In the first part of the paper we continue the study of solutions to Schrödinger equations with a time singularity in the dispersive relation and in the periodic setting. In the second we show that if the Schrödinger operator involves a Laplace operator with variable coefficients with a particular dependence on the space variables, then one can prove Strichartz estimates at the same regularity as that needed for constant coefficients. Our work presents a two dimensional analysis, but we expect that with the obvious adjustments similar results are available in higher dimensions.
Citation: Serena Federico, Gigliola Staffilani. Sharp Strichartz estimates for some variable coefficient Schrödinger operators on R×T2[J]. Mathematics in Engineering, 2022, 4(4): 1-23. doi: 10.3934/mine.2022033
[1] | José Antonio Vélez-Pérez, Panayotis Panayotaros . Wannier functions and discrete NLS equations for nematicons. Mathematics in Engineering, 2019, 1(2): 309-326. doi: 10.3934/mine.2019.2.309 |
[2] | Luca Franzoi, Riccardo Montalto . Time almost-periodic solutions of the incompressible Euler equations. Mathematics in Engineering, 2024, 6(3): 394-406. doi: 10.3934/mine.2024016 |
[3] | Federico Bernini, Simone Secchi . Existence of solutions for a perturbed problem with logarithmic potential in R2. Mathematics in Engineering, 2020, 2(3): 438-458. doi: 10.3934/mine.2020020 |
[4] | Roberto Feola, Felice Iandoli, Federico Murgante . Long-time stability of the quantum hydrodynamic system on irrational tori. Mathematics in Engineering, 2022, 4(3): 1-24. doi: 10.3934/mine.2022023 |
[5] | María Ángeles García-Ferrero, Angkana Rüland . Strong unique continuation for the higher order fractional Laplacian. Mathematics in Engineering, 2019, 1(4): 715-774. doi: 10.3934/mine.2019.4.715 |
[6] | Alessandro Michelangeli, Raffaele Scandone . On real resonances for three-dimensional Schrödinger operators with point interactions. Mathematics in Engineering, 2021, 3(2): 1-14. doi: 10.3934/mine.2021017 |
[7] | Prashanta Garain, Kaj Nyström . On regularity and existence of weak solutions to nonlinear Kolmogorov-Fokker-Planck type equations with rough coefficients. Mathematics in Engineering, 2023, 5(2): 1-37. doi: 10.3934/mine.2023043 |
[8] | Silvia Cingolani, Marco Gallo, Kazunaga Tanaka . On fractional Schrödinger equations with Hartree type nonlinearities. Mathematics in Engineering, 2022, 4(6): 1-33. doi: 10.3934/mine.2022056 |
[9] | Marco Bramanti, Sergio Polidoro . Fundamental solutions for Kolmogorov-Fokker-Planck operators with time-depending measurable coefficients. Mathematics in Engineering, 2020, 2(4): 734-771. doi: 10.3934/mine.2020035 |
[10] | Chiara Gavioli, Pavel Krejčí . Deformable porous media with degenerate hysteresis in gravity field. Mathematics in Engineering, 2025, 7(1): 35-60. doi: 10.3934/mine.2025003 |
In the first part of the paper we continue the study of solutions to Schrödinger equations with a time singularity in the dispersive relation and in the periodic setting. In the second we show that if the Schrödinger operator involves a Laplace operator with variable coefficients with a particular dependence on the space variables, then one can prove Strichartz estimates at the same regularity as that needed for constant coefficients. Our work presents a two dimensional analysis, but we expect that with the obvious adjustments similar results are available in higher dimensions.
This paper is devoted to the analysis of Schrödinger operators on the two dimensional torus T2 with variable coefficients depending either on time or on space.
In the time-dependent case, being, in particular, a time-degenerate case, a time-weighted version of Bourgain's sharp L4-Strichartz estimate will be derived and applied to the corresponding semilinear initial value problem (IVP). As for the (special) space-dependent nondegenerate case, the sharp local well-posedness in Hε(T2), for any ε>0, of the corresponding nonlinear cubic initial value problem will be shown.
