+}}. Then, there exists T>0T>0 and a local unique Sobolev solution u𝒞([0,T],Lq(n))u\in\mathcal{C}([0,T],L^{q}(\mathbb{R}^{n})) to (1.5) for all 2q2n[n4]+2\leq q\leq\frac{2n}{[n-4]_{+}}, excluding the case q=q=\infty when n=4n=4.

Remark 2.4.

We point out that Theorem 2.2 was already obtained in [22] (see Theorem 4.14.1). In fact, in [16] (see Theorem 1.1), applying Strichartz type estimates, the authors proved the local existence of energy solutions 𝒞([0,T),H2(n))𝒞1([0,T),L2(n))\mathcal{C}([0,T),H^{2}(\mathbb{R}^{n}))\cap\mathcal{C}^{1}([0,T),L^{2}(\mathbb{R}^{n})) for 1<αn+4[n4]+1<\alpha\leq\frac{n+4}{[n-4]_{+}} to (1.5), i.e., they proved that a local existence result is still valid in the endpoint α=n+4[n4]+\alpha=\frac{n+4}{[n-4]_{+}}. In this paper we give an alternative (short) proof using the derived estimates to the linear Cauchy problem, but obtaining solution in the space 𝒞([0,T),Lq(n))\mathcal{C}([0,T),L^{q}(\mathbb{R}^{n})). We also remark that since H2LqH^{2}\subset L^{q} for 2q2n[n4]+2\leq q\leq\frac{2n}{[n-4]_{+}}, using Theorem 4.14.1 of [22], we conclude that the solution obtained in Theorem 2.2 must belong to 𝒞([0,T),H2(n))\mathcal{C}([0,T),H^{2}(\mathbb{R}^{n})).

Under the assumption of small energy data, one may conclude a priori uniform bounds for the energy solutions to (1.5) (the same reasoning holds true for the semilinear Klein-Gordon scalar equation, see [20]). Hence, the local solution given by Theorem 2.2 and Remark 2.4 can be prolonged in order the get the following result.

Theorem 2.3.

(see, for instance, Theorem 4.1 in [16]) Let us assume that 1<αn+4[n4]+1<\alpha\leq\frac{n+4}{[n-4]_{+}}. Then, there exists a small constant ε>0\varepsilon>0 such that for any data (u0,u1)H2(n)×L2(n)(u_{0},u_{1})\in H^{2}(\mathbb{R}^{n})\times L^{2}(\mathbb{R}^{n}) satisfying the assumption u0H2+u1L2<ε,\|u_{0}\|_{H^{2}}+\|u_{1}\|_{L^{2}}<\varepsilon, there is a uniquely determined global (in time) energy solution 𝒞([0,),H2(n))𝒞1([0,),L2(n))\mathcal{C}([0,\infty),H^{2}(\mathbb{R}^{n}))\cap\mathcal{C}^{1}([0,\infty),L^{2}(\mathbb{R}^{n})) to (1.5).

Next, inspired by [18], we give a (counterpart of Theorem 2.2) non-existence result for local weak solution to (1.5) for large values of α\alpha. Before to state the result we need to recall the definition of weak solution.

Definition 2.1.

We say that uLlocα([0,T)×n)u\in L^{\alpha}_{\text{loc}}([0,T)\times\mathbb{R}^{n}), T>0T>0, is a weak solution to (1.5) with u00u_{0}\equiv 0 when

0Tnu(t,x){t2+Δx2+1}ψ(t,x)𝑑x𝑑t=0Tn|u(t,x)|αψ(t,x)𝑑x𝑑t+nu1(x)ψ(0,x)𝑑x𝑑t\int_{0}^{T}\int_{\mathbb{R}^{n}}u(t,x)\{\partial^{2}_{t}+\Delta^{2}_{x}+1\}\psi(t,x)dxdt=\int_{0}^{T}\int_{\mathbb{R}^{n}}|u(t,x)|^{\alpha}\psi(t,x)dxdt+\int_{\mathbb{R}^{n}}u_{1}(x)\psi(0,x)dxdt

for every ψ𝒞c([0,T)×n)\psi\in\mathcal{C}^{\infty}_{c}([0,T)\times\mathbb{R}^{n}).

Theorem 2.4.

Let m[1,2]m\in[1,2] and suppose n>2mn>2m. If α>n+2mn2m\alpha>\frac{n+2m}{n-2m}, then there exists 0fLm(n)0\neq f\in L^{m}(\mathbb{R}^{n}) such that if there exists a local weak solution to (1.5) with u00u_{0}\equiv 0 and u1=λfu_{1}=\lambda f, for some λ0\lambda\geq 0, then λ=0\lambda=0.

Remark 2.5.

Thanks to Theorems 2.2-2.3 and Theorem 2.4 with m=2m=2, we conclude that for n5n\geq 5, α=n+4n4\alpha=\frac{n+4}{n-4} is a critical exponent for the local (and global) existence of weak solution to (1.5).

Remark 2.6.

If we assume u1L1(n)L1+ϵ(n)u_{1}\in L^{1}(\mathbb{R}^{n})-L^{1+\epsilon}(\mathbb{R}^{n}) for any ϵ>0\epsilon>0 and use the estimates to the linear Cauchy problem in Theorem 2.1, one can still get a local existence result for all n2<α<n[n2]+\frac{n}{2}<\alpha<\frac{n}{[n-2]_{+}} to (1.5) when the space dimension n3n\leq 3 but not for n4n\geq 4. Indeed, in order to get a local unique solution u𝒞([0,T),Lq(n))u\in\mathcal{C}([0,T),L^{q}(\mathbb{R}^{n})) we have to assume two compatibility conditions, namely, 1n2(11q)01-\frac{n}{2}\left(1-\frac{1}{q}\right)\geq 0 in (2.2) and dpl(1,q)=n2q<1d_{\text{pl}}(1,q)=\frac{n}{2q}<1, that is,

n2<qn[n2]+.\frac{n}{2}<q\leq\frac{n}{[n-2]_{+}}.

For this reason, unlike in Theorem 2.5, in the next Theorem 2.6 we have to assume u1L1(n)Lr(n)u_{1}\in L^{1}(\mathbb{R}^{n})\cap L^{r}(\mathbb{R}^{n}), for some r(1,2]r\in(1,2].

In the following results we apply the obtained estimates to the linear Cauchy problem to derive the global (in time) existence of solution to (1.5) under the assumption of small datum u1L1(n)u_{1}\in L^{1}(\mathbb{R}^{n}). Since the L1LqL^{1}-L^{q} estimates change according whether qq is smaller or larger than 33, we have to work with the two following candidates to critical exponents

αc=1+4nandα~c=2+2n.\alpha_{c}=1+\frac{4}{n}\qquad\text{and}\qquad\tilde{\alpha}_{c}=2+\frac{2}{n}. (2.7)

We point out that αcα~c\alpha_{c}\leq\tilde{\alpha}_{c} if, and only if, n2n\leq 2. Here we use the expression critical exponents in the sense that in the supercritical case, we prove the existence of a global in time solution that satisfies the same optimal decay estimate as the solution of the associated linear Cauchy problem. However, it is not clear what happens in the subcritical case assuming only u1L1(n)L1+ϵ(n)u_{1}\in L^{1}(\mathbb{R}^{n})-L^{1+\epsilon}(\mathbb{R}^{n}) for any ϵ>0\epsilon>0.

Theorem 2.5.

Let n=1,2n=1,2 and assume u00u_{0}\equiv 0 and α>αc\alpha>\alpha_{c}, where αc\alpha_{c} is given in (2.7). Then, there exists a small constant ε>0\varepsilon>0 such that for any datum u1L1(n)u_{1}\in L^{1}(\mathbb{R}^{n}) satisfying the assumption u1L1<ε,\|u_{1}\|_{L^{1}}<\varepsilon, there is a uniquely determined global (in time) Sobolev solution

u𝒞([0,),L2(n)L(n))\displaystyle u\in\mathcal{C}([0,\infty),L^{2}(\mathbb{R}^{n})\cap L^{\infty}(\mathbb{R}^{n}))

to (1.5). Moreover, the solution satisfies the following decay estimates:

u(t,)LqC(1+t)n4(11q)+n4β(1,q)(log(1+t))γu1L1,\displaystyle\|u(t,\cdot)\|_{L^{q}}\leq C(1+t)^{-\frac{n}{4}\left(1-\frac{1}{q}\right)+\frac{n}{4}\beta(1,q)}(\log(1+t))^{\gamma}\,\|u_{1}\|_{L^{1}},

for all q[2,]q\in[2,\infty], where γ\gamma is defined in (2.2), and the constant C>0C>0 does not depend on the initial datum.

Due to Remark 2.6, in the next result we have to ask the additional hypothesis u1Lr(n)u_{1}\in L^{r}(\mathbb{R}^{n}), for some r(1,2]r\in(1,2]. In order to avoid further restrictions on the space of solution and on the upper bound for α\alpha, we shall assume u1L1(n)L2(n)u_{1}\in L^{1}(\mathbb{R}^{n})\cap L^{2}(\mathbb{R}^{n}).

Theorem 2.6.

Let n=3,4n=3,4 and assume u00u_{0}\equiv 0 and α>α~c\alpha>\tilde{\alpha}_{c}, where α~c\tilde{\alpha}_{c} is defined in (2.7). Then, there exists a small constant ε>0\varepsilon>0 such that for any u1𝒜:=L1(n)L2(n)u_{1}\in\mathcal{A}:=L^{1}(\mathbb{R}^{n})\cap L^{2}(\mathbb{R}^{n}) satisfying the assumption

u1𝒜:=u1L1+u1L2<ε,\|u_{1}\|_{\mathcal{A}}:=\|u_{1}\|_{L^{1}}+\|u_{1}\|_{L^{2}}<\varepsilon,

there exists a uniquely determined global (in time) Sobolev solution

u𝒞([0,),L2(n)Lp~(n)),\displaystyle u\in\mathcal{C}([0,\infty),L^{2}(\mathbb{R}^{n})\cap L^{\tilde{p}}(\mathbb{R}^{n})),

for any p~[2,)\tilde{p}\in[2,\infty) to (1.5). Moreover, the following decay estimate holds:

u(t,)LqC(1+t)n4(11q)+n4β(1,q)(log(1+t))γu1𝒜,\displaystyle\|u(t,\cdot)\|_{L^{q}}\leq C(1+t)^{-\frac{n}{4}\left(1-\frac{1}{q}\right)+\frac{n}{4}\beta(1,q)}(\log(1+t))^{\gamma}\,\|u_{1}\|_{\mathcal{A}},

for all q[2,p~]q\in[2,\tilde{p}], where γ\gamma is defined as in (2.2), and the constant C>0C>0 does not depend on the initial datum.

Remark 2.7.

When n=3n=3 we can replace p~\tilde{p} by \infty in Theorem 2.6. Moreover, if u1L1(n)L1+ϵ(n)u_{1}\in L^{1}(\mathbb{R}^{n})-L^{1+\epsilon}(\mathbb{R}^{n}) for any ϵ>0\epsilon>0, one can still get a global (in time) small data Sobolev solution in 𝒞([0,),L2(3)L3(3))\mathcal{C}([0,\infty),L^{2}(\mathbb{R}^{3})\cap L^{3}(\mathbb{R}^{3})) to (1.5) for α~c(3)=83<α3\tilde{\alpha}_{c}(3)=\frac{8}{3}<\alpha\leq 3.

Remark 2.8.

We get the restriction n4n\leq 4 in Theorem 2.6 because dpl(1,α~c)>1d_{\text{pl}}(1,\tilde{\alpha}_{c})>1 for n5n\geq 5. We also point out that dpl(1,2)=1d_{\text{pl}}(1,2)=1 when n=4n=4, so in this situation we need to use LrL2L^{r}-L^{2} estimates for some r(1,2]r\in(1,2].

3 Proof of Theorem 2.1

Let us apply partial Fourier transform with respect to the spatial variable xx to the linear problem (1.3). Then, u^(t,ξ)=xξ(u(t,x))\widehat{u}(t,\xi)=\mathcal{F}_{x\rightarrow\xi}(u(t,x)) solves

{u^tt+(1+|ξ|4)u^=0,(t,ξ)(0,)×n,u^(0,ξ)=0,u^t(0,ξ)=u^1(ξ),ξn.\begin{cases}\widehat{u}_{tt}+(1+|\xi|^{4})\widehat{u}=0,&\,\,\,(t,\xi)\in(0,\infty)\times\mathbb{R}^{n},\\ \widehat{u}(0,\xi)=0,\quad\widehat{u}_{t}(0,\xi)=\widehat{u}_{1}(\xi),&\,\,\,\xi\in\mathbb{R}^{n}.\end{cases} (3.1)

The characteristic roots λ±(|ξ|)\lambda_{\pm}(|\xi|) to the equation in (3.1) are given by

λ±(|ξ|)=±i1+|ξ|4.\lambda_{\pm}(|\xi|)=\pm i\sqrt{1+|\xi|^{4}}\,.

The solution to (3.1) is then given by

u^(t,ξ)=eλ+(|ξ|)teλ(|ξ|)tλ+(|ξ|)λ(|ξ|)=:K^(t,|ξ|)u^1(ξ).\widehat{u}(t,\xi)=\underbrace{\frac{e^{\lambda_{+}(|\xi|)t}-e^{\lambda_{-}(|\xi|)t}}{\lambda_{+}(|\xi|)-\lambda_{-}(|\xi|)}}_{{=:\widehat{K}(t,|\xi|)}}\widehat{u}_{1}(\xi).

Our assignment in this part is to process estimates for the Fourier multiplier K^(t,|ξ|)\widehat{K}(t,|\xi|). Precisely, the kernel can be written by

K^(t,|ξ|)=sin(1+|ξ|4)t1+|ξ|4.\widehat{K}(t,|\xi|)=\frac{\sin\big{(}\sqrt{1+|\xi|^{4}}\big{)}t}{\sqrt{1+|\xi|^{4}}}. (3.2)

Then, the solution to (1.3) can be expressed by

u(t,x)=ξx1(sin(1+|ξ|4)t1+|ξ|4)(x)u1(x).u(t,x)=\mathcal{F}_{\xi\rightarrow x}^{-1}\bigg{(}\frac{\sin\big{(}\sqrt{1+|\xi|^{4}}\big{)}t}{\sqrt{1+|\xi|^{4}}}\bigg{)}\ast_{(x)}u_{1}(x).

Let χ𝒞c([0,))\chi\in\mathcal{C}_{c}^{\infty}([0,\infty)) with χ(ρ)=1\chi(\rho)=1 for ρ[0,1]\rho\in[0,1] and χ(ρ)=0\chi(\rho)=0 for ρ2\rho\geq 2. Then, we divide our considerations in two cases, small and large frequencies as follows:

K^0(t,)=χ(2|ξ|)K^(t,),K^(t,)=(1χ(|ξ|))K^(t,).\widehat{K}_{0}(t,\cdot)=\chi(2|\xi|)\widehat{K}(t,\cdot),\qquad\ \widehat{K}_{\infty}(t,\cdot)=(1-\chi(|\xi|))\widehat{K}(t,\cdot).

3.1 Estimates for high frequencies

Since at high frequencies one may expect the same behaviour of the free plate equation, we apply the change of variables by ξ=t12η\xi=t^{-\frac{1}{2}}\eta in order to get

K^(t,)Mpq=t1n2(1p1q)m(t,)Mpq,\|\widehat{K}_{\infty}(t,\cdot)\|_{M_{p}^{q}}=t^{1-\frac{n}{2}\left(\frac{1}{p}-\frac{1}{q}\right)}\|m_{\infty}(t,\cdot)\|_{M_{p}^{q}},

where mm_{\infty} is the radial multiplier

m(t,η)=sin(|η|21+(t/|η|2)2)|η|21+(t/|η|2)2φ(t,η)=sin(ω(t,|η|))ω(t,|η|)φ(t,η)m_{\infty}(t,\eta)=\frac{\sin\big{(}|\eta|^{2}\sqrt{1+(t/|\eta|^{2})^{2}}\big{)}}{|\eta|^{2}\sqrt{1+(t/|\eta|^{2})^{2}}}\varphi_{\infty}(t,\eta)=\frac{\sin(\omega_{\infty}(t,|\eta|))}{\omega_{\infty}(t,|\eta|)}\varphi_{\infty}(t,\eta)

with

ω(t,|η|):=|η|21+(t/|η|2)2,\omega_{\infty}(t,|\eta|):=|\eta|^{2}\sqrt{1+(t/|\eta|^{2})^{2}},

and

φ(t,η)=(1χ)(t12|η|).\varphi_{\infty}(t,\eta)=(1-\chi)(t^{-\frac{1}{2}}|\eta|).