Before describing in detail the operators under consideration and the problems addressed in this work, we recall that the celebrated sharp L4-Strichartz estimate on T2 proved by Bourgain [1] for the rational torus and by Bourgain-Demeter [3] in the general case states that:
‖eitΔxu0‖L4t([0,1]×T2)≤Cϵ‖u0‖Hϵ(T2),ε>0. | (1.1) |
Here and in the rest of the paper we shall refer to (1.1) as Bourgain's sharp Strichartz estimate on T2.
Let us start by introducing Schrödinger equations with time variable coefficients, a topic that has attracted the interest of mathematicians and physicists in the last decades. For instance, equations of the form
i∂tu+b(t)Δxu+h(t)u|u|2+v(t,x)u=0, | (1.2) |
under some assumptions on the time dependent functions, have been recently studied in the context of Bose-Einstein condensations and nonlinear optics in the Euclidian setting. Exact one and two soliton solutions for (1.2) have been derived in [14], whereas similar related problems have been studied, for example, in [5,19,20] (see also references therein). Interestingly it was shown in [20] that the nature of the soliton solution (canonical soliton or deformed canonical soliton) to Eq (1.2) is influenced by the choice of the time dependent coefficients b,h and v.
Time-degenerate equations of the form (1.2) (with b(0)=0), still in the Euclidean setting, have been investigated in connection with other mathematical problems. Local well-posedness results for the homogeneous IVP associated with time-degenerate equations of the form (1.2) (with h=v=0) have been derived in [6]. The validity of smoothing estimates and the local well-posedness of the nonlinear IVP for time-degenerate equations of the form (1.2) was proved by the authors in [9], while Strichartz estimates for the same kind of time-degenerate equations (with v=0) have been derived in [8] by the first author and Ruzhansky. All the results listed hold for space variable in Rd. In the present work instead we are interested on investigating Strichartz estimates and well-posedness in the periodic setting.
Let us now go back to the Bourgain's sharp Strichartz estimate (1.1). We would like to stress that, in spite of the fact that a deep analysis of Strichartz estimates on the torus and on more general manifolds has been carried out in the last years, as we will recall below, Bourgain's sharp result remains unproved for variable coefficient Schrödinger operators on T2, even in the simplest case when the Laplace operator is replaced by an elliptic operator with coefficients which are smooth and almost constant.
In the Euclidean setting Strichartz estimates for non-degenerate space-variable coefficient Schrödinger operators were proved by the second author and Tataru in [18], and, in the non-elliptic case, by Salort in [17]. Smoothing estimates for non-degenerate space-variable coefficient Schrödinger operators have been considered by Kenig, Ponce, Rolvung and Vega, see, for instance, [11] and references therein. Smoothing and Strichartz estimates for asymptotically flat Schrödinger operators were derived by Marzuola, Metcalfe and Tataru in [15].
In the manifolds setting Strichartz estimates on general compact manifolds have been proved by Burq, Gérard and Tzvetkov in [4] and by Hani in [10], and by Mizutani and Tzvetkov in [16] in the non-elliptic case. Specifically, Theorem 1 in [4] states the following:
Let (M,g) be a Riemannian compact manifold of dimension d≥1 and Δ be the Laplace Beltrami operator on M. Given p,q satisfying the scaling condition 2/p+d/q=d/2, and p≥2, q<∞, the solution v of
i∂tv(t,x)+Δv(t,x)=0,v(0,x)=v0(x), |
satisfies for any finite time interval I,
‖v(t)‖Lp(I)Lq(M)≤C(I)‖v0‖H1p(M). | (1.3) |
It is clear that Boungain's sharp result on T2 is not covered by (1.3), since, for p=q=2 and M=T2, (1.3) gives
‖v(t)‖L4(I)L4(T2)≤C(I)‖v0‖H14(T2). |
However, still in [4], it is shown that estimate (1.3) is actually sharp when M is the sphere Sd with d≥3.
The observations and results mentioned above motivated our analysis of the following linear Schrödinger equations with variable coefficients in T2:
i∂tu+g′(t)Δxu=0, | (1.4) |
and
i∂tu+a1(x1)∂2x1u+a2(x2)∂2x2u=0, | (1.5) |
where g′, representing the derivative of a strictly monotone function g∈C1(R), with g(0)=0, is such that g′(0)=0, while ai∈C∞(T), for i=1,2, are strictly positive functions*.
* Here we assume smoothness for a1 and a2, but much less regularity is required for our analysis.
For the time-degenerate Eq (1.4) (which is now considered in a non-Euclidean setting), a suitable weighted version of Bourgain's sharp-Strichartz estimate on T2 will be obtained. As a consequence of the aforementioned inequality, a local well-posedness result for the semilinear IVP associated with (1.4) will be obtained.
As for Eq (1.5), by exploiting a change of variables in combination with the use of a gauge transform, we will be able to apply Bourgain's sharp-Strichartz estimate to obtain the local well-posedness of the corresponding cubic IVP in Hε(T2),ε>0 (the same result as the one we observe in the non-degenerate constant coefficients case).
Although the study of (1.4) and (1.5) is interesting on its own term, the analysis of these operators shows that sharp Strichartz estimates on T2 can be obtained even when some variable (possibly degenerate in time) coefficients are present, and as a consequence also sharp local well-posedness results in the same settings hold true.
Indeed, while the result in [4] allows to prove local well-posedness in Hs(T2), with s>1/4, of the cubic IVP associated with the constant coefficient Schrödinger operator on T2, here we manage to combine different techniques to obtain sharp local well-posedness in Hs, s>0, when some variable coefficients are present.
For completeness we list below the main results of this paper.
Let g∈C∞(Rt) be a strictly monotone function such that g(0)=g′(0)=0, let f be a smooth function such that f∈H1,bg(R) for b∈(1/2,1), where H1,bg(R) is as in Definition 2.5, and let ˜Xs,bg(R×T2) be the Banach space as in Definition 2.2. Then the following theorems hold.
Theorem 1.1. Let s>0 and b∈(1/2,1). Then, for every u0∈Hs(T2), there exists a unique solution of the IVP
{i∂tu+g′(t)Δxu=g′(t)|u|2u,u(0,x)=u0(x), | (1.6) |
in the time interval [−T,T] for a suitable time T=T(‖u0‖Hs). Moreover the solution u satisfies
u∈C([−T,T];Hs) |
and, for I closed neighborhood of [−T,T], and χI a smooth cutoff function such that χI≡1 on [−T,T], we have that there exists b∈(1/2,1) such that
χIu∈˜Xs,bg(R×T2). |
Theorem 1.2. Let s>0 and b∈(1/2,1). Then, for every u0∈Hs(T2), there exists a unique solution of the IVP
{i∂tu+g′(t)Δxu=f(t)|u|2u,u(0,x)=u0(x), | (1.7) |
in the time interval [−T,T] for a suitable time T=T(‖u0‖Hs). Moreover the solution u satisfies
u∈C([−T,T];Hs) |
and, for a closed neighborhood I of [−T,T], we have that there exists b∈(1/2,1) such that
χIu∈˜Xs,bg(R×T2) |
with χI being a smooth cutoff function such that χI≡1 on [−T,T].
Let a1,a2∈C∞(T) be two real valued strictly positive functions and Xs,bΦ,˜α(R×T2) the Banach space in Definition 2.4. Then the following local well-posedness result holds.
Theorem 1.3. Let s>0 and b∈(1/2,1). Then, for every u0∈Hs(T2), there exists a unique solution of the IVP
{i∂tu+a1(x1)∂2x1u+a2(x2)∂2x2u=u|u|2,u(0,x)=u0(x), | (1.8) |
in the time interval [−T,T] for a suitable time T=T(‖u0‖Hs). Moreover the solution u satisfies
u∈C([−T,T];Hs) |
and, for a closed neighborhood I of [−T,T], we have that there exists b∈(1/2,1) such that
χIu∈Xs,bΦ(R×T2) |
with χI being a smooth cutoff function such that χI≡1 on [−T,T].