Now, we divide again our considerations in two cases, new small and large frequencies as follows:

m,0(t,η)\displaystyle m_{\infty,0}(t,\eta) =χ(|η|)m(t,η),\displaystyle=\chi(|\eta|)m_{\infty}(t,\eta),
m,1(t,η)\displaystyle m_{\infty,1}(t,\eta) =(1χ(|η|))m(t,η).\displaystyle=\big{(}1-\chi(|\eta|)\big{)}m_{\infty}(t,\eta).

We notice that this division is effective only for t(0,1)t\in(0,1), since m,1(t,η)=m(t,η)m_{\infty,1}(t,\eta)=m_{\infty}(t,\eta) for t>1t>1.

Proposition 3.1.

On the support of φ(t,ρ)\varphi_{\infty}(t,\rho) the following estimates hold:

|ρω(t,ρ)|\displaystyle|\partial_{\rho}^{\ell}\omega_{\infty}(t,\rho)| Cρ2,,\displaystyle\leq C_{\ell}\rho^{2-\ell},\qquad\ell\in\mathbb{N}, (3.3)
|ρω(t,ρ)|\displaystyle|\partial_{\rho}^{\ell}\omega_{\infty}(t,\rho)| cρ2,=1,2,\displaystyle\geq c_{\ell}\rho^{2-\ell},\,\,\quad\,\,\,\,\,\,\ell=1,2, (3.4)
|ρ1w(t,ρ)|\displaystyle\big{|}\partial_{\rho}^{\ell}\frac{1}{w_{\infty}(t,\rho)}\big{|} C~ρ2,.\displaystyle\leq\widetilde{C}_{\ell}\rho^{-2-\ell},\quad\,\,\,\,\ell\in\mathbb{N}. (3.5)
Proof.

Observing that

ω(t,|η|)=|η|4+t2,\omega_{\infty}(t,|\eta|)=\sqrt{|\eta|^{4}+t^{2}},

and t2ρ4t^{2}\leq\rho^{4}, we get

ρω(t,ρ)\displaystyle\partial_{\rho}\omega_{\infty}(t,\rho) =2ρ3ρ4+t22ρ,\displaystyle=\frac{2\rho^{3}}{\sqrt{\rho^{4}+t^{2}}}\geq\sqrt{2}\rho,
ρ2ω(t,ρ)\displaystyle\partial_{\rho}^{2}\omega_{\infty}(t,\rho) =ρ(2ρ3ρ4+t2)=2ρ2ρ4+t22(32ρ4ρ4+t2)1c.\displaystyle=\partial_{\rho}\Big{(}\frac{2\rho^{3}}{\sqrt{\rho^{4}+t^{2}}}\Big{)}=\underbrace{\frac{2\rho^{2}}{\sqrt{\rho^{4}+t^{2}}}}_{\geq\sqrt{2}}\underbrace{\left(3-\frac{2\rho^{4}}{\rho^{4}+t^{2}}\right)}_{\geq 1}\geq c.

In the same manner one can prove (3.3) and (3.5). ∎

Lemma 3.1.

For any 1pq1\leq p\leq q\leq\infty and t[0,1]t\in[0,1], we have

m,0(t,)MpqC.\displaystyle\|m_{\infty,0}(t,\cdot)\|_{M_{p}^{q}}\leq C.
Proof.

Since χ(|η|)𝒞c\chi(|\eta|)\in\mathcal{C}_{c}^{\infty} is independent of tt, we get (see Theorem 1.81.8 in [17])

m,0(t,)MpqCmin{m,0(t,)Mpp,m,0(t,)Mqq}.\|m_{\infty,0}(t,\cdot)\|_{M_{p}^{q}}\leq C\min\big{\{}\|m_{\infty,0}(t,\cdot)\|_{M_{p}^{p}},\|m_{\infty,0}(t,\cdot)\|_{M_{q}^{q}}\big{\}}.

Let (p,q){(1,1),(,),(1,)}(p,q)\notin\{(1,1),(\infty,\infty),(1,\infty)\}. We know that |sinρ|Cρ(1+ρ)1|\sin\rho|\leq C\rho(1+\rho)^{-1}, that is, |sinρ|ρ|\sin\rho|\leq\rho for small ρ\rho and |sinρ|1|\sin\rho|\leq 1 for large ρ\rho. Moreover, it holds

|ρksincρ|ρk,|\partial_{\rho}^{k}\operatorname{sinc\,}\rho|\lesssim\rho^{-k},

for ρ1\rho\leq 1, where sincρ:=ρ1sinρ\operatorname{sinc\,}\rho:=\rho^{-1}sin\ \rho. Therefore, using the estimates (3.3), (3.5) from Proposition 3.1 and chain rule we may conclude that

|ρksincω(t,ρ)|ρk.|\partial_{\rho}^{k}\operatorname{sinc\,}\omega_{\infty}(t,\rho)|\lesssim\rho^{-k}.

Thus, by Mikhlin-Hörmander theorem (cf. Theorem A.2) we find that the multiplier norm is bounded by a uniform constant with respect to tt. Now we consider (p,q)=(1,)(p,q)=(1,\infty). We have

m,0(t,)M1\displaystyle\|m_{\infty,0}(t,\cdot)\|_{M_{1}^{\infty}} =ηx1(χ(|η|)sincω(t,|η|)φ(t,η))L\displaystyle=\big{\|}\mathcal{F}_{\eta\to x}^{-1}\big{(}\chi(|\eta|)\operatorname{sinc\,}\omega_{\infty}(t,|\eta|)\varphi_{\infty}(t,\eta)\big{)}\big{\|}_{L^{\infty}}
neixηχ(|η|)sincω(t,|η|)φ(t,η)𝑑ηL\displaystyle\lesssim\big{\|}\int_{\mathbb{R}^{n}}e^{ix\cdot\eta}\chi(|\eta|)\operatorname{sinc\,}\omega_{\infty}(t,|\eta|)\varphi_{\infty}(t,\eta)d\eta\big{\|}_{L^{\infty}}
|η|1|sincω(t,|η|)|𝑑η1.\displaystyle\lesssim\int_{|\eta|\leq 1}|\operatorname{sinc\,}\omega_{\infty}(t,|\eta|)|d\eta\lesssim 1.

Finally, we consider the case (p,q)=(1,1)(p,q)=(1,1) (and so by duality also (p,q)=(,)(p,q)=(\infty,\infty)). In particular, thanks to (L1)M11\mathscr{F}(L^{1})\subset M_{1}^{1}, it is enough to prove 1m,0L1\mathscr{F}^{-1}m_{\infty,0}\in L^{1}. Since m,0𝒞c([0,1]×n)m_{\infty,0}\in\mathcal{C}_{c}([0,1]\times\mathbb{R}^{n}), we have that

|ρkm,0|Ck,k.|\partial_{\rho}^{k}m_{\infty,0}|\leq C_{k},\quad k\in\mathbb{N}.

Then, for any integer N0N\geq 0, by NN integration by parts

|1m,0(x)|=|(2π)n|ξ|1eixξm(ξ)𝑑ξ||x|N|ξ|1𝑑ξ|x|N.\displaystyle\left|\mathscr{F}^{-1}m_{\infty,0}\,(x)\right|=\left|(2\pi)^{-n}\int_{|\xi|\leq 1}e^{ix\cdot\xi}\,m(\xi)\,d\xi\right|\lesssim|x|^{-N}\int_{|\xi|\leq 1}\,d\xi\lesssim|x|^{-N}.

Choosing N=0N=0 for |x|<1|x|<1 and N=n+1N=n+1 for |x|1|x|\geq 1, we can conclude the proof. ∎

Proposition 3.2.

The determinant of the Hessian matrix satisfies the following estimate:

|detHω(t,|η|)|c.|\det H_{\omega_{\infty}(t,|\eta|)}|\geq c.
Proof.

We follow the proof of Proposition 5.15.1 in [10]. We have

ηkηj(ω(t,|η|))=(ρ1ρωδjk+ρ1ρ(ρ1ρω)ηjηk)ρ=|η|.\partial_{\eta_{k}}\partial_{\eta_{j}}(\omega_{\infty}(t,|\eta|))=\big{(}\rho^{-1}\partial_{\rho}\omega_{\infty}\delta_{j}^{k}+\rho^{-1}\partial_{\rho}(\rho^{-1}\partial_{\rho}\omega_{\infty})\eta_{j}\eta_{k}\big{)}_{\rho=|\eta|}.

Then, the Hessian matrix is

Hω=αIn+βηη/|η|2withα=|η|1ρω,β=ρ2ωα,H_{\omega_{\infty}}=\alpha I_{n}+\beta\eta\otimes\eta/|\eta|^{2}\quad\text{with}\quad\alpha=|\eta|^{-1}\partial_{\rho}\omega_{\infty}\,,\,\,\,\beta=\partial_{\rho}^{2}\omega_{\infty}-\alpha,

where InI_{n} is the identity matrix and ηη\eta\otimes\eta is the matrix with entries (ηkηj)j,k(\eta_{k}\eta_{j})_{j,k}. Finally, we arrive at

detHω=(1+βα1)αn=(ρ2ω)(|η|1ρω)n1c,\det H_{\omega_{\infty}}=(1+\beta\alpha^{-1})\alpha^{n}=(\partial_{\rho}^{2}\omega_{\infty})(|\eta|^{-1}\partial_{\rho}\omega_{\infty})^{n-1}\geq c,

due to estimates in (3.4). ∎

Lemma 3.2.

We have the following estimates:

m,1(t,)MpqC for 1pq such that dpl(p,q)<1,\|m_{\infty,1}(t,\cdot)\|_{M_{p}^{q}}\leq C\qquad\text{ for }1\leq p\leq q\leq\infty\text{ such that }d_{\text{pl}}(p,q)<1,

for any 0<t<10<t<1, and

m,1(t,)MpqCtdpl1 for 1pq such that dpl(p,q)<1,\|m_{\infty,1}(t,\cdot)\|_{M_{p}^{q}}\leq Ct^{d_{\text{pl}}-1}\qquad\text{ for }1\leq p\leq q\leq\infty\text{ such that }d_{\text{pl}}(p,q)<1,

for any t1t\geq 1. Moreover, the estimates remain true for 1<p2q<1<p\leq 2\leq q<\infty such that dpl(p,q)=1d_{\text{pl}}(p,q)=1.

Proof.

Since Mpq=Mqp\|\cdot\|_{M_{p}^{q}}=\|\cdot\|_{M_{q^{\prime}}^{p^{\prime}}}, it is enough to prove Lemma 3.2 for 1p21\leq p\leq 2 and pqpp\leq q\leq p^{\prime}.

We are going to use Littman’s lemma (cf. Lemma A.1) to estimate the norm

m,1(t,)ϕk()M1=ξx1[(eiω(t,|ξ|)eiω(t,|ξ|))ϕk(ξ)f(t,|ξ|)]L,\|m_{\infty,1}(t,\cdot)\phi_{k}(\cdot)\|_{M_{1}^{\infty}}=\big{\|}\mathcal{F}_{\xi\to x}^{-1}\big{[}\big{(}e^{i\omega_{\infty}(t,|\xi|)}-e^{-i\omega_{\infty}(t,|\xi|)}\big{)}\phi_{k}(\xi)f(t,|\xi|)\big{]}\big{\|}_{L^{\infty}},

where

f(t,|ξ|)=(1χ)(|ξ|)φ(t,|ξ|)ω(t,|ξ|).f(t,|\xi|)=\frac{(1-\chi)(|\xi|)\varphi_{\infty}(t,|\xi|)}{\omega_{\infty}(t,|\xi|)}.

Let η:=2kξ\eta:=2^{-k}\xi and dη=2kndξd\eta=2^{-kn}d\xi. Then, we set w(t,2kη):=22kw~(t,η)w_{\infty}(t,2^{k}\eta):=2^{2k}\widetilde{w}_{\infty}(t,\eta), with w~(t,η):=w(22kt,η)=t224k+|η|4\widetilde{w}_{\infty}(t,\eta):=w_{\infty}(2^{-2k}t,\eta)=\sqrt{t^{2}2^{-4k}+|\eta|^{4}}. The functions w~\widetilde{w}_{\infty} and (w~)1(\widetilde{w}_{\infty})^{-1} are uniformly bounded with respect to tt and kk, together with their derivatives on the support of ϕ\phi. Hence,

ξx1[(eiω(t,|ξ|)\displaystyle\big{\|}\mathcal{F}_{\xi\to x}^{-1}\big{[}\big{(}e^{i\omega_{\infty}(t,|\xi|)} eiω(t,|ξ|))ϕk(ξ)1w(t,ξ)]L\displaystyle-e^{-i\omega_{\infty}(t,|\xi|)}\big{)}\phi_{k}(\xi)\frac{1}{w_{\infty}(t,\xi)}\big{]}\big{\|}_{L^{\infty}}
=2k(n2)ηx1[ei22kw~(t,η)ϕ(|η|)1w~(t,η)]L\displaystyle=2^{k(n-2)}\big{\|}\mathcal{F}_{\eta\to x}^{-1}\big{[}e^{i2^{2k}\widetilde{w}_{\infty}(t,\eta)}\phi(|\eta|)\frac{1}{\widetilde{w}_{\infty}(t,\eta)}\big{]}\big{\|}_{L^{\infty}}
C2k(n2)(1+22k)n222k.\displaystyle\leq C2^{k(n-2)}(1+2^{2k})^{-\frac{n}{2}}\approx 2^{-2k}.

Thus, Young’s inequality implies

ξx1(m,1(t,ξ)ϕk(ξ)(u1))L\displaystyle\big{\|}\mathcal{F}_{\xi\to x}^{-1}\big{(}m_{\infty,1}(t,\xi)\phi_{k}(\xi)\mathcal{F}(u_{1})\big{)}\big{\|}_{L^{\infty}} =ξx1(m,1(t,ξ)ϕk(ξ))u1L\displaystyle=\big{\|}\mathcal{F}_{\xi\to x}^{-1}\big{(}m_{\infty,1}(t,\xi)\phi_{k}(\xi)\big{)}\ast u_{1}\big{\|}_{L^{\infty}}
ξx1(m,1(t,ξ)ϕk(ξ))Lu1L1\displaystyle\leq\big{\|}\mathcal{F}_{\xi\to x}^{-1}\big{(}m_{\infty,1}(t,\xi)\phi_{k}(\xi)\big{)}\big{\|}_{L^{\infty}}\|u_{1}\|_{L^{1}}
C22ku1L1,\displaystyle\leq C2^{-2k}\|u_{1}\|_{L^{1}},

which allow us to conlcude that

m,1(ξ)ϕk(|ξ|)M1C22k.\|m_{\infty,1}(\xi)\phi_{k}(|\xi|)\|_{M_{1}^{\infty}}\leq C2^{-2k}. (3.6)

Now using the fact that M22=LM_{2}^{2}=L^{\infty}, we obtain

m,1(t,)ϕkM22=m,1(t,)ϕkLC22k.\displaystyle\|m_{\infty,1}(t,\cdot)\phi_{k}\|_{M_{2}^{2}}=\|m_{\infty,1}(t,\cdot)\phi_{k}\|_{L^{\infty}}\leq C2^{-2k}. (3.7)

In this way, from (3.6), (3.7) and Riesz-Thorin interpolation theorem (cf. Theorem A.1) (L2L2L^{2}-L^{2} and L1LL^{1}-L^{\infty}) we get

m,1(t,)ϕkMp0q0C22kwith1p0+1q0=1.\|m_{\infty,1}(t,\cdot)\phi_{k}\|_{M_{p_{0}}^{q_{0}}}\leq C2^{-2k}\qquad\text{with}\qquad\frac{1}{p_{0}}+\frac{1}{q_{0}}=1.