Remark 1.4. Applying the strategies employed in the results above one can also derive the local well-posedness of the IVP
{i∂tu+g′(t)(a1(x1)∂2x1u+a2(x2)∂2x2u)=f(t)u|u|2,u(0,x)=u0(x), |
with solution in a suitable Xs,bg,Φ space (see Definition 2.4).
Under the same assumptions on the functions g and f the following results hold.
Theorem 1.5. Let s>0 and b∈(1/2,1). Then, for every u0∈Hs(T), there exists a unique solution of the IVP
{i∂tu+g′(t)Δxu=g′(t)|u|4u,u(0,x)=u0(x), | (1.9) |
in the time interval [−T,T] for a suitable time T=T(‖u0‖Hs). Moreover the solution u satisfies
u∈C([−T,T];Hs) |
and, for I closed neighborhood of [−T,T], and χI a smooth cutoff function such that χI≡1 on [−T,T], we have that there exists b∈(1/2,1) such that
χIu∈˜Xs,bg(R×T). |
Theorem 1.6. Let s>0 and b∈(1/2,1). Then, for every u0∈Hs(T), there exists a unique solution of the IVP
{i∂tu+g′(t)Δxu=f(t)|u|4u,u(0,x)=u0(x), | (1.10) |
in the time interval [−T,T] for a suitable time T=T(‖u0‖Hs). Moreover the solution u satisfies
u∈C([−T,T];Hs) |
and, for a closed neighborhood I of [−T,T], we have that there exists b∈(1/2,1) such that
χIu∈˜Xs,bg(R×T) |
with χI being a smooth cutoff function such that χI≡1 on [−T,T].
Let a∈C∞(T) be a real valued strictly positive function and Xs,bΦ,˜α(R×T) as in Definition 2.4. Then the following local well-posedness result holds.
Theorem 1.7. Let s>0 and b∈(1/2,1). Then, for every u0∈Hs(T), there exists a unique solution of the IVP
{i∂tu+a(x)Δxu=u|u|4,u(0,x)=u0(x), | (1.11) |
in the time interval [−T,T] for a suitable time T=T(‖u0‖Hs). Moreover the solution u satisfies
u∈C([−T,T];Hs) |
and, for a closed neighborhood I of [−T,T], we have that there exists b∈(1/2,1) such that
χIu∈Xs,bΦ,˜α(R×T) |
with χI being a smooth cutoff function such that χI≡1 on [−T,T].
Remark 1.8. Applying the strategies employed in the results above one can also derive the local well-posedness of the IVP
{i∂tu+g′(t)a(x)Δxu=f(t)u|u|4,u(0,x)=u0(x), |
with solution in a suitable ˜Xs,bg,Φ,˜α space (see Definition 2.4).
Below we will give detailed proofs of the two dimensional case (x∈T2), that have straightforward applications in the one dimensional case (x∈T). However we expect equivalent results to be true in higher dimensions after applying suitable adjustments in our assumptions and arguments.
Let us remark that in the sequel we will also give very precise proofs of some standard multilinear estimates contained, for instance, in [1,12]. Even if some of these results are well-known and standard, detailed proofs allow to measure in a more precise way some key parameters appearing in the contraction argument, as, for instance, the parameter δ associated with the length of the time interval of existence, and, more importantly, some (dangerous) exponents appearing in the contraction argument. It is only with a sharp evaluation of these quantities that the contraction argument can be performed successfully and get sharp local well-posedness results.
We conclude this introduction by giving the plan of the paper. Section 2 is devoted to the definition of the Xs,bg-spaces which are central in our analysis. In Section 3 we derive the suitable weighted formulation of Strichartz and multilinear estimates via the use of Xs,bg-spaces. This section also contains the proof of Theorem 1.1, 1.2, 1.5 and 1.6. Finally Section 4 will be devoted to the study of (1.8) and (1.11), that is, in particular, to the proof of Theorem 1.3 and 1.7.
Notation. We shall use the notation A≲B to indicate that there exists an absolute constant c>0 such that A≤cB.
In this section we shall briefly recall some tools we will be using throughout the paper.
Given a function u∈C∞(Td), the Fourier and anti-Fourier transforms of u are defined as
ˆu(k)=(Fu)(k)=∫Tde−ik⋅xu(x)dx, |
and
u(x)=(F−1ˆu)(x)=∑k∈Zdeik⋅xˆu(k), |
and, by Plancherel's Theorem, we have the identity
‖u‖L2(Td)=‖ˆu‖ℓ2(Zd). |
Similarly, given a space-time dependent function v(t,x), with (t,x)∈R×Td, we can write v(t,x) by means of the Fourier representation formula as
v(t,x)=∑k∈Zd∫Re−i(k⋅x+tτ)ˆv(k,τ)dτ, |
where here ˆv(k,τ) (and in the rest of the paper) is the space-time Fourier transform.
As Bourgain did in his pioneering work in the periodic setting [1], also in our case we will make use of the Xs,b-spaces recalled below.
Definition 2.1. (Xs,b Spaces). Let X be the space of functions u on R×Td such that
(i) u:R×Td→C;
(ii) t→u(t,x) is S(R) for all x∈Td;
(iii) x→u(⋅,x) is C∞(Td).
Then, for s,b∈R, we denote by Xs,b the completion of the space X with respect to the norm
‖u‖Xs,b:=(∑k∈Zd(1+|k|)2s∫R(1−|τ−|k|2|)2b|ˆu(τ,k)|2dτ)1/2. |
We will now introduce a new notion of Xs,b space subordinate to a function g∈C∞(Rt). Let g∈C∞(Rt) be a strictly monotone function such that g(0)=0, then we define the modified Fourier and inverse Fourier transform subordinate to g as
(˜Fu)(τ):=∫Re−ig(t)τu(t)dt |
and
(˜F−1v)(t):=g′(t)∫Reig(t)τv(τ)dτ, |
where u,v∈L1(R)∩L2(R). Note that, on denoting by ˜u(τ):=(˜Fu)(τ) we have that the following properties hold
● u(t)=(˜F−1˜u)(t);
● ˜F(∂tu)(τ)=(−iτ)˜F(g′u)(τ);
● ‖˜u‖L2(Rτ)=‖1√|g′|u‖L2(Rt);
● ‖~g′u‖L2(Rτ)=‖√|g′|u‖L2(Rt);
● ‖˜F(√|g′|u)‖L2(Rτ)=‖u‖L2(Rt);
When no confusion arises we shall also use the notation ˜u(τ,k) for the space-time transform
˜u(τ,k):=∫R×Tde−i(g(t)τ+k⋅x)u(t,x)dtdx, |
namely, the modified Fourier transform in time and the standard Fourier transform in space of a function u on R×Td.
With this definition at our disposal we can now define the Xs,bg and the ˜Xs,bg spaces subordinate to g.
Definition 2.2. (Xs,bg and ˜Xs,bg spaces). Given a strictly monotone function g∈C∞(Rt), with g(0)=0, we define the space Xs,bg as the completion of the space X in Definition 2.1 with respect to the norm
‖u‖Xs,bg:=(∑k∈Zd(1+|k|)2s∫R(1+|τ−|k|2|)2b|˜u(τ,k)|2dτ)1/2. |
With the same function g we define the spaces ˜Xs,bg as
˜Xs,bg:={u∈Xs,bg;g′u∈Xs,bg}. |
Remark 2.3. Note that, by a simple change of variables, we have the following relations between Xs,bg, ˜Xs,bg and Xs,b spaces:
‖u‖Xs,bg=‖1g′∘g−1u(g−1(⋅),⋅)‖Xs,b |
and
‖u‖˜Xs,bg:=‖g′u‖Xs,bg=‖u(g−1(⋅),⋅)‖Xs,b, |
provided that g′u and u are smooth enough. Note also that
‖u‖X0,0g=‖˜u‖L2τℓ2k=‖1√|g′|u‖L2t,x. |
These simple observations will be crucial in the next section where the time-degenerate case described by Eq (1.4) will be treated. We shall also explain below how these spaces are related with the solution of the IVP associated with (1.4).
We conclude this section by defining some other spaces we will be using throughout the paper.