Thanks to (3.3), we have

ξβ(m,1(t,)ϕk)L2Cβ(2k1|ξ|2k+1|ξ|4+2|β|𝑑ξ)12Cn,β2k(|β|2+n2),\|\partial_{\xi}^{\beta}(m_{\infty,1}(t,\cdot)\phi_{k})\|_{L^{2}}\leq C_{\beta}\Big{(}\int_{2^{k-1}\leq|\xi|\leq 2^{k+1}}|\xi|^{-4+2|\beta|}d\xi\Big{)}^{\frac{1}{2}}\leq C_{n,\beta}2^{k\left(|\beta|-2+\frac{n}{2}\right)},

so, from Bernstein’s Theorem (cf. Theorem A.3) we obtain the estimate

m,1(t,)ϕkM11m,1(t,)ϕkL21n2N|β|=Nξβ(m,1(t,)ϕk)L2n2NC2k(n2),\|m_{\infty,1}(t,\cdot)\phi_{k}\|_{M_{1}^{1}}\leq\|m_{\infty,1}(t,\cdot)\phi_{k}\|_{L^{2}}^{1-\frac{n}{2N}}\sum_{|\beta|=N}\|\partial_{\xi}^{\beta}\big{(}m_{\infty,1}(t,\cdot)\phi_{k}\big{)}\|_{L^{2}}^{\frac{n}{2N}}\leq C2^{k(n-2)}, (3.8)

where we choose N>n/2N>n/2.

Using (3.6), (3.8) and again Riesz-Thorin interpolation theorem (Lp0Lq0L^{p_{0}}-L^{q_{0}} and L1L1L^{1}-L^{1}), we conclude

m,1(t,)ϕkMpqC2k(nθ2)=C2k(n(1p+1q1)2),\|m_{\infty,1}(t,\cdot)\phi_{k}\|_{M_{p}^{q}}\leq C2^{k(n\theta-2)}=C2^{k\left(n\left(\frac{1}{p}+\frac{1}{q}-1\right)-2\right)},

where 0<θ<10<\theta<1, 1p=1θp0+θ\frac{1}{p}=\frac{1-\theta}{p_{0}}+\theta and 1q=1θq0+θ\frac{1}{q}=\frac{1-\theta}{q_{0}}+\theta.

Now, we have that

m,1(t,)Mpqk=k1(t)m,1(t,)ϕkMpq,\displaystyle\|m_{\infty,1}(t,\cdot)\|_{M_{p}^{q}}\leq\sum_{k=k_{1}(t)}^{\infty}\|m_{\infty,1}(t,\cdot)\phi_{k}\|_{M_{p}^{q}},

where k1(t)=min{k: 2k+1t12}.k_{1}(t)=\min{\{k\in\mathbb{N}\,:\,2^{k+1}\geq t^{\frac{1}{2}}\}}. We notice that if 0<t<10<t<1, then k1(t)=0k_{1}(t)=0 and

m,1(t,)Mpqk=0m,1(t,)ϕkMpqCk=02k(n(1p+1q1)2),\displaystyle\|m_{\infty,1}(t,\cdot)\|_{M_{p}^{q}}\leq\sum_{k=0}^{\infty}\|m_{\infty,1}(t,\cdot)\phi_{k}\|_{M_{p}^{q}}\leq C\sum_{k=0}^{\infty}2^{k\left(n\left(\frac{1}{p}+\frac{1}{q}-1\right)-2\right)},

which converges if and only if

n(1p+1q1)2<0,n\left(\frac{1}{p}+\frac{1}{q}-1\right)-2<0,

or equivalently dpl<1.d_{\text{pl}}<1. Otherwise, if t1t\geq 1, we have

m,1(t,)Mpq\displaystyle\|m_{\infty,1}(t,\cdot)\|_{M_{p}^{q}}\leq k=k1(t)m,1(t,)ϕkMpq\displaystyle\sum_{k=k_{1}(t)}^{\infty}\|m_{\infty,1}(t,\cdot)\phi_{k}\|_{M_{p}^{q}}
C 22(dpl1)k1(t)Ctdpl1,\displaystyle\leq C\,2^{2(d_{\text{pl}}-1)k_{1}(t)}\leq Ct^{d_{\text{pl}}-1},

where we used that dpl<1d_{\text{pl}}<1 and k1(t)log2t121.k_{1}(t)\geq\log_{2}t^{\frac{1}{2}}-1.

Moreover, we can use embedding for Besov spaces to treat the situation where dpl=1d_{\text{pl}}=1 and 1<p2q<1<p\leq 2\leq q<\infty. Indeed, in this case we have

LpBp,20,Bq,20Lq.\displaystyle L^{p}\hookrightarrow B^{0}_{p,2},\quad B^{0}_{q,2}\hookrightarrow L^{q}.

Then, it holds

1(m,1(t,)f^)LqC11(m,1(t,)f^)Bq,20C2fBp,20C3fLp,\displaystyle\|\mathscr{F}^{-1}(m_{\infty,1}(t,\cdot)\widehat{f})\|_{L^{q}}\leq C_{1}\|\mathscr{F}^{-1}(m_{\infty,1}(t,\cdot)\widehat{f})\|_{B^{0}_{q,2}}\leq C_{2}\|f\|_{B^{0}_{p,2}}\leq C_{3}\|f\|_{L^{p}},

where the second inequality holds, since

1(ϕkm,1(t,)f^)Lq2Cj=k1k+11(ϕjf^)Lp2,\displaystyle\|\mathscr{F}^{-1}(\phi_{k}m_{\infty,1}(t,\cdot)\widehat{f})\|_{L^{q}}^{2}\leq C\sum\limits_{j=k-1}^{k+1}\|\mathscr{F}^{-1}(\phi_{j}\widehat{f})\|_{L^{p}}^{2},

so that

1(m,1(t,)f^)Bq,20=(j=01(ϕkm,1(t,)f^)Lq2)123C(k=01(ϕjf^)Lp2)12.\displaystyle\|\mathscr{F}^{-1}(m_{\infty,1}(t,\cdot)\widehat{f})\|_{B^{0}_{q,2}}=\left(\sum\limits_{j=0}^{\infty}\|\mathscr{F}^{-1}(\phi_{k}m_{\infty,1}(t,\cdot)\widehat{f})\|_{L^{q}}^{2}\right)^{\frac{1}{2}}\leq 3C\left(\sum\limits_{k=0}^{\infty}\|\mathscr{F}^{-1}(\phi_{j}\widehat{f})\|_{L^{p}}^{2}\right)^{\frac{1}{2}}.

Summing up, we have proven the following lemma.

Lemma 3.3.

Let 1pq1\leq p\leq q\leq\infty. Then, we have

K^(t,)MpqC{t1n2(1p1q)+dpl1,foranyt1,t1n2(1p1q),t(0,1),\|\widehat{K}_{\infty}(t,\cdot)\|_{M_{p}^{q}}\leq C\begin{cases}t^{1-\frac{n}{2}\left(\frac{1}{p}-\frac{1}{q}\right)+d_{\text{pl}}-1},&\ {\text{f}or\ any}\ t\geq 1,\\ t^{1-\frac{n}{2}\left(\frac{1}{p}-\frac{1}{q}\right)},&\ t\in(0,1),\end{cases}

provided that dpl(p,q)<1d_{\text{pl}}(p,q)<1. Moreover, these estimates remain true for 1<p2q<1<p\leq 2\leq q<\infty such that dpl(p,q)=1d_{\text{pl}}(p,q)=1.

3.2 Estimates for small frequencies

It is clear that K~0(t,)Mpq\tilde{K}_{0}(t,\cdot)\in M_{p}^{q} for all t[0,1]t\in[0,1] and for all 1pq1\leq p\leq q\leq\infty. Moreover,

K^0(t,)MpqCt,t[0,1].\|\widehat{K}_{0}(t,\cdot)\|_{M_{p}^{q}}\leq Ct,\quad t\in[0,1].

To estimate K^0(t,)Mpq\|\widehat{K}_{0}(t,\cdot)\|_{M_{p}^{q}} for t1t\geq 1, we are going to use again the stationary phase method as in the previous subsection. So, in order to get uniformly estimates w.r.t. time for the phase function, we perform the change of variable η=t14ξ\eta=t^{\frac{1}{4}}\xi. Thus, we obtain

K^0(t,)Mpq=tn4(1p1q)m0(t,)Mpq,\|\widehat{K}_{0}(t,\cdot)\|_{M_{p}^{q}}=t^{-\frac{n}{4}\left(\frac{1}{p}-\frac{1}{q}\right)}\|m_{0}(t,\cdot)\|_{M_{p}^{q}},

where we consider the radial multiplier

m0(t,η)=sin(t1+t1|η|4)1+t1|η|4)φ0(t,η)=sin(ω(t,|η|))1+t1|η|4)φ0(t,η),m_{0}(t,\eta)=\frac{\sin\big{(}t\sqrt{1+t^{-1}|\eta|^{4}}\big{)}}{\sqrt{1+t^{-1}|\eta|^{4})}}\varphi_{0}(t,\eta)=\frac{\sin(\omega(t,|\eta|))}{\sqrt{1+t^{-1}|\eta|^{4})}}\varphi_{0}(t,\eta),

where

φ0(t,η)=χ(t14|η|),ω(t,η)=t1+t1|η|4.\varphi_{0}(t,\eta)=\chi(t^{-\frac{1}{4}}|\eta|),\quad\omega(t,\eta)=t\sqrt{1+t^{-1}|\eta|^{4}}.

Now, we divide again our considerations in two cases, new small and large frequencies as follows:

m0,0(t,η)\displaystyle m_{0,0}(t,\eta) =χ(|η|)m0(t,η),\displaystyle=\chi(|\eta|)m_{0}(t,\eta),
m0,1(t,η)\displaystyle m_{0,1}(t,\eta) =(1χ(|η|))m0(t,η).\displaystyle=\big{(}1-\chi(|\eta|)\big{)}m_{0}(t,\eta).
Proposition 3.3.

For t1ρ4<1t^{-1}\rho^{4}<1 the following estimates hold:

|ρω(t,ρ)|\displaystyle|\partial_{\rho}^{\ell}\omega(t,\rho)| Cρ4,{0},\displaystyle\leq C_{\ell}\rho^{4-\ell},\qquad\ell\in\mathbb{N}\setminus\{0\}, (3.9)
|ρω(t,ρ)|\displaystyle|\partial_{\rho}^{\ell}\omega(t,\rho)| cρ4,=1,2,\displaystyle\geq c_{\ell}\rho^{4-\ell},\,\,\qquad\ell=1,2, (3.10)
|ρtw(t,ρ)|\displaystyle\big{|}\partial_{\rho}^{\ell}\frac{t}{w(t,\rho)}\big{|} C~ρ,.\displaystyle\leq\widetilde{C}_{\ell}\rho^{-\ell},\quad\,\,\,\,\,\,\,\,\,\ell\in\mathbb{N}. (3.11)
Proof.

We may estimate

ρω(t,ρ)\displaystyle\partial_{\rho}\omega(t,\rho) =2ρ3t1ρ4+1ρ3,\displaystyle=\frac{2\rho^{3}}{\sqrt{t^{-1}\rho^{4}+1}}\approx\rho^{3},
ρ2ω(t,ρ)\displaystyle\partial_{\rho}^{2}\omega(t,\rho) =ρ(2ρ3t1ρ4+1)=6ρ2t1ρ4+14ρ6t1(t1ρ4+1)t1ρ4+1ρ2.\displaystyle=\partial_{\rho}\Big{(}\frac{2\rho^{3}}{\sqrt{t^{-1}\rho^{4}+1}}\Big{)}=\frac{6\rho^{2}}{\sqrt{t^{-1}\rho^{4}+1}}-\frac{4\rho^{6}t^{-1}}{(t^{-1}\rho^{4}+1)\sqrt{t^{-1}\rho^{4}+1}}\approx\rho^{2}.

In the same way, we conclude (3.9). On the other hand, we have

|ρ1w(t,ρ)|w()w2+(w)w+1t2ρ4+t1ρ3,\big{|}\partial_{\rho}^{\ell}\frac{1}{w(t,\rho)}\big{|}\lesssim\frac{w^{(\ell)}}{w^{2}}+\frac{(w^{\prime})^{\ell}}{w^{\ell+1}}\lesssim t^{-2}\rho^{4-\ell}+t^{-\ell-1}\rho^{3\ell},

and so, using t1ρ4<1t^{-1}\rho^{4}<1, we obtain (3.11) and this concludes the proof. ∎

Proposition 3.4.

The determinant of the Hessian matrix satisfies the following estimate:

|detHω(t,|η|)|ρ2n.|\det H_{\omega(t,|\eta|)}|\geq\rho^{2n}.
Proof.

Following the proof of Proposition 3.2 and due to estimates (3.10), we arrive at

detHω=(1+βα1)αn=(ρ2ω)(|η|1ρω)n1ρ2n.\det H_{\omega}=(1+\beta\alpha^{-1})\alpha^{n}=(\partial_{\rho}^{2}\omega)(|\eta|^{-1}\partial_{\rho}\omega)^{n-1}\geq\rho^{2n}.

Lemma 3.4.

For any 1pq1\leq p\leq q\leq\infty, we have

m0,0(t,)MpqC.\displaystyle\|m_{0,0}(t,\cdot)\|_{M_{p}^{q}}\leq C.
Proof.

Since χ𝒞c\chi\in\mathcal{C}_{c}^{\infty} is independent of tt, we get (see Theorem 1.81.8 in [17])

m0,0(t,)MpqCmin{m0,0(t,)Mpp,m0,0(t,)Mqq}.\|m_{0,0}(t,\cdot)\|_{M_{p}^{q}}\leq C\min\big{\{}\|m_{0,0}(t,\cdot)\|_{M_{p}^{p}},\|m_{0,0}(t,\cdot)\|_{M_{q}^{q}}\big{\}}.

Let (p,q){(1,1),(,),(1,)}(p,q)\notin\{(1,1),(\infty,\infty),(1,\infty)\}. Using estimates (3.9) and (3.11), we have

|ρksinw(t,ρ)|ρ4k,k1,|\partial_{\rho}^{k}\sin w(t,\rho)|\leq\rho^{4-k},\quad k\geq 1,

for ρ1\rho\leq 1. Then, from Leibniz rule we may conclude that (ρ<1\rho<1)

|ρksinw(t,ρ)t/w(t,ρ)|1kρ4ρk++Cρkρk.|\partial_{\rho}^{k}\sin w(t,\rho)\,t/w(t,\rho)|\lesssim\sum_{1\leq\ell\leq k}\rho^{4-\ell}\rho^{-k+\ell}+C\rho^{-k}\lesssim\rho^{-k}.

Thus, by Mikhlin-Hörmander theorem (cf. Theorem A.2) we find that the multiplier norm is bounded by a uniform constant with respect to tt.

Now we consider (p,q)=(1,)(p,q)=(1,\infty). Then, we have

m0,0(t,)M1\displaystyle\|m_{0,0}(t,\cdot)\|_{M_{1}^{\infty}} =ηx1(χ(|η|)sinω(t,|η|)/1+t1|η|4)L\displaystyle=\big{\|}\mathcal{F}_{\eta\to x}^{-1}\big{(}\chi(|\eta|)\sin\omega(t,|\eta|)/\sqrt{1+t^{-1}|\eta|^{4}}\big{)}\big{\|}_{L^{\infty}}
|η|1|sinω(t,|η|)|𝑑η1.\displaystyle\lesssim\int_{|\eta|\leq 1}|\sin\omega(t,|\eta|)|d\eta\lesssim 1.