Definition 2.4 (Xs,bΦ, Xs,bg,Φ, Xs,bΦ,˜α, Xs,bg,Φ,˜α, and ˜Xs,bg,Φ,˜α spaces). Let Φ∈C∞(Td) and let g be as in Definition 2.2. Let also ˜α:R×Td→R×Td be such that ˜α(t,x):=(t,α(x)), where α:Td→Td is a diffeomorphism. Then we define the spaces Xs,bΦ, Xs,bg,Φ, Xs,bΦ,˜α, Xs,bg,Φ,˜α, and ˜Xs,bg,Φ,˜α as
Xs,bΦ(R×Td):={f:R×Td→C;eΦf∈Xs,b(R×Td)},Xs,bg,Φ(R×Td):={f:R×Td→C;eΦf∈Xs,bg(R×Td)},˜Xs,bg,Φ(R×Td):={f:R×Td→C;eΦf∈˜Xs,bg(R×Td)},Xs,bΦ,˜α(R×Td):={f:R×Td→C;(eΦf)∘˜α∈Xs,b(R×Td)},Xs,bg,Φ,˜α(R×Td):={f:R×Td→C;(eΦf)∘˜α∈Xs,bg(R×Td)},˜Xs,bg,Φ,˜α(R×Td):={f:R×Td→C;(eΦf)∘˜α∈˜Xs,bg(R×Td)}. |
Definition 2.5 (Hp,b and Hp,bg spaces). Let p∈[1,∞) and b∈R, then we define the spaces Hp,b(R) and Hp,bg(R) as
Hp,b(R):={f∈Lp(R);ˆf,^Dbf∈Lp(R)} |
equipped with the norm
‖f‖pHp,b:=∫R⟨τ⟩pb|ˆf(τ)|pdτ, |
with ⟨τ⟩:=(1+|τ|2)1/2, and
Hp,bg(R):={f∈Lp(R);‖f‖Hp,bg<∞}, |
where ‖f‖pHp,bg:=∫R⟨τ⟩pb|˜f(τ)|pdτ.
Here we focus on the analysis of the two-dimensional case, that is when x∈T2. We will use the square torus to conduct our calculations, so that the symbol of the Laplacian is simply −|k|2. The argument will be the same for any other torus. The one-dimensional case follows similarly and is described at the end of this section.
We consider the time-degenerate Schrödinger equation on R×T2
i∂tu−g′(t)Δxu=0, |
where g′ is the derivative of a strictly monotone function g∈C∞(R), with g(0)=0, and such that g′(0)=0. Additionally, g is supposed to have power growth, that is there exists α>0 such that |g(j)(t)|≲(1+|t|)α−j.
It is easy to see that the solution of the IVP associated with the equation under consideration and with initial datum u(0,x)=u0 is given by
u(t,x)=S(t)u0:=∑k∈Z2e−ig(t)|k|2ˆu0(k), |
where S(t):=S(t,0):=eig(t)Δx represents the so called solution operator giving the solution at time t starting at time 0. More generally, given a space-dependent function φ, we have
S(t,s)φ:=∑k∈Z2e−i(g(t)−g(s))|k|2ˆφ(k), |
where
● S(t,t)=I∀t∈R;
● S(t,s)=S(t,r)S(r,s),∀r,s,t∈R;
● S(t,s)Δx=ΔxS(t,s).
Moreover by Duhamel's principle the solution of the inhomogeneous IVP
{i∂tu−g′(t)Δxu=f(t,x)u(0,x)=u0(x) |
is given by
˜S(t)u0:=S(t)u0+∫t0S(t,s)f(s)ds. |
Now, considering ˜F as the time-Fourier transform subordinate to g and applying ˜Ft→τFx→k to Eq (1.4), we obtain that
˜Ft→τFx→k(i∂tu−g′(t)Δxu)=0 |
which is equivalent to
(−τ+|k|2)~g′u(τ,k)=0, |
meaning that the modified Fourier transform of the product of g′ and the solution u is supported on the paraboloid τ=|k|2.