Finally we consider the case (p,q)=(1,1)(p,q)=(1,1). Using ρ<1\rho<1, t>1t>1 and the smoothness of the multiplier, we notice that we may improve estimate (3.11), obtaining

|ρktw(t,ρ)|Ck,\big{|}\partial_{\rho}^{k}\frac{t}{w(t,\rho)}\big{|}\leq C_{k},

so that we may estimate

|ρkm0,0(t,ρ)|C~k.|\partial_{\rho}^{k}m_{0,0}(t,\rho)|\lesssim\tilde{C}_{k}.

So, we may conclude the proof with integration by parts argument, as in the proof of Lemma 3.1. ∎

Lemma 3.5.

We have the following estimates:

m0,1(t,)MpqCtn4β(p,q) for 1pq,\|m_{0,1}(t,\cdot)\|_{M_{p}^{q}}\leq Ct^{\frac{n}{4}\beta(p,q)}\qquad\text{ for }1\leq p\leq q\leq\infty,

for any t1t\geq 1, where β=β(p,q)\beta=\beta(p,q) is defined in Theorem 2.1.

Proof.

Since Mpq=Mqp\|\cdot\|_{M_{p}^{q}}=\|\cdot\|_{M_{q^{\prime}}^{p^{\prime}}} it is enough to prove Lemma 3.5 for 1p21\leq p\leq 2 and pqpp\leq q\leq p^{\prime}.

We are going to use Littman’s lemma (cf. Lemma A.1) in order to find an estimate to the norm

m0,1(t,)ϕkM1=ξx1[(eiω(t,|ξ|)eiω(t,|ξ|))ϕk(ξ)φ0(t,η)(1χ)(η)1+t1|ξ|4]L.\|m_{0,1}(t,\cdot)\phi_{k}\|_{M_{1}^{\infty}}=\left\|\mathcal{F}_{\xi\to x}^{-1}\left[\big{(}e^{i\omega(t,|\xi|)}-e^{-i\omega(t,|\xi|)}\big{)}\phi_{k}(\xi)\frac{\varphi_{0}(t,\eta)(1-\chi)(\eta)}{\sqrt{1+t^{-1}|\xi|^{4}}}\right]\right\|_{L^{\infty}}.

Let η:=2kξ\eta:=2^{-k}\xi and dη=2kndξd\eta=2^{-kn}d\xi. Then, we set w(t,2kη):=24kw~(t,η)w(t,2^{k}\eta):=2^{4k}\widetilde{w}(t,\eta), with

w~(t,η):=w(24kt,η)=t228k+t24k|η|4.\widetilde{w}(t,\eta):=w(2^{-4k}t,\eta)=\sqrt{t^{2}2^{-8k}+t2^{-4k}|\eta|^{4}}.

Thus, we may conclude

ξx1[(eiω(t,|ξ|)\displaystyle\big{\|}\mathcal{F}_{\xi\to x}^{-1}\big{[}\big{(}e^{i\omega(t,|\xi|)} eiω(t,|ξ|))ϕk(ξ)11+t1|ξ|4]L\displaystyle-e^{-i\omega(t,|\xi|)}\big{)}\phi_{k}(\xi)\frac{1}{\sqrt{1+t^{-1}|\xi|^{4}}}\big{]}\big{\|}_{L^{\infty}}
2knηx1[ei24kw~(t,η)ϕ(|η|)11+t1|2kη|4]L\displaystyle\lesssim 2^{kn}\big{\|}\mathcal{F}_{\eta\to x}^{-1}\big{[}e^{i2^{4k}\widetilde{w}(t,\eta)}\phi(|\eta|)\frac{1}{\sqrt{1+t^{-1}|2^{k}\eta|^{4}}}\big{]}\big{\|}_{L^{\infty}}
C2kn(1+24k)n22kn,\displaystyle\leq C2^{kn}(1+2^{4k})^{-\frac{n}{2}}\approx 2^{-kn},

due to the fact that the derivatives of the function w~\widetilde{w} are uniformly bounded with respect to tt and kk on the support of ϕ\phi and also the amplitude function (1+t1|2kη|4)1/2(1+t^{-1}|2^{k}\eta|^{4})^{-1/2} is uniformly bounded with respect to t and k, together with all its η\eta-derivatives on the support of ϕ\phi.

From Young’s inequality we get

ξx1(m0,1(t,ξ)ϕk(ξ)(u1))L\displaystyle\big{\|}\mathcal{F}_{\xi\to x}^{-1}\big{(}m_{0,1}(t,\xi)\phi_{k}(\xi)\mathcal{F}(u_{1})\big{)}\big{\|}_{L^{\infty}} =ξx1(m0,1(t,ξ)ϕk(ξ))u1L\displaystyle=\big{\|}\mathcal{F}_{\xi\to x}^{-1}\big{(}m_{0,1}(t,\xi)\phi_{k}(\xi)\big{)}\ast u_{1}\big{\|}_{L^{\infty}}
ξx1(m0,1(t,ξ)ϕk(ξ))Lu1L1\displaystyle\leq\big{\|}\mathcal{F}_{\xi\to x}^{-1}\big{(}m_{0,1}(t,\xi)\phi_{k}(\xi)\big{)}\big{\|}_{L^{\infty}}\|u_{1}\|_{L^{1}}
C2knu1L1,\displaystyle\leq C2^{-kn}\|u_{1}\|_{L^{1}},

which allows to conclude the following:

m0,1(ξ)ϕk(|ξ|)M1C2kn.\|m_{0,1}(\xi)\phi_{k}(|\xi|)\|_{M_{1}^{\infty}}\leq C2^{-kn}. (3.12)

Using the fact that M22=LM_{2}^{2}=L^{\infty}, we obtain

m0,1(t,)ϕkM22=m0,1(t,)ϕkLC.\displaystyle\|m_{0,1}(t,\cdot)\phi_{k}\|_{M_{2}^{2}}=\|m_{0,1}(t,\cdot)\phi_{k}\|_{L^{\infty}}\leq C. (3.13)

Now, using (3.12), (3.13) and Riesz-Thorin interpolation theorem (cf. Theorem A.1) (L2L2L^{2}-L^{2} and L1LL^{1}-L^{\infty}) we obtain

m0,1(t,)ϕkMp0q0C2kn(12q0)with1p0+1q0=1.\|m_{0,1}(t,\cdot)\phi_{k}\|_{M_{p_{0}}^{q_{0}}}\leq C2^{-kn\big{(}1-\frac{2}{q}_{0}\big{)}}\qquad\text{with}\qquad\frac{1}{p_{0}}+\frac{1}{q_{0}}=1.

Finally, since we may estimate

|ρsin(w(t,ρ))|ρ3,\displaystyle|\partial_{\rho}^{\ell}sin(w(t,\rho))|\lesssim\rho^{3\ell},
|ρm0,1(t,ρ)|ρ3,\displaystyle|\partial_{\rho}^{\ell}m_{0,1}(t,\rho)|\lesssim\rho^{3\ell},

we have

m0,1(t,)ϕkL22kn/2.\|m_{0,1}(t,\cdot)\phi_{k}\|_{L^{2}}\lesssim 2^{kn/2}.

Hence, from Bernstein’s Theorem (cf. Theorem A.3) we get

m0,1(t,)ϕkM11m0,1(t,)ϕkL21n2N|β|=Nξβ(m0,1(t,)ϕk)L2n2NC22kn.\|m_{0,1}(t,\cdot)\phi_{k}\|_{M_{1}^{1}}\leq\|m_{0,1}(t,\cdot)\phi_{k}\|_{L^{2}}^{1-\frac{n}{2N}}\sum_{|\beta|=N}\|\partial_{\xi}^{\beta}\big{(}m_{0,1}(t,\cdot)\phi_{k}\big{)}\|_{L^{2}}^{\frac{n}{2N}}\leq C2^{2kn}. (3.14)

Using (3.12), (3.14) and again Riesz-Thorin interpolation theorem (Lp0Lq0L^{p_{0}}-L^{q_{0}} and L1L1L^{1}-L^{1}), we conclude

m0,1(t,)ϕkMpqC2kn(1p+3q2),1p+1q1.\|m_{0,1}(t,\cdot)\phi_{k}\|_{M_{p}^{q}}\leq C2^{kn\left(\frac{1}{p}+\frac{3}{q}-2\right)},\quad\frac{1}{p}+\frac{1}{q}\geq 1.

Now, we have that

m0,1(t,)Mpqk=0k0(t)m0,1(t,)ϕkMpq,\displaystyle\|m_{0,1}(t,\cdot)\|_{M_{p}^{q}}\leq\sum_{k=0}^{k_{0}(t)}\|m_{0,1}(t,\cdot)\phi_{k}\|_{M_{p}^{q}},

where k0(t)=max{k: 2k1t14}.k_{0}(t)=\max{\{k\in\mathbb{Z}\,:\,2^{k-1}\leq t^{\frac{1}{4}}\}}. Hence, we have

m0,1(t,)Mpqk=0k0(t)m0(t,)ϕkMpqCk=0k0(t)2kn(1p+3q2),\displaystyle\|m_{0,1}(t,\cdot)\|_{M_{p}^{q}}\leq\sum_{k=0}^{k_{0}(t)}\|m_{0}(t,\cdot)\phi_{k}\|_{M_{p}^{q}}\leq C\sum_{k=0}^{k_{0}(t)}2^{kn\left(\frac{1}{p}+\frac{3}{q}-2\right)},

which converges (and with an estimate independent of time) if and only if

1p+3q2<0.\frac{1}{p}+\frac{3}{q}-2<0.

In the situation where 1p+3q2=0\frac{1}{p}+\frac{3}{q}-2=0 we may use embeddings for Besov spaces as in the previous section. Let 1<p2q<1<p\leq 2\leq q<\infty (the line 1p+3q2=0\frac{1}{p}+\frac{3}{q}-2=0 in the 1p1q\frac{1}{p}-\frac{1}{q} plane is localized in the zone 1p2q31\leq p\leq 2\leq q\leq 3). Since

LpBp,max{p,2}0,Bq,min{2,q}0Lq,\displaystyle L^{p}\hookrightarrow B^{0}_{p,\max\{p,2\}},\quad B^{0}_{q,\min\{2,q\}}\hookrightarrow L^{q},

we have

1(m0,1(t,)f^)LqC11(m0,1(t,)f^)Bq,20C2fBp,20C3fLp,\displaystyle\|\mathscr{F}^{-1}(m_{0,1}(t,\cdot)\widehat{f})\|_{L^{q}}\leq C_{1}\|\mathscr{F}^{-1}(m_{0,1}(t,\cdot)\widehat{f})\|_{B^{0}_{q,2}}\leq C_{2}\|f\|_{B^{0}_{p,2}}\leq C_{3}\|f\|_{L^{p}},

where the second inequality holds, since

1(ψkm0,1(t,)f^)Lq2Cj=k1k+11(ψjf^)Lp2,\displaystyle\|\mathscr{F}^{-1}(\psi_{k}m_{0,1}(t,\cdot)\widehat{f})\|_{L^{q}}^{2}\leq C\sum\limits_{j=k-1}^{k+1}\|\mathscr{F}^{-1}(\psi_{j}\widehat{f})\|_{L^{p}}^{2},

so that

1(m0,1(t,)f^)Bq,20=(k=0k01(ψkm0,1(t,)f^)Lq2)123C(k=0k01(ψkf^)Lp2)12.\displaystyle\|\mathscr{F}^{-1}(m_{0,1}(t,\cdot)\widehat{f})\|_{B^{0}_{q,2}}=\left(\sum\limits_{k=0}^{k_{0}}\|\mathscr{F}^{-1}(\psi_{k}m_{0,1}(t,\cdot)\widehat{f})\|_{L^{q}}^{2}\right)^{\frac{1}{2}}\leq 3C\left(\sum\limits_{k=0}^{k_{0}}\|\mathscr{F}^{-1}(\psi_{k}\widehat{f})\|_{L^{p}}^{2}\right)^{\frac{1}{2}}.

In the case p=1p=1 and q=3q=3, using only the second embedding, we obtain

1(m0,1(t,)f^)LqC11(m0,1(t,)f^)Bq,min{2,3}0C2(k=0k01(ψkf^)L12)12\displaystyle\|\mathscr{F}^{-1}(m_{0,1}(t,\cdot)\widehat{f})\|_{L^{q}}\leq C_{1}\|\mathscr{F}^{-1}(m_{0,1}(t,\cdot)\widehat{f})\|_{B^{0}_{q,\min\{2,3\}}}\leq C_{2}\left(\sum\limits_{k=0}^{k_{0}}\|\mathscr{F}^{-1}(\psi_{k}\widehat{f})\|_{L^{1}}^{2}\right)^{\frac{1}{2}}
C3fL1(k=0k01)12C4(logt)12fL1.\displaystyle\leq C_{3}\|f\|_{L^{1}}\left(\sum\limits_{k=0}^{k_{0}}1\right)^{\frac{1}{2}}\leq C_{4}(\log t)^{\frac{1}{2}}\|f\|_{L^{1}}.

Finally, we assume β(p,q)=1p+3q2>0\beta(p,q)=\frac{1}{p}+\frac{3}{q}-2>0. In this case, we may estimate

m0,1(t,)Mpq2{k0(t)+1}nβ(p,q)12nβ(p,q)12k0(t)nβ(p,q)tnβ(p,q)4.\displaystyle\|m_{0,1}(t,\cdot)\|_{M_{p}^{q}}\leq\frac{2^{\{k_{0}(t)+1\}n\beta(p,q)}-1}{2^{n\beta(p,q)}-1}\lesssim 2^{k_{0}(t)n\beta(p,q)}\approx t^{\frac{n\beta(p,q)}{4}}.

Summing up, we have proven the following result.

Lemma 3.6.

Let 1pq1\leq p\leq q\leq\infty. Then, setting β(p,q)\beta(p,q) as in Theorem 2.1, we have

K^0(t,)MpqC{tn4(1p1q)+β(p,q)n4,foranyt1,t,t[0,1),\|\widehat{K}_{0}(t,\cdot)\|_{M_{p}^{q}}\leq C\begin{cases}t^{-\frac{n}{4}\left(\frac{1}{p}-\frac{1}{q}\right)+\beta(p,q)\frac{n}{4}},&\ {\text{f}or\ any}\ t\geq 1,\\ t,&\ t\in[0,1),\end{cases}

provided that in the case (p,q)=(1,3)(p,q)=(1,3) or (p,q)=(3/2,)(p,q)=(3/2,\infty) with dpl<1d_{\text{pl}}<1, estimate (3.6) is replaced by

K^0(t,)MpqCtn4(1p1q)(logt)12,\|\widehat{K}_{0}(t,\cdot)\|_{M_{p}^{q}}\leq Ct^{-\frac{n}{4}\left(\frac{1}{p}-\frac{1}{q}\right)}(\log t)^{\frac{1}{2}},

for any t1t\geq 1.

Proof of Theorem 2.1.

Due to

1n2(1p1q)+dpl1n4(1p1q)+β(p,q)n4,1-\frac{n}{2}\left(\frac{1}{p}-\frac{1}{q}\right)+d_{\text{pl}}-1\leq-\frac{n}{4}\left(\frac{1}{p}-\frac{1}{q}\right)+\beta(p,q)\frac{n}{4},

the proof of Theorem 2.1 follows by Lemma 3.3 and Lemma 3.6. ∎

4 Proof of local and global in time existence results

Let us first introduce some notations which will be used in the proof of Theorems 2.2, 2.5 and 2.6. Let ulin(t,x)u^{\operatorname{lin}}(t,x) be the solution to (1.1) and K=K(t,x)K=K(t,x) be the fundamental solution to the linear Cauchy problem (1.3). Let now NN be the operator

N:uX(T)Nu(t,x)=ulin(t,x)+unl(t,x),N:u\in X(T)\rightarrow Nu(t,x)=u^{\operatorname{lin}}(t,x)+u^{\operatorname{nl}}(t,x),

where X(T)X(T) is a suitable space where we search for solutions and

unl(t,x)\displaystyle u^{\operatorname{nl}}(t,x) =0tK(tτ,x)(x)|u(τ,x)|α𝑑τ.\displaystyle=\int_{0}^{t}K(t-\tau,x)\ast_{(x)}|u(\tau,x)|^{\alpha}d\tau.