Under the previous assumptions on the function g we now prove the frequency localized (time-)weighted Strichartz estimate from which the full (time-)weighted Strichartz estimate will follow. Afterwards, we shall also give detailed proofs of some classical multilinear estimates holding in the constant coefficients case. We will then translate these estimates in their suitable (time-)weighted version to be used in the time-degenerate case. Some of the standard multilinear estimates we prove here will be used in their original form in the analyis of the space-variable coefficients case studied in Section 4.
For simplicity we assume that g is strictly increasing, where, recall, g(0)=g′(0)=0. This assumption is useful when we make certain change of variables. Note that the vanishing of g′ is essential in order to have a degeneracy in the time-dependent Schrödinger operator.
Proposition 3.1. Let ϕN∈L2(T2) be such that suppˆϕN⊆B(0,N):={k∈Z2;|k|≤N}, and let I⊂R be a finite interval centered at zero. Then, for any ϵ>0, we have
‖g′(t)1/4S(t)ϕN‖L4(I×T2)≲Nϵ‖ϕN‖L2(T2). | (3.1) |
Proof. To prove the result we apply a change of variable in time and get
‖g′(t)1/4S(t)ϕN‖L4(I×T2)=(∫I×Tdg′(t)|eig(t)ΔϕN(x)|4dxdt)1/4 |
=t′=g(t)(∫I′×Td|eit′ΔϕN(x)|4dxdt′)1/4=‖eit′ΔϕN‖L4(I′×T2) |
≲Nϵ‖ϕN‖L2(T2), |
where the last inequality follows from the standard case when g′(t)=1 (see [1] for the rational case and [3] for the general case). This shows (3.1) and concludes the proof.
Proposition 3.2. Let ϕN1,ϕN2∈L2(T2), N1>>N2, such that suppˆϕNi⊆B(0,Ni):={k∈Z2;|k|≤Ni}, and let I⊂R be a finite interval centered at zero. Then, for any ε>0,
‖g′(t)1/2(S(t)ϕN1)(S(t)ϕN2)‖L2(I×T2)≲min(N1,N2)ϵ‖ϕN1‖L2(T2)‖ϕN2‖L2(T2). | (3.2) |
Proof. The proof is straightforward and follows by simply observing that
‖g′(t)1/2(S(t)ϕN1)(S(t)ϕN2)‖L2(I×T2)=‖(eitΔϕN1)(eitΔϕN2)‖L2(I′×T2) |
≲Nϵ2‖ϕN1‖L2(T2)‖ϕN2‖L2(T2), |
where, once more, the last inequality is due to the application of classical results (see [1] for rational tori and [7] for the general case).
A Strichartz estimate similar to the one in Proposition 3.2 in any dimension and any p≥2, with g(t)=1, was proved by Bourgain and Demeter [3], and hence, using the Littlewood-Paley decomposition one gets the full (time-)weighted Lp-Strichartz estimate on I×T2 for p≥2 below.
Theorem 3.3. Let I be a finite time interval. Then, for p≥2,
‖g′(t)1/pS(t)ϕ‖Lp(I×Td)≲‖ϕ‖L2(Td),p<2(d+2)d; |
‖g′(t)1/pS(t)ϕ‖Lp(I×Td)≲‖ϕ‖Hs(Td),s>0,p=2(d+2)d; |
‖g′(t)1/pS(t)ϕ‖Lp(I×T2)≲‖ϕ‖Hs(T2),s>d2−d+2p,p>2(d+2)d. |
We now restrict ourselves again to the case d=2.
Proposition 3.4. Let uN be a function on I×T2 such that the space-Fourier transform ˆuN(t,k) is supported in B(0,N):={k∈Z2;|k|≤N}. Then, for any s1>0 and b1>1−min{s1,1/2}2 (14<b1<12+, s1>1−2b1), we have
‖χ[0,1]g′(t)1/4u‖L4(R×T2)≲Ns1‖χ[0,1]g′(t)u‖X0,b1g. |
Proof. By applying the previous change of variables we have
‖χ[0,1]g′(t)1/4u‖L4(R×T2)=‖χ[0,g(1)]u(g−1(t),⋅)‖L4(R×T2) |
≲Ns1‖χ[0,g(1)]u(g−1(t),⋅)‖X0,b1=Ns1‖χ[0,1]g′(t)u‖X0,b1g, |
where in the last line we applied the result in the standard case g′(t)=1 in [1] and the relation between Xs,b and Xs,bg norms.