Thanks to Duhamel’s principle, the solution of (1.5) is given by the fixed point of the operator NN. To prove that NN admits a fixed point, we are going to show that the mapping NN satisfies the following estimates:

NuX(T)\displaystyle\|Nu\|_{X(T)} C0(u0,u1)𝒜+C1(t)uX(T)α,\displaystyle\leq C_{0}\|(u_{0},u_{1})\|_{\mathcal{A}}+C_{1}(t)\|u\|_{X(T)}^{\alpha}, (4.1)
NuNvX(T)\displaystyle\|Nu-Nv\|_{X(T)} C2(t)uvX(T)(uX(T)α1+vX(T)α1),\displaystyle\leq C_{2}(t)\|u-v\|_{X(T)}\Big{(}\|u\|_{X(T)}^{\alpha-1}+\|v\|_{X(T)}^{\alpha-1}\Big{)}, (4.2)

where 𝒜\mathcal{A} stands for the space which the data (u0,u1)(u_{0},u_{1}) belong to. In this way, by using contraction argument, we get simultaneously a unique solution to Nu=uNu=u locally in time for large data and globally in time for small data. To prove the local in time existence we use that C1(t)C_{1}(t) and C2(t)C_{2}(t) tend to zero as tt tends to zero, whereas to prove global in time existence we use that max{C1(t),C2(t)}C\max\{C_{1}(t),C_{2}(t)\}\leq C for all t[0,)t\in[0,\infty) for a suitable nonnegative constant CC.

Proof of Theorem 2.2.

Let r=2n[n4]+r=\frac{2n}{[n-4]_{+}}. For n4n\neq 4 we define the Banach space

X(T):={u𝒞([0,T],L2(n)Lr(n)):uX(T):=supt[0,T],q[2,r]u(t,)Lq<},X(T):=\bigl{\{}u\in\mathcal{C}([0,T],L^{2}(\mathbb{R}^{n})\cap L^{r}(\mathbb{R}^{n}))\,:\ \|u\|_{X(T)}:=\sup_{t\in[0,T],q\in[2,r]}\|u(t,\cdot)\|_{L^{q}}<\infty\bigr{\}}\,,

and for n=4n=4 we consider

X(T):={u𝒞([0,T],L2(n)Lr(n)):uX(T):=supt[0,T],q[2,r)u(t,)Lq<}.X(T):=\bigl{\{}u\in\mathcal{C}([0,T],L^{2}(\mathbb{R}^{n})\cap L^{r}(\mathbb{R}^{n}))\,:\ \|u\|_{X(T)}:=\sup_{t\in[0,T],q\in[2,r)}\|u(t,\cdot)\|_{L^{q}}<\infty\bigr{\}}\,.

Due to the hypothesis (u0,u1)H2(n)×L2(n)(u_{0},u_{1})\in H^{2}(\mathbb{R}^{n})\times L^{2}(\mathbb{R}^{n}), it follows from Gagliardo-Nirenberg inequality (see Proposition A.1) that

ulin(t,)Lqulin(t,)H2˙θulin(t,)L21θE(ulin)(t)=E(ulin)(0)u0H2+u1L2,\|u^{\operatorname{lin}}(t,\cdot)\|_{L^{q}}\lesssim\|u^{\operatorname{lin}}(t,\cdot)\|_{\dot{H^{2}}}^{\theta}\|u^{\operatorname{lin}}(t,\cdot)\|_{L^{2}}^{1-\theta}\lesssim E(u^{\operatorname{lin}})(t)=E(u^{\operatorname{lin}})(0)\leq\|u_{0}\|_{H^{2}}+\|u_{1}\|_{L^{2}},

for all t0t\geq 0 and q[2,r]q\in[2,r] when n4n\neq 4 and for all q[2,r)q\in[2,r) when n=4n=4. Here θ=n2(121q)[0,1]\theta=\frac{n}{2}\left(\frac{1}{2}-\frac{1}{q}\right)\in[0,1]. Therefore ulinX(T)u^{\operatorname{lin}}\in X(T).

In the following argument we have to exclude again the case q=rq=r when n=4n=4. Let us fix α(1,n+4[n4]+)\alpha\in\left(1,\frac{n+4}{[n-4]_{+}}\right). We then note that for any q[2,r]q\in[2,r] there exists p=p(α,q)(2nn+4,2]p=p(\alpha,q)\in\left(\frac{2n}{n+4},2\right] such that

2pαranddpl(p,q)<1.2\leq p\alpha\leq r\quad\text{and}\quad d_{\text{pl}}(p,q)<1.

So, using Minkowski integral inequality and again the estimates of Theorem 2.1 we get

unl(t,)Lq\displaystyle\|u^{\operatorname{nl}}(t,\cdot)\|_{L^{q}} 0tK(ts,)(x)|u(s,)|αLq𝑑s\displaystyle\leq\int_{0}^{t}\|K(t-s,\cdot)\ast_{(x)}|u(s,\cdot)|^{\alpha}\|_{L^{q}}\,ds
0t(ts)1n2(1p1q)u(s,)Lpαα𝑑s\displaystyle\lesssim\int_{0}^{t}(t-s)^{1-\frac{n}{2}\left(\frac{1}{p}-\frac{1}{q}\right)}\|u(s,\cdot)\|^{\alpha}_{L^{p\alpha}}\,ds
uX(T)α0t(ts)1n2(1p1q)𝑑st2n2(1p1q)uX(T)α,\displaystyle\lesssim\|u\|_{X(T)}^{\alpha}\int_{0}^{t}(t-s)^{1-\frac{n}{2}\left(\frac{1}{p}-\frac{1}{q}\right)}\,ds\lesssim t^{2-\frac{n}{2}\left(\frac{1}{p}-\frac{1}{q}\right)}\|u\|^{\alpha}_{X(T)},

for any q[2,r]q\in[2,r]. Since p>2nn+4p>\frac{2n}{n+4} we get

2n2p+n2q>0,q[2,r].2-\frac{n}{2p}+\frac{n}{2q}>0,\quad\forall q\in[2,r].

Therefore, NN maps X(T)X(T) into itself and, if TT is sufficiently small, then the existence of a unique local solution follows by a contraction argument (for more details, see the proof of Theorem 2.5). ∎

In order to prove the global (in time) existence results with data u00u_{0}\equiv 0 and u1L1u_{1}\in L^{1} the following lemma comes into play.

Lemma 4.1.

Let ν>1\nu>-1 and μ\mu\in\mathbb{R}. Then, it holds

0t(ts)ν(1+s)μ𝑑s{(1+t)ν, if μ<1,(1+t)νlog(e+t), if μ=1,(1+t)1+ν+μ, if μ>1,\int_{0}^{t}(t-s)^{\nu}\,(1+s)^{\mu}\,ds\lesssim\begin{cases}(1+t)^{\nu},&\mbox{ if }\,\,\mu<-1,\\ (1+t)^{\nu}\,log(e+t),&\mbox{ if }\,\,\mu=-1,\\ (1+t)^{1+\nu+\mu},&\mbox{ if }\,\,\mu>-1,\end{cases}

and

0t(ts)νec(ts)(1+s)μ𝑑s(1+t)μ.\int_{0}^{t}(t-s)^{\nu}\,e^{-c(t-s)}(1+s)^{\mu}\,ds\lesssim(1+t)^{\mu}.

Moreover, the estimate is also valid if (ts)ν(t-s)^{\nu} is replaced by (1+ts)ν(1+t-s)^{\nu} in the integral.

Proof of Theorem 2.5.

We introduce the solution space X(T)X(T) by

X(T):={u𝒞([0,T],L2(n)L(n)):uX(T)<},\displaystyle X(T):=\big{\{}u\in\mathcal{C}\big{(}[0,T],L^{2}(\mathbb{R}^{n})\cap L^{\infty}(\mathbb{R}^{n})\big{)}:\|u\|_{X(T)}<\infty\big{\}},

where

uX(T)\displaystyle\|u\|_{X(T)} :=supt[0,T],q[2,3)(1+t)n4(11q)n4β(1,q)uLq+(1+t)n4(113)(log(1+t))12uL3\displaystyle:=\sup\limits_{t\in[0,T],\,q\in[2,3)}(1+t)^{\frac{n}{4}\left(1-\frac{1}{q}\right)-\frac{n}{4}\beta(1,q)}\|u\|_{L^{q}}+(1+t)^{\frac{n}{4}\left(1-\frac{1}{3}\right)}\,(\log(1+t))^{-\frac{1}{2}}\,\|u\|_{L^{3}}
+supt[0,T],q(3,](1+t)n4(11q)uLq.\displaystyle\quad+\sup\limits_{t\in[0,T],q\in(3,\infty]}(1+t)^{\frac{n}{4}\left(1-\frac{1}{q}\right)}\|u\|_{L^{q}}.

Since u1L1(n)=𝒜u_{1}\in L^{1}(\mathbb{R}^{n})=\mathcal{A}, from (2.2) we immediately conclude

ulinX(T)=K(t,x)u1(x)X(T)C0u1𝒜.\|u^{\operatorname{lin}}\|_{X(T)}=\|K(t,x)\ast u_{1}(x)\|_{X(T)}\leq C_{0}\|u_{1}\|_{\mathcal{A}}.

Therefore, it remains to prove

unlX(T)uX(T)α.\|u^{\operatorname{nl}}\|_{X(T)}\lesssim\|u\|_{X(T)}^{\alpha}.

For the sake of brevity we shall consider only the case q3q\neq 3. Using the estimates from (2.2), we have the following:

unl(t,)Lq\displaystyle\|u^{\operatorname{nl}}(t,\cdot)\|_{L^{q}} 0tK(tτ,x)(x)|u(τ,x)|αLq𝑑τ\displaystyle\lesssim\int_{0}^{t}\big{\|}K(t-\tau,x)\ast_{(x)}|u(\tau,x)|^{\alpha}\big{\|}_{L^{q}}d\tau
0t(1+tτ)n4(11q)+n4β(1,q)|u(τ,x)|αL1𝑑τ\displaystyle\lesssim\int_{0}^{t}(1+t-\tau)^{-\frac{n}{4}\left(1-\frac{1}{q}\right)+\frac{n}{4}\beta(1,q)}\,\||u(\tau,x)|^{\alpha}\|_{L^{1}}d\tau
0t(1+tτ)n4(11q)+n4β(1,q)uLαα𝑑τ\displaystyle\lesssim\int_{0}^{t}(1+t-\tau)^{-\frac{n}{4}\left(1-\frac{1}{q}\right)+\frac{n}{4}\beta(1,q)}\|u\|_{L^{\alpha}}^{\alpha}d\tau
uX(T)α0t(1+tτ)n4(11q)+n4β(1,q)(1+τ)α(n4(11α)+n4β(1,α))𝑑τ.\displaystyle\lesssim\|u\|_{X(T)}^{\alpha}\int_{0}^{t}(1+t-\tau)^{-\frac{n}{4}\left(1-\frac{1}{q}\right)+\frac{n}{4}\beta(1,q)}(1+\tau)^{-\alpha\left(\frac{n}{4}(1-\frac{1}{\alpha})+\frac{n}{4}\beta(1,\alpha)\right)}d\tau.

We notice that for n=1,2n=1,2 it holds

n4(11q)+n4β(1,q)>1.-\frac{n}{4}\Big{(}1-\frac{1}{q}\Big{)}+\frac{n}{4}\beta(1,q)>-1.

Moreover, since α>1+4n>3\alpha>1+\frac{4}{n}>3 ( for n=1,2n=1,2) we have β(1,α)=0\beta(1,\alpha)=0 and therefore

α>1+4nn4(α1)<1.\alpha>1+\frac{4}{n}\iff-\frac{n}{4}\left(\alpha-1\right)<-1.

So, we can use Lemma 4.1 to get

unl(t,)Lq(1+t)n4(1nq)+n4β(1,q)uX(T)α.\displaystyle\|u^{\operatorname{nl}}(t,\cdot)\|_{L^{q}}\lesssim(1+t)^{-\frac{n}{4}\left(1-\frac{n}{q}\right)+\frac{n}{4}\beta(1,q)}\|u\|_{X(T)}^{\alpha}.

Next we turn to the proof of the Lipschitz condition (4.2). Due to the fact that

||u(τ,x)|α|v(τ,x)|α||u(s,x)v(s,x)|(|u(s,x)|α1+|v(s,x)|α1),|\,|u(\tau,x)|^{\alpha}-|v(\tau,x)|^{\alpha}|\lesssim|u(s,x)-v(s,x)|\big{(}|u(s,x)|^{\alpha-1}+|v(s,x)|^{\alpha-1}\big{)},

from Hölder’s inequality we obtain

|u(s,)|α|v(s,)|αL1\displaystyle\big{\|}|u(s,\cdot)|^{\alpha}-|v(s,\cdot)|^{\alpha}\big{\|}_{L^{1}} u(s,)v(s,)Lα(u(s,)Lαα1+v(s,)Lαα1).\displaystyle\lesssim\|u(s,\cdot)-v(s,\cdot)\|_{L^{\alpha}}\big{(}\|u(s,\cdot)\|_{L^{\alpha}}^{\alpha-1}+\|v(s,\cdot)\|_{L^{\alpha}}^{\alpha-1}\big{)}.

Thus, following the same ideas to what we did to estimate unl(t,)Lq\|u^{\operatorname{nl}}(t,\cdot)\|_{L^{q}}, we can obtain (4.2). In this way the proof of Theorem 2.5 is completed. ∎

Proof of Theorem 2.6..

We introduce the solution space X(T)X(T) by

X(T):={u𝒞([0,T],L2(n)Lp~(n)):uX(T)<},\displaystyle X(T):=\big{\{}u\in\mathcal{C}\big{(}[0,T],L^{2}(\mathbb{R}^{n})\cap L^{\tilde{p}}(\mathbb{R}^{n})\big{)}:\|u\|_{X(T)}<\infty\big{\}},

where we set p~=\tilde{p}=\infty for n=3n=3 and p~<\tilde{p}<\infty for n=4n=4, and the corresponding norm

uX(T)\displaystyle\|u\|_{X(T)} :=supt[0,T],q[2,3)(1+t)n4(11q)n4β(1,q)uLq+(1+t)n4(113)(log(1+t))12uL3\displaystyle:=\sup\limits_{t\in[0,T],q\in[2,3)}(1+t)^{\frac{n}{4}\left(1-\frac{1}{q}\right)-\frac{n}{4}\beta(1,q)}\|u\|_{L^{q}}+(1+t)^{\frac{n}{4}\left(1-\frac{1}{3}\right)}\,(\log(1+t))^{-\frac{1}{2}}\,\|u\|_{L^{3}}
+supt[0,T],q(3,p~](1+t)n4(11q)uLq.\displaystyle\quad+\sup\limits_{t\in[0,T],q\in(3,\tilde{p}]}(1+t)^{\frac{n}{4}\left(1-\frac{1}{q}\right)}\|u\|_{L^{q}}.

On one hand we have d(1,q)=n2q<1d(1,q)=\frac{n}{2q}<1 for n=3n=3 or n=4n=4 with q>2q>2, and on the other hand d(2,q)=n4n2q<1d(2,q)=\frac{n}{4}-\frac{n}{2q}<1 when n=3n=3 or n=4n=4 with 2q<2\leq q<\infty. Hence, we get

ulinX(T)u1L1L2.\displaystyle\|u^{\operatorname{lin}}\|_{X(T)}\lesssim\|u_{1}\|_{L^{1}\cap L^{2}}.

Now we deal with the nonlinear part. For the sake of brevity, we consider only the case q3q\neq 3. Using L1L2LqL^{1}\cap L^{2}-L^{q} estimates for q(2,p~]q\in(2,\tilde{p}], we get

unl(t,)Lq\displaystyle\|u^{\operatorname{nl}}(t,\cdot)\|_{L^{q}} 0tK(tτ,x)|u(τ,)|αLq𝑑τ\displaystyle\lesssim\int_{0}^{t}\big{\|}K(t-\tau,x)\ast|u(\tau,\cdot)|^{\alpha}\big{\|}_{L^{q}}\,d\tau
0t(1+tτ)n4(11q)+n4β(1,q)|u(τ,)|αL1L2𝑑τ\displaystyle\lesssim\int_{0}^{t}(1+t-\tau)^{-\frac{n}{4}\left(1-\frac{1}{q}\right)+\frac{n}{4}\beta(1,q)}\,\|\,|u(\tau,\cdot)|^{\alpha}\|_{L^{1}\cap L^{2}}d\tau
0t(1+tτ)n4(11q)+n4β(1,q)(u(τ,)Lαα+u(τ,)L2αα)𝑑τ.\displaystyle\lesssim\int_{0}^{t}(1+t-\tau)^{-\frac{n}{4}\left(1-\frac{1}{q}\right)+\frac{n}{4}\beta(1,q)}\big{(}\|u(\tau,\cdot)\|_{L^{\alpha}}^{\alpha}+\|u(\tau,\cdot)\|_{L^{2\alpha}}^{\alpha}\big{)}d\tau.

We note that in both cases, q3q\leq 3 or q>3q>3, we have

n4(11q)+n4β(1,q)>1-\frac{n}{4}\left(1-\frac{1}{q}\right)+\frac{n}{4}\beta(1,q)>-1

and

u(τ,)Lαα(1+τ)σ(log(1+τ))γuX(T)α,\|u(\tau,\cdot)\|_{L^{\alpha}}^{\alpha}\lesssim(1+\tau)^{-\sigma}(\log(1+\tau))^{\gamma}\|u\|_{X(T)}^{\alpha},

with

σ={n4(α1)α>3,n(α21)α3.\sigma=\begin{cases}\frac{n}{4}\left(\alpha-1\right)&\alpha>3,\\ n\left(\frac{\alpha}{2}-1\right)&\alpha\leq 3.\end{cases} (4.3)

Moreover, we can see that u(τ,)L2αα\|u(\tau,\cdot)\|_{L^{2\alpha}}^{\alpha} decays faster than u(τ,)Lαα\|u(\tau,\cdot)\|_{L^{\alpha}}^{\alpha}, since

n4(112α)n2+nα-\frac{n}{4}\left(1-\frac{1}{2\alpha}\right)\leq-\frac{n}{2}+\frac{n}{\alpha}

if α3\alpha\leq 3 and by definition of uX(T)u\in X(T) for α3\alpha\geq 3. Thus, using α>2+2n\alpha>2+\frac{2}{n} we conclude that σ>1\sigma>1, with σ\sigma given by (4.3) and Lemma 4.1 implies

unl(t,)Lq(1+t)n4(1nq)+β4uX(T)α,\displaystyle\|u^{\operatorname{nl}}(t,\cdot)\|_{L^{q}}\lesssim(1+t)^{-\frac{n}{4}\left(1-\frac{n}{q}\right)+\frac{\beta}{4}}\|u\|_{X(T)}^{\alpha},

for q(2,p~]q\in(2,\tilde{p}]. Finally, we consider the case q=2q=2. We have

unl(t,)L2\displaystyle\|u^{\operatorname{nl}}(t,\cdot)\|_{L^{2}} 0tK(tτ,x)|u(τ,)|αL2𝑑τ\displaystyle\lesssim\int_{0}^{t}\big{\|}K(t-\tau,x)\ast|u(\tau,\cdot)|^{\alpha}\big{\|}_{L^{2}}\,d\tau
0t|u(τ,)|αL2𝑑τ0tu(τ,)L2αα𝑑τuX(T)α0t(1+τ)n4(112α)α𝑑τ.\displaystyle\lesssim\int_{0}^{t}\,\|\,|u(\tau,\cdot)|^{\alpha}\|_{L^{2}}d\tau\lesssim\int_{0}^{t}\|u(\tau,\cdot)\|_{L^{2\alpha}}^{\alpha}\,d\tau\lesssim\|u\|_{X(T)}^{\alpha}\,\int_{0}^{t}(1+\tau)^{-\frac{n}{4}\left(1-\frac{1}{2\alpha}\right)\alpha}\,d\tau.

Now, since α>α~c=2+2n\alpha>\tilde{\alpha}_{c}=2+\frac{2}{n}, we get

n4(112α)α<1-\frac{n}{4}\left(1-\frac{1}{2\alpha}\right)\alpha<-1

and so, using Lemma 4.1, we conclude

unl(t,)L2uX(T)α.\|u^{\operatorname{nl}}(t,\cdot)\|_{L^{2}}\lesssim\|u\|_{X(T)}^{\alpha}.

Following as in the proof of Theorem 2.5, one may conclude the Lipschitz condition (4.2) and this concludes the proof.

5 Non-local existence result

This section is devoted to the proof of Theorem 2.4. Let η(t)𝒞c([0,))\eta(t)\in\mathcal{C}^{\infty}_{c}([0,\infty)) and ϕ(x)𝒞c(n)\phi(x)\in\mathcal{C}^{\infty}_{c}(\mathbb{R}^{n}) such that 0η,ϕ10\leq\eta,\phi\leq 1 and

η(t)={1,0t12,0,t1,ϕ(x)={1,|x|12,0,|x|1.\eta(t)=\begin{cases}1,\quad 0\leq t\leq\frac{1}{2},\\ 0,\quad t\geq 1,\end{cases}\quad\phi(x)=\begin{cases}1,\quad|x|\leq\frac{1}{2},\\ 0,\quad|x|\geq 1.\end{cases}

We remark that for any r>1r>1 there exists Cr>0C_{r}>0 such that

|t2η(t)|Crη(t)1r,|Δx2ϕ(x)|Crϕ(x)1r.|\partial^{2}_{t}\eta(t)|\leq C_{r}\eta(t)^{\frac{1}{r}},\quad|\Delta^{2}_{x}\phi(x)|\leq C_{r}\phi(x)^{\frac{1}{r}}.

For a parameter τ(0,1]\tau\in(0,1] we define

ψτ(t,x)=ητ(t)ϕτ(x),ητ(t)=η(tτ)andϕτ(x)=ϕ(xτ12).\psi_{\tau}(t,x)=\eta_{\tau}(t)\phi_{\tau}(x),\quad\eta_{\tau}(t)=\eta\left(\frac{t}{\tau}\right)\quad\text{and}\quad\phi_{\tau}(x)=\phi\left(\frac{x}{\tau^{\frac{1}{2}}}\right).

In the sequel we consider the following positive quantity:

I(τ)=0Tn|u(t,x)|αψτ(t,x)𝑑x𝑑t.I(\tau)=\int_{0}^{T}\int_{\mathbb{R}^{n}}|u(t,x)|^{\alpha}\psi_{\tau}(t,x)dxdt.

In the following Lemma 5.1 we shall employ the well-known test function method to derive an integral inequality.

Lemma 5.1.

Suppose that uLlocα([0,T)×n)u\in L^{\alpha}_{\text{loc}}([0,T)\times\mathbb{R}^{n}) is a weak solution to (1.5) with u00u_{0}\equiv 0. Then, there exists C=C(α,n)>0C=C(\alpha,n)>0 such that

|0Tnu(t,x){t2+Δx2+1}ψτ(t,x)𝑑x𝑑t|Cτ2+(1+n2)1αI(τ)1α,\left|\int_{0}^{T}\int_{\mathbb{R}^{n}}u(t,x)\{\partial^{2}_{t}+\Delta^{2}_{x}+1\}\psi_{\tau}(t,x)dxdt\right|\leq C\tau^{-2+\left(1+\frac{n}{2}\right)\frac{1}{\alpha^{\prime}}}I(\tau)^{\frac{1}{\alpha}},

where α\alpha^{\prime} denotes the conjugate exponent of α\alpha.

Proof.

From Hölder inequality we get

|0Tnu(t,x)t2ψτ(t,x)dxdt|\displaystyle\left|\int_{0}^{T}\int_{\mathbb{R}^{n}}u(t,x)\partial^{2}_{t}\psi_{\tau}(t,x)dxdt\right| τ20Tn|u(t,x)|ψτ(t,x)1α|η′′(tτ)|ητ(t)1αϕτ(x)1αϕτ(x)𝑑x𝑑t\displaystyle\leq\tau^{-2}\int_{0}^{T}\int_{\mathbb{R}^{n}}|u(t,x)|\psi_{\tau}(t,x)^{\frac{1}{\alpha}}\left|\eta^{\prime\prime}\left(\frac{t}{\tau}\right)\right|\eta_{\tau}(t)^{-\frac{1}{\alpha}}\phi_{\tau}(x)^{\frac{1}{\alpha^{\prime}}}\phi_{\tau}(x)dxdt
Cατ2I(τ)1α{0τ|x|τ2𝑑x𝑑t}1α\displaystyle\leq C_{\alpha}\tau^{-2}I(\tau)^{\frac{1}{\alpha}}\left\{\int_{0}^{\tau}\int_{|x|\leq\frac{\tau}{2}}dxdt\right\}^{\frac{1}{\alpha^{\prime}}}
Cα,nτ2+(1+n2)1αI(τ)1α.\displaystyle\leq C_{\alpha,n}\tau^{-2+\left(1+\frac{n}{2}\right)\frac{1}{\alpha^{\prime}}}I(\tau)^{\frac{1}{\alpha}}.

Analogously, we have

|0Tnu(t,x)Δx2ψτ(t,x)𝑑x𝑑t|Cα,nτ2+(1+n2)1αI(τ)1α.\left|\int_{0}^{T}\int_{\mathbb{R}^{n}}u(t,x)\Delta^{2}_{x}\psi_{\tau}(t,x)dxdt\right|\leq C_{\alpha,n}\tau^{-2+\left(1+\frac{n}{2}\right)\frac{1}{\alpha^{\prime}}}I(\tau)^{\frac{1}{\alpha}}.

On the other hand, since τ1\tau\leq 1 we get

|0Tnu(t,x)ψτ(t,x)𝑑x𝑑t|Cα,nτ(1+n2)1αI(τ)1αCα,nτ2+(1+n2)1αI(τ)1α.\left|\int_{0}^{T}\int_{\mathbb{R}^{n}}u(t,x)\psi_{\tau}(t,x)dxdt\right|\leq C_{\alpha,n}\tau^{\left(1+\frac{n}{2}\right)\frac{1}{\alpha^{\prime}}}I(\tau)^{\frac{1}{\alpha}}\leq C_{\alpha,n}\tau^{-2+\left(1+\frac{n}{2}\right)\frac{1}{\alpha^{\prime}}}I(\tau)^{\frac{1}{\alpha}}.

This completes the proof. ∎

Now let us consider the following function:

f(x)={|x|k,|x|1,0,|x|>1,f(x)=\begin{cases}|x|^{-k},\quad|x|\leq 1,\\ 0,\qquad\,\,\,\,\,|x|>1,\end{cases} (5.1)

with k<nk<n. Hence fL1(n)f\in L^{1}(\mathbb{R}^{n}).

Lemma 5.2.

Let f=f(x)f=f(x) be defined as above. If k<nk<n, then there exists C=C(n,k)>0C=C(n,k)>0 such that

nf(x)ϕτ(x)𝑑xCτn2k2.\int_{\mathbb{R}^{n}}f(x)\phi_{\tau}(x)dx\geq C\tau^{\frac{n}{2}-\frac{k}{2}}.
Proof.

We have

nf(x)ϕ(xτ12)𝑑x=|x|τ12|x|kϕ(xτ12)𝑑x=τn2k2|x|1|x|kϕ(x)𝑑x=Cτn2k2.\int_{\mathbb{R}^{n}}f(x)\phi\left(\frac{x}{\tau^{\frac{1}{2}}}\right)dx=\int_{|x|\leq\tau^{\frac{1}{2}}}|x|^{-k}\phi\left(\frac{x}{\tau^{\frac{1}{2}}}\right)dx=\tau^{\frac{n}{2}-\frac{k}{2}}\int_{|x|\leq 1}|x|^{-k}\phi(x)dx=C\tau^{\frac{n}{2}-\frac{k}{2}}.

Proof of Theorem 2.4..

Let f=f(x)f=f(x) given by (5.1). In order to conclude that fL1(n)Lm(n)f\in L^{1}(\mathbb{R}^{n})\cap L^{m}(\mathbb{R}^{n}) we shall assume

k<nm.k<\frac{n}{m}.

Suppose now that there exist a weak solution u(t,x)Llocα([0,T)×n)u(t,x)\in L^{\alpha}_{\text{loc}}([0,T)\times\mathbb{R}^{n}) to (1.5) with u00u_{0}\equiv 0 and u1(x)=λf(x)u_{1}(x)=\lambda f(x) for some λ0\lambda\geq 0. Combining Lemma 5.1 with Lemma 5.2 we get

Cλτn2k2+I(τ)Cτ2+(1+n2)1αI(τ)1α.C\lambda\tau^{\frac{n}{2}-\frac{k}{2}}+I(\tau)\leq C\tau^{-2+\left(1+\frac{n}{2}\right)\frac{1}{\alpha^{\prime}}}I(\tau)^{\frac{1}{\alpha}}.

Young’s inequality for products allow us to conclude

λCτ2α+1+k2.\lambda\leq C\tau^{-2\alpha^{\prime}+1+\frac{k}{2}}.

Therefore, if

2α+1+k2>0k>2α+1α1,-2\alpha^{\prime}+1+\frac{k}{2}>0\iff k>2\frac{\alpha+1}{\alpha-1},

then we conclude that λ=0\lambda=0 letting τ0+\tau\to 0^{+}. To have a non-empty range for kk we then need n>2mn>2m and

n2m>α+1α1α>n+2mn2m.\frac{n}{2m}>\frac{\alpha+1}{\alpha-1}\iff\alpha>\frac{n+2m}{n-2m}.

Appendix

A Auxiliary results

To handle with LpLqL^{p}-L^{q} estimates we state the important Riesz-Thorin interpolation theorem.

Theorem A.1 (Riesz-Thorin interpolation theorem, see Theorem 24.5.1 in [12]).

Let 1p0,p1,q0,q11\leq p_{0},p_{1},q_{0},q_{1}\leq\infty. If TT is a linear continuous operator in the space (Lp0Lq0)(Lp1Lq1)\mathcal{L}\big{(}L^{p_{0}}\mapsto L^{q_{0}}\big{)}\cap\mathcal{L}\big{(}L^{p_{1}}\mapsto L^{q_{1}}\big{)}, then

T(LpθLqθ)T\in\mathcal{L}\big{(}L^{p_{\theta}}\mapsto L^{q_{\theta}}\big{)}

for any θ(0,1)\theta\in(0,1), where

1pθ=1θp0+θp1and1qθ=1θq0+θq1.\frac{1}{p_{\theta}}=\frac{1-\theta}{p_{0}}+\frac{\theta}{p_{1}}\quad\text{and}\quad\frac{1}{q_{\theta}}=\frac{1-\theta}{q_{0}}+\frac{\theta}{q_{1}}.

Moreover, the following norm estimates hold:

T(LpθLqθ)T(Lp0Lq0)1θT(Lp1Lq1)θ.\|T\|_{\mathcal{L}(L^{p_{\theta}}\mapsto L^{q_{\theta}})}\leq\|T\|_{\mathcal{L}(L^{p_{0}}\mapsto L^{q_{0}})}^{1-\theta}\|T\|_{\mathcal{L}(L^{p_{1}}\mapsto L^{q_{1}})}^{\theta}.

A key result for multipliers in MpM_{p} with p(1,)p\in(1,\infty) is the Mikhlin-Hörmander multiplier theorem.

Theorem A.2 (Mikhlin-Hörmander multiplier theorem, see Theorem 2.5 in [17]).

Let 1<p<1<p<\infty and k=[n/2]+1k=[n/2]+1. Assume that m𝒞k(n/0)m\in\mathcal{C}^{k}(\mathbb{R}^{n}/{0}) and

|ξγm(ξ)|C|ξ||γ|,|γ|k.|\partial_{\xi}^{\gamma}m(\xi)|\leq C|\xi|^{-|\gamma|},\qquad|\gamma|\leq k.

Then, mMppm\in M_{p}^{p}.

The classical integer version of the Gagliardo–Nirenberg inequality can be stated as follows.

Proposition A.1 (Classical Gagliardo-Nirenberg inequality, see Proposition 25.5.4 in [12]).

Let j,mj,m\in\mathbb{N} with j<mj<m, and let u𝒞cm(n)u\in\mathcal{C}^{m}_{c}(\mathbb{R}^{n}), i.e. u𝒞mu\in\mathcal{C}^{m} with compact support. Let θ[j/m,1]\theta\in[j/m,1], and let p,q,rp,q,r in [1,][1,\infty] be such that

jnq=(mnr)θnp(1θ).j-\frac{n}{q}=\Big{(}m-\frac{n}{r}\Big{)}\theta-\frac{n}{p}(1-\theta).

Then,

DjuLqCn,m,j,p,r,θDmuLrθuLp1θ\displaystyle\|D^{j}u\|_{L^{q}}\leq C_{n,m,j,p,r,\theta}\|D^{m}u\|_{L^{r}}^{\theta}\ \|u\|_{L^{p}}^{1-\theta} (A.1)

provided that (mnr)j\big{(}m-\frac{n}{r}\big{)}-j\not\in\mathbb{N}, that is nr>mj\frac{n}{r}>m-j or nr\frac{n}{r}\not\in\mathbb{N}.

If (mnr)j\big{(}m-\frac{n}{r}\big{)}-j\in\mathbb{N}, then (A.1) holds provided that θ[jm,1)\theta\in[\frac{j}{m},1).

To handle with L1LL^{1}-L^{\infty} estimate, we have the following lemma.

Theorem A.3 (Bernstein’s Theorem, see Theorem 1.2 in [29]).

Let n1n\geq 1 and N>n/2N>n/2. We assume that mHNm\in H^{N}, then 1mL1\mathcal{F}^{-1}m\in L^{1} and there exists a constant CC such that

1mL1CmL21n2NDNmL2n2N.\|\mathcal{F}^{-1}m\|_{L^{1}}\leq C\|m\|_{L^{2}}^{1-\frac{n}{2N}}\|D^{N}m\|_{L^{2}}^{\frac{n}{2N}}.

The estimates provided by Theorem A.3 will be used together with the estimates for 1mL\|\mathcal{F}^{-1}m\|_{L^{\infty}}, provided by the following application of Littman’s lemma, based on stationary phase methods.

Lemma A.1 (A Littman type lemma, Lemma 2.1 in [25]).

Let us consider for ττ0\tau\geq\tau_{0}, τ0\tau_{0} is a large positive number, the oscillatory integral

ξx1(eiτω(ξ)v(ξ)).\mathcal{F}^{-1}_{\xi\rightarrow x}\big{(}e^{-i\tau\omega(\xi)}v(\xi)\big{)}.

The amplitude function v=v(ξ)v=v(\xi) is supposed to belong to 𝒞c(ξn)\mathcal{C}_{c}^{\infty}(\mathbb{R}_{\xi}^{n}) with support {ξn:|ξ|[1/2,2]}\{\xi\in\mathbb{R}^{n}:|\xi|\in[1/2,2]\}. The function ω=ω(ξ)\omega=\omega(\xi) is 𝒞\mathcal{C}^{\infty} in a neighborhood of the support of vv. Moreover, the Hessian Hω(ξ)H_{\omega}(\xi) is non-singular, that is, detHω(ξ)0\det H_{\omega}(\xi)\neq 0, on the support of vv. Then, the following LLL^{\infty}-L^{\infty} estimate holds:

ξx1(eiτω(ξ)v(ξ))L(ξn)C(1+τ)n2|α|n+1Dξαv(ξ)L(ξn),\big{\|}\mathcal{F}^{-1}_{\xi\rightarrow x}\big{(}e^{-i\tau\omega(\xi)}v(\xi)\big{)}\big{\|}_{L^{\infty}(\mathbb{R}_{\xi}^{n})}\leq C(1+\tau)^{-\frac{n}{2}}\sum_{|\alpha|\leq n+1}\big{\|}D_{\xi}^{\alpha}v(\xi)\big{\|}_{L^{\infty}(\mathbb{R}_{\xi}^{n})},

where the constant CC depends on 𝒞n+2\mathcal{C}^{n+2} norm of the phase function ω\omega, on the lower bound for |detHω(ξ)||\det H_{\omega}(\xi)| and on the diameter of the support of vv.

In [30, Proposition 2.5, Chapter 8], one can find a simple proof of Lemma A.1, from which it is easy to check that the statement remains valid whenever ω\omega and vv depend on some parameter, with a constant CC, uniform with respect to the parameter, provided that the derivatives of ω\omega and vv, as well as |detHω(t,ξ)||\det H_{\omega}(t,\xi)| can be estimated uniformly with respect to the parameter.

B Proof of Lemma 2.1

This Appendix is devoted to discuss the proof of Lemma 2.1. Before to proceed, we need to recall some facts about Bessel functions. As usual, for ν12\nu\geq-\frac{1}{2} we denote by JνJ_{\nu} the Bessel function

Jν(z)=(z2)νk=0(1)k(14z2)kk!Γ(ν+k+1).J_{\nu}(z)=\left(\frac{z}{2}\right)^{\nu}\sum_{k=0}^{\infty}(-1)^{k}\frac{\left(\frac{1}{4}z^{2}\right)^{k}}{k!\Gamma(\nu+k+1)}.

The Bessel functions satisfy the following asymptotic behaviour:

Jν(z)=cνz12cos(zνπ/2π/4)+O(z32), as z.J_{\nu}(z)=c_{\nu}z^{-\frac{1}{2}}\cos(z-\nu\pi/2-\pi/4)+O(z^{-\frac{3}{2}}),\quad\text{ as }\,z\to\infty. (B.1)

When fL1(n)f\in L^{1}(\mathbb{R}^{n}) is radial, its Fourier transform can be computed in the following way:

f^(ξ)=0f(r)rn1(r|ξ|)2n2Jn22(r|ξ|)𝑑r.\widehat{f}(\xi)=\int_{0}^{\infty}f(r)r^{n-1}(r|\xi|)^{\frac{2-n}{2}}J_{\frac{n-2}{2}}(r|\xi|)dr.

We also need the following result concerning oscillatory integrals (see Lemma 44 in [23]).

Lemma B.1.

Let h:[β,γ]h:[\beta,\gamma]\to\mathbb{R} be a concave or convex function and we consider

G=βγeih(y)𝑑y.G=\int_{\beta}^{\gamma}e^{ih(y)}dy.

If |h(y)|λ|h^{\prime}(y)|\geq\lambda on [β,γ][\beta,\gamma], then there exists C>0C>0 independent of λ\lambda such that

|G|Cλ1.|G|\leq C\lambda^{-1}.
Proof of Lemma 2.1.

We recall that

K^(t,ξ)=sin(t1+|ξ|4)1+|ξ|4=sin(t|ξ|2)|ξ|2,ξ:=1+|ξ|2.\widehat{K}(t,\xi)=\frac{\sin(t\sqrt{1+|\xi|^{4}})}{\sqrt{1+|\xi|^{4}}}=\frac{\sin(t\langle|\xi|^{2}\rangle)}{\langle|\xi|^{2}\rangle},\quad\langle\xi\rangle:=\sqrt{1+|\xi|^{2}}.

Next, we fix f𝒮(n)f\in\mathcal{S}(\mathbb{R}^{n}) such that f^𝒞c(n)\widehat{f}\in\mathcal{C}_{c}^{\infty}(\mathbb{R}^{n}), f^(ξ)\widehat{f}(\xi) is radial and f^\widehat{f} vanishes near the origin. From the representation of the Fourier transform in terms of Bessel functions we have

K(t,x)f(x)=|x|1n/20sin(tr2)r2f^(r)J(n2)/2(|x|r)rn2𝑑r.K(t,x)\ast f(x)=|x|^{1-n/2}\int_{0}^{\infty}\frac{\sin(t\langle r^{2}\rangle)}{\langle r^{2}\rangle}\widehat{f}(r)\,J_{(n-2)/2}(|x|r)\,r^{\frac{n}{2}}\,dr.

Let us fix a,ba,b such that 0<a<b0<a<b. We shall restrict our analysis in the zone xnx\in\mathbb{R}^{n} such that

ta<|x|<tb.ta<|x|<tb.

Hence, we have |x|/t1|x|/t\approx 1. From (B.1) we get

K(t,x)f(x)\displaystyle K(t,x)\ast f(x) =cn|x|12n20sin(tr2)r2f^(r)cos(|x|rn14π)rn212𝑑r\displaystyle=c_{n}|x|^{\frac{1}{2}-\frac{n}{2}}\int_{0}^{\infty}\frac{\sin(t\langle r^{2}\rangle)}{\langle r^{2}\rangle}\widehat{f}(r)\cos\left(|x|r-\frac{n-1}{4}\pi\right)r^{\frac{n}{2}-\frac{1}{2}}dr
+|x|1n20sin(tr2)r2f^(r)H(|x|r)rn2𝑑r,\displaystyle\quad+|x|^{1-\frac{n}{2}}\int_{0}^{\infty}\frac{\sin(t\langle r^{2}\rangle)}{\langle r^{2}\rangle}\widehat{f}(r)H(|x|r)r^{\frac{n}{2}}dr,

where H(|x|r)C(|x|r)32H(|x|r)\leq C(|x|r)^{-\frac{3}{2}} for |x|r|x|r large. Since f^\widehat{f} has compact support away from zero, we obtain

K(t,x)f(x)\displaystyle K(t,x)\ast f(x) =cn|x|12n20sin(tr2)r2f^(r)cos(|x|rn14π)rn212𝑑r+G(|x|)|x|1n2,\displaystyle=c_{n}|x|^{\frac{1}{2}-\frac{n}{2}}\int_{0}^{\infty}\frac{\sin(t\langle r^{2}\rangle)}{\langle r^{2}\rangle}\widehat{f}(r)\cos\left(|x|r-\frac{n-1}{4}\pi\right)r^{\frac{n}{2}-\frac{1}{2}}dr+G(|x|)|x|^{1-\frac{n}{2}},

where G(|x|)|x|1n2C|x|12n2G(|x|)|x|^{1-\frac{n}{2}}\leq C|x|^{-\frac{1}{2}-\frac{n}{2}} for all large |x||x|.

Since we are interested in at|x|btat\leq|x|\leq bt, we conclude that G(|x|)|x|1n2t12n2G(|x|)|x|^{1-\frac{n}{2}}\lesssim t^{-\frac{1}{2}-\frac{n}{2}}. So, the estimate (2.6) must come from

I~=|x|12n20sin(tr2)r2f^(r)cos(|x|rn14π)rn212𝑑r.\tilde{I}=|x|^{\frac{1}{2}-\frac{n}{2}}\int_{0}^{\infty}\frac{\sin(t\langle r^{2}\rangle)}{\langle r^{2}\rangle}\widehat{f}(r)\cos\left(|x|r-\frac{n-1}{4}\pi\right)r^{\frac{n}{2}-\frac{1}{2}}dr.

Using the formula

sin(a+b)+sin(ab)=2sin(a)cos(b),\sin(a+b)+\sin(a-b)=2\sin(a)\cos(b),

we write

sin(tr2)cos(|x|rn14π)=12sin(tr2+|x|rn14π)+12sin(tr2|x|r+n14π).\sin(t\langle r^{2}\rangle)\cos\left(|x|r-\frac{n-1}{4}\pi\right)=\frac{1}{2}\sin\left(t\langle r^{2}\rangle+|x|r-\frac{n-1}{4}\pi\right)+\frac{1}{2}\sin\left(t\langle r^{2}\rangle-|x|r+\frac{n-1}{4}\pi\right).

Hence, we have

I~=c~n|x|12n2{I++I},\tilde{I}=\tilde{c}_{n}|x|^{\frac{1}{2}-\frac{n}{2}}\{I_{+}+I_{-}\},

where

I±=0sin(tr2±(|x|rn14π))f^(r)rn212r2𝑑r.I_{\pm}=\int_{0}^{\infty}\sin\left(t\langle r^{2}\rangle\pm\left(|x|r-\frac{n-1}{4}\pi\right)\right)\frac{\widehat{f}(r)r^{\frac{n}{2}-\frac{1}{2}}}{\langle r^{2}\rangle}dr.

Setting g(r)=f^(r)rn212r2g(r)=\frac{\widehat{f}(r)r^{\frac{n}{2}-\frac{1}{2}}}{\langle r^{2}\rangle}, we have

I±(t)=0(2i)1{eih±(t,r)eih±(t,r)}g(r)𝑑r,I_{\pm}(t)=\int_{0}^{\infty}(2i)^{-1}\{e^{ih_{\pm}(t,r)}-e^{-ih_{\pm}(t,r)}\}g(r)dr,

where

h±(t,r)=tr2±{|x|rn14π}.h_{\pm}(t,r)=t\langle r^{2}\rangle\pm\{|x|r-\frac{n-1}{4}\pi\}.

Observe that

rh±(t,r)=2tr3r2±|x|,r2h±(t,r)=2tr2r23{3+r4}.\partial_{r}h_{\pm}(t,r)=\frac{2tr^{3}}{\langle r^{2}\rangle}\pm|x|,\quad\partial^{2}_{r}h_{\pm}(t,r)=\frac{2tr^{2}}{\langle r^{2}\rangle^{3}}\{3+r^{4}\}.

Since gg has compact support away from zero and at<|x|<btat<|x|<bt, we have

r2h±>0,rh+(t,r)t.\partial^{2}_{r}h_{\pm}>0,\quad\partial_{r}h_{+}(t,r)\gtrsim t.

Therefore, Lemma B.1 implies

|I+|t1|x|12n2|I+|t12n2.|I_{+}|\lesssim t^{-1}\implies|x|^{\frac{1}{2}-\frac{n}{2}}|I_{+}|\lesssim t^{-\frac{1}{2}-\frac{n}{2}}.

Hence, we must get (2.6) from |x|12n2I|x|^{\frac{1}{2}-\frac{n}{2}}I_{-}. Let us write

h(t,r)=th(r),h(r):=r2|x|tr+(n1)π4t.h_{-}(t,r)=th(r),\quad h(r):=\langle r^{2}\rangle-\frac{|x|}{t}r+\frac{(n-1)\pi}{4t}.

Since r2h(r)>0\partial^{2}_{r}h(r)>0 we obtain that rh()\partial_{r}h(\cdot) has a single zero r0r_{0} which satisfies the equation

2r03r02=|x|t.\frac{2r_{0}^{3}}{\langle r_{0}^{2}\rangle}=\frac{|x|}{t}. (B.2)

In particular, we have

|x|t=2r03r022r0r0a2.\frac{|x|}{t}=\frac{2r_{0}^{3}}{\langle r_{0}^{2}\rangle}\leq 2r_{0}\implies r_{0}\geq\frac{a}{2}. (B.3)

On the other hand, (using r0a2r_{0}\geq\frac{a}{2})

|x|t=2r03r022r0r021+r022r0a241+a24r0b24+a2a2.\frac{|x|}{t}=\frac{2r_{0}^{3}}{\langle r_{0}^{2}\rangle}\geq 2r_{0}\frac{r^{2}_{0}}{1+r_{0}^{2}}\geq 2r_{0}\frac{\frac{a^{2}}{4}}{1+\frac{a^{2}}{4}}\implies r_{0}\leq\frac{b}{2}\frac{4+a^{2}}{a^{2}}. (B.4)

That is, since |x|t1\frac{|x|}{t}\approx 1, we conclude that r0r_{0} is bounded away from zero and infinity. In this way, we can choose ff such that r0suppf^r_{0}\in\operatorname{supp\,}\widehat{f} and we can write h(r)=(rr0)h0(r)h^{\prime}(r)=(r-r_{0})h_{0}(r), with h0(r0)0h_{0}(r_{0})\neq 0. Now we split our analysis in two cases: |rr0|t13|r-r_{0}|\leq t^{-\frac{1}{3}} or |rr0|t13|r-r_{0}|\geq t^{-\frac{1}{3}} (t>1t>1).

If |rr0|t13|r-r_{0}|\geq t^{-\frac{1}{3}} then |th(r)|ct23|th^{\prime}(r)|\geq ct^{\frac{2}{3}}. From Lemma B.1 we conclude

|x|12n2||rr0|t13sin(th(r))g(r)𝑑r|Ct(1223)n2.|x|^{\frac{1}{2}-\frac{n}{2}}\left|\int_{|r-r_{0}|\geq t^{-\frac{1}{3}}}\sin(th(r))g(r)dr\right|\leq Ct^{\left(\frac{1}{2}-\frac{2}{3}\right)-\frac{n}{2}}.

Therefore, to conclude the proof it suffices to prove that

|rr0|t13sin(th(r))g(r)𝑑rt12.\int_{|r-r_{0}|\leq t^{-\frac{1}{3}}}\sin(th(r))g(r)dr\gtrsim t^{-\frac{1}{2}}.

We write Taylor’s expansion up to the order three to h(r)h(r)

T3(r)=h(r0)+h′′(r0)2(rr0)2+h(3)(r0)6(rr0)3=h(r0)+α(rr0)2+β(rr0)3,T_{3}(r)=h(r_{0})+\frac{h^{\prime\prime}(r_{0})}{2}(r-r_{0})^{2}+\frac{h^{(3)}(r_{0})}{6}(r-r_{0})^{3}=h(r_{0})+\alpha(r-r_{0})^{2}+\beta(r-r_{0})^{3},

where α>0\alpha>0. Thus, we have

|rr0|t13sin(th(r))g(r)𝑑r\displaystyle\int_{|r-r_{0}|\leq t^{-\frac{1}{3}}}\sin(th(r))g(r)dr =|rr0|t13(sin(th(r))sin(tT3(r)))g(r)𝑑r+|rr0|t13sin(T3(r))g(r)𝑑r\displaystyle=\int_{|r-r_{0}|\leq t^{-\frac{1}{3}}}\big{(}\sin(th(r))-\sin(tT_{3}(r))\big{)}g(r)dr+\int_{|r-r_{0}|\leq t^{-\frac{1}{3}}}\sin(T_{3}(r))g(r)dr
=|rr0|t13sin(T3(r))g(r)𝑑r+O(t23).\displaystyle=\int_{|r-r_{0}|\leq t^{-\frac{1}{3}}}\sin(T_{3}(r))g(r)dr+O(t^{-\frac{2}{3}}).

Indeed, using the Lagrange remainder for some r¯(r0t13,r0+t13)\overline{r}\in(r_{0}-t^{-\frac{1}{3}},r_{0}+t^{-\frac{1}{3}}) we get

|sin(th(r))sin(tT3(r))|t|h(r)T3(r)|t|h(4)(r¯)|4!(rr0)4t|rr0|4.\big{|}\sin(th(r))-\sin(tT_{3}(r))\big{|}\leq t|h(r)-T_{3}(r)|\lesssim t\frac{|h^{(4)}(\overline{r})|}{4!}(r-r_{0})^{4}\lesssim t|r-r_{0}|^{4}.

Hence, we have

||rr0|t13(sin(th(r))sin(tT3(r)))g(r)𝑑r|t|rr0|t13(rr0)4𝑑rt23.\left|\int_{|r-r_{0}|\leq t^{-\frac{1}{3}}}\big{(}\sin(th(r))-\sin(tT_{3}(r))\big{)}g(r)dr\right|\lesssim t\int_{|r-r_{0}|\leq t^{-\frac{1}{3}}}(r-r_{0})^{4}dr\lesssim t^{-\frac{2}{3}}.

Applying a similar idea to the function g(r)g(r), we conclude that it suffices to prove

g(r0)|rr0|t13sin(T3(r))𝑑r=g(r0)t13t13sin(th(r0)+tαs2+tβs3)𝑑st12.g(r_{0})\int_{|r-r_{0}|\leq t^{-\frac{1}{3}}}\sin(T_{3}(r))dr=g(r_{0})\int_{-t^{-\frac{1}{3}}}^{t^{-\frac{1}{3}}}\sin\big{(}th(r_{0})+t\alpha s^{2}+t\beta s^{3}\big{)}ds\gtrsim t^{-\frac{1}{2}}.

Since α>0\alpha>0 we obtain that the function F(s)=sα+βsF(s)=s\sqrt{\alpha+\beta s} is well-defined for small values of ss. Moreover, again assuming ss small, we have

F(s)=α+βssβ2α+βsα>0.F^{\prime}(s)=\sqrt{\alpha+\beta s}-\frac{s\beta}{2\sqrt{\alpha+\beta s}}\gtrsim\sqrt{\alpha}>0.

So, FF defines a diffeomorphism near the origin: F:(δ,δ)(F(δ),F(δ))F:(-\delta,\delta)\to(F(-\delta),F(\delta)). Changing the variables we get

t13t13sin(th(r0)+tαs2+tβs3)𝑑s\displaystyle\int_{-t^{-\frac{1}{3}}}^{t^{-\frac{1}{3}}}\sin\big{(}th(r_{0})+t\alpha s^{2}+t\beta s^{3}\big{)}ds =F(t13)F(t13)sin(th(r0)+ts2)|(F1)(s)|𝑑s\displaystyle=\int_{F(-t^{-\frac{1}{3}})}^{F(t^{-\frac{1}{3}})}\sin\big{(}th(r_{0})+ts^{2}\big{)}|(F^{-1})^{\prime}(s)|ds
dt13dt13sin(th(r0)+ts2)|(F1)(s)|𝑑s,\displaystyle\geq\int_{-dt^{-\frac{1}{3}}}^{dt^{-\frac{1}{3}}}\sin\big{(}th(r_{0})+ts^{2}\big{)}|(F^{-1})^{\prime}(s)|ds,

where d=(21α)12d=(2^{-1}\alpha)^{\frac{1}{2}}. Applying Taylor’s formula to |(F1)(s)||(F^{-1})^{\prime}(s)| we see that it is enough to prove that

dt13dt13sin(th(r0)+ts2)𝑑st12.\displaystyle\int_{-dt^{-\frac{1}{3}}}^{dt^{-\frac{1}{3}}}\sin\big{(}th(r_{0})+ts^{2}\big{)}ds\gtrsim t^{-\frac{1}{2}}. (B.5)

Using Lemma B.1 once more, the integral (B.5) may be considered over the whole real line with an error of order O(t23)O(t^{-\frac{2}{3}}). Now we note that

sin(th(r0)+ts2)𝑑s\displaystyle\int_{-\infty}^{\infty}\sin\big{(}th(r_{0})+ts^{2}\big{)}ds =t12sin(th(r0)+s2)𝑑s\displaystyle=t^{-\frac{1}{2}}\int_{-\infty}^{\infty}\sin\big{(}th(r_{0})+s^{2}\big{)}ds
=t12(2i)1ei(th(r0)+s2)ei(th(r0)+s2)ds\displaystyle=t^{-\frac{1}{2}}(2i)^{-1}\int_{-\infty}^{\infty}e^{i\left(th(r_{0})+s^{2}\right)}-e^{-i\left(th(r_{0})+s^{2}\right)}ds
=t12π2(2i)1{eith(r0)(1+i)eith(r0)(1i)}\displaystyle=t^{-\frac{1}{2}}\sqrt{\frac{\pi}{2}}(2i)^{-1}\big{\{}e^{ith(r_{0})}(1+i)-e^{-ith(r_{0})}(1-i)\big{\}}
=t12π2{cos(th(r0))+sin(th(r0))}\displaystyle=t^{-\frac{1}{2}}\sqrt{\frac{\pi}{2}}\big{\{}\cos(th(r_{0}))+\sin(th(r_{0}))\big{\}}
=t12πsin(th(r0)+π4).\displaystyle=t^{-\frac{1}{2}}\sqrt{\pi}\sin\left(th(r_{0})+\frac{\pi}{4}\right).

Moreover

th(r0)+π4=t(r02|x|tr0+(n1)π4t)+π4.th(r_{0})+\frac{\pi}{4}=t\left(\langle r_{0}^{2}\rangle-\frac{|x|}{t}r_{0}+\frac{(n-1)\pi}{4t}\right)+\frac{\pi}{4}.

Since r0r_{0} satisfies (B.2), (B.3) and (B.4), we have

r02|x|tr0=|x|tr0(2r02t2|x|21)ar0(a22b21).\langle r_{0}^{2}\rangle-\frac{|x|}{t}r_{0}=\frac{|x|}{t}r_{0}\left(2r_{0}^{2}\frac{t^{2}}{|x|^{2}}-1\right)\geq ar_{0}\left(\frac{a^{2}}{2b^{2}}-1\right).
r02|x|tr0=r02(1|x|2t212r02),\langle r_{0}^{2}\rangle-\frac{|x|}{t}r_{0}=\langle r^{2}_{0}\rangle\left(1-\frac{|x|^{2}}{t^{2}}\frac{1}{2r^{2}_{0}}\right),

and

12b2a21|x|2t212r0212a2b2a4(4+a2)2.1-\frac{2b^{2}}{a^{2}}\leq 1-\frac{|x|^{2}}{t^{2}}\frac{1}{2r^{2}_{0}}\leq 1-\frac{2a^{2}}{b^{2}}\frac{a^{4}}{(4+a^{2})^{2}}.

Hence, choosing 0<a<b0<a<b such that b2a34+a2b\leq\frac{{\sqrt{2}}a^{3}}{4+a^{2}} we have that r02|x|tr0\langle r_{0}^{2}\rangle-\frac{|x|}{t}r_{0} is bounded away from zero for all xx and tt such that at|x|tbat\leq|x|\leq tb. Hence, it is possible to take a sequence 1tk1\leq t_{k}\to\infty such that

sin(tkh(r0)+π4)12,k.\sin\left(t_{k}h(r_{0})+\frac{\pi}{4}\right)\geq\frac{1}{2},\quad k\in\mathbb{N}.

Therefore, we get

sin(tkh(r0)+tks2)𝑑stk12.\int_{-\infty}^{\infty}\sin\big{(}t_{k}h(r_{0})+t_{k}s^{2}\big{)}ds\gtrsim t_{k}^{-\frac{1}{2}}.

This concludes the proof. ∎

Acknowledgments

Alexandre A. Junior has been supported by Grant 2022/0171232022/01712-3 and Halit S. Aslan has been supported by Grant 2021/0174332021/01743-3 from São Paulo Research Foundation (FAPESP). Marcelo R. Ebert is partially supported by "Conselho Nacional de Desenvolvimento Científico e Tecnológico (CNPq)", grant number 304408/2020-4.

References

  • [1] P. Brenner, On LpLqL^{p}-L^{q} estimates for the wave-equation. Math. Z. 145 (1975), no. 3, 251–254.
  • [2] T. Cazenave, Uniform estimates for solutions of nonlinear Klein-Gordon equations. J. Funct. Anal. 60 (1985), 36–55.
  • [3] R.C. Charão, C.R. da Luz, Asymptotic properties for a semilinear plate equation in unbounded domains. J. Hyperbolic Differ. Equ. 6 (2009), 269–294.
  • [4] R.C. Charão, C.R. da Luz, R. Ikehata, New decay rates for a problem of plate dynamics with fractional damping. J. Hyperbolic Differ. Equ. 10, 3 (2013), 563–575.
  • [5] R.C. Charão, J.L. Horbach, R. Ikehata, Optimal decay rates and asymptotic profile for the plate equation with structural damping. J. Math. Anal. Appl. 440 (2016), 529–560.
  • [6] P.G. Ciarlet, A justification of the von Kármán equations. Arch. Rational Mech. Anal. 73 (1980), 349–389.
  • [7] E. Cordero, D. Zucco, Strichartz estimates for the vibrating plate equation. J. Evol. Equ. 11 (2011), 827–845.
  • [8] E. Cordero, D. Zucco, The Cauchy problem for the vibrating plate equation in modulation spaces. J. Pseudo-Differ. Oper. Appl. 2 (2011), 343–354.
  • [9] M. D’Abbicco, M.R. Ebert. A new phenomenon in the critical exponent for structurally damped semi-linear evolution equations. Nonlinear Anal. 149 (2017) 1–40.
  • [10] M. D’Abbicco, M.R. Ebert, LpLqL^{p}-L^{q} estimates for a parameter-dependent multiplier with oscillatory and diffusive components. J. Math. Anal. Appl. 504 (2021).
  • [11] M. D’Abbicco, M.R. Ebert, The critical exponent for semilinear-evolution equations with a strong non-effective damping. Nonlinear Anal. 215 (2022).
  • [12] M.R. Ebert, M. Reissig, Methods for partial differential equations. Qualitative properties of solutions, phase space analysis, semilinear models. Birkhäuser/Springer, Cham, 2018.
  • [13] M.R. Ebert, L.M. Lourenço, The critical exponent for evolution models with power non-linearity, in: Trends in Mathematics, New Tools for Nonlinear PDEs and Applications, Birkhäuser Basel, 2019, pp. 153–177.
  • [14] M.R. Ebert, C.R. da Luz, M.F. Palma, The influence of data regularity in the critical exponent for a class of semilinear evolution equations. Nonlinear Differ. Equ. Appl. 27, 44 (2020).
  • [15] Z. Guo, L. Peng, B. Wang, Decay estimates for a class of wave equations. J. Funct. Anal. 254 (2008), 1642–1660.
  • [16] E. Hebey, B. Pausader, An introduction to fourth order nonlinear wave equations.
    http://hebey.u-cergy.fr/HebPausSurvey.pdf
  • [17] L. Hörmander, Estimates for translation invariant operators in LpL^{p} spaces. Acta Math. 104 (1960), 93–140.
  • [18] M. Ikeda, T. Inui, A remark on non-existence results for the semi-linear damped Klein-Gordon equations. RIMS Kôkyûroku Bessatsu B56 (2016), 11–30.
  • [19] G. Karch, Selfsimilar profiles in large time asymptotics of solutions to damped wave equations. Studia Math. 143 (2000) 175–197.
  • [20] M. Keel, T. Tao, Small data blow-up for semi-linear Klein-Gordon equations. Amer. J. Math. 121 (1997), 629–669.
  • [21] I. Lasiecka, Uniform stability of a full von Kármán system with nonlinear boundary feedback. SIAM J. Control Optim. 36 (1998), 1376–1422.
  • [22] S. Levandosky, Decay estimates for the fourth-order wave equations. J. Differential Equations 143 (1998), 360–413.
  • [23] B. Marshall, W. Strauss, S. Wainger, LpLqL^{p}-L^{q} estimates for the Klein-Gordon equation. J. Math. Pures Appl. 59 (1980) 417–440.
  • [24] A. Miyachi, On some singular Fourier multipliers. J. Fac. Sci. Univ. Tokyo Sect. IA Math. 28 (1981), no. 2, 267–315.
  • [25] H. Pecher, LpL^{p}-Abschätzungen und klassische Lösungen für nichtlineare Wellengleichungen, I. Math. Z. 150 (1976), 159–183.
  • [26] J.C. Peral, LpL^{p} estimates for the wave equation. J. Functional Analysis 36 (1980), no. 1, 114–145.
  • [27] D.T. Pham, M. Kainane, M. Reissig, Global existence for semi-linear structurally damped σ\sigma-evolution models. J. Math. Anal. Appl. 431 (2015) 569–596.
  • [28] J.P. Puel, M. Tucsnak, Global existence for full von Kármán system. Appl. Math. Optim. 34 (1996), 139–160.
  • [29] S. Sjöstrand, On the Riesz means of the solution of the Shrödinger equation. Ann. Sc. Norm. Super. Pisa, Cl. Sci. (3) 24(2) (1970), 331–348.
  • [30] E.M. Stein, R. Shakarchi, Functional Analysis: Introduction to further topics in analysis, vol. 4, Princeton University Press, 2011.
  • [31] R.S. Strichartz, Convolutions with kernels having singularities on a sphere. Trans. Amer. Math. Soc. 148 (1970), 461–471.
  • [32] Y. Sugitani, S. Kawashima, Decay estimates of solutions to a semi-linear dissipative plate equation. J. Hyperbolic Differ. Equ. 7 (2010) 471–501.