Proposition 3.5. Assume that |I′|:=|g(I)|=δ, then
‖χI(t)g′(t)S(t)u0‖Xs,bg≲δ1/2−b‖u0‖Hs,∀u0∈Hs(T2), | (3.3) |
‖g′(t)∫t0g′(s)S(t,s)w(s)ds‖Xs,bg≲‖g′(t)w‖Xs,b−1g, | (3.4) |
‖χI(t)g′(t)|u|2u‖Xs,b−1g≲‖χIg′(t)u‖2Xs,b′g‖χIg′(t)u‖Xs,bg,forb>1/2,1/4<b′<b,s>0; | (3.5) |
‖χIg′(t)u‖Xs,b′g≲δb−b′8‖g′(t)u‖Xs,bg. | (3.6) |
Proof. First observe that (3.3), (3.4), (3.5) and (3.6) hold true in the standard case when g(t)=t (see [1]). Now, by using the definition of modified Fourier transform (in time) with respect to g, we have
˜Ft→τFx→k(χIg′(t)S(t)u0)(τ,k)=∫Re−ig(t)τ−ig(t)|k|2χIg′(t)ˆu0(k)dt |
=t′=g(t)∫Re−it′(τ+|k|2)χg(I)(t′)ˆu0(k)dt′=Ft→τFx→k(χg(I)eitΔu0)(τ,k), |
which, in particular, gives
‖χI(t)g′(t)S(t)u0‖Xs,bg=‖χI(g−1(t))eitΔu0‖Xs,b |
≲|g(I)|1/2−b‖u0‖Hs, |
where in the last line we applied (3.3) in the standard case g(t)=t to conclude (3.3) in our case.
To prove (3.4) we apply the previous strategy, that is, after a change of variable we get
‖g′(t)∫t0g′(s)S(t,s)w(s)ds‖Xs,bg=‖∫t0ei(t−s′)Δw(g−1(s′))ds′‖Xs,b |
≲‖w(g−1(s′))‖Xs,b−1=‖g′(t)w‖Xs,b−1g, |
where in the last line we applied (3.4) in the standard case and the relation between Xs,b and Xs,bg spaces.
Inequality (3.5) follows again from the standard case, indeed,
‖χI(t)g′(t)|u|2u‖Xs,b−1g=‖χg(I)(t)|u(g−1(t))|2u(g−1(t))‖Xs,b−1 |
≲‖χg(I)(t)u(g−1(t))‖2Xs,b′‖χg(I)(t)u(g−1(t))‖Xs,b |
=‖χI(t)g′(t)u‖2Xs,b′g‖χI(t)g′(t)u‖Xs,bg, |
which proves (3.5).
Finally, (3.6) can be proved again by means of the standard case, so we omit the details. This concludes the proof.
More generally one has that the following trilinear estimate holds.
Proposition 3.6. Let s>0, b>1/2 and 1/4<b′<b. Let also χI be a smooth cutoff function as before. Then
‖g′(t)χIv1χIv2χIv3‖Xs,b−1g≲‖χIg′(t)v1‖Xs,bg‖χIg′(t)v2‖Xs,b′g‖χIg′(t)v3‖Xs,b′g | (3.7) |
Proof. By using the identity
‖g′(t)χIv1χIv2χIv3‖Xs,b−1g=‖χg(I)w1χg(I)w2χg(I)w3‖Xs,b−1, |
with wi(t,x)=wi(g−1(t),x), i=1,2,3, the proof follows from the standard case g(t)=t (see [1]).
We will now prove some results holding in the classical constant coefficients case from which the suitable formulation in our degenerate case will be derived.
Proposition 3.7. Let u∈Xs,b(R×T2) and let f=f(t) be such that f∈H1,b(R). Then, for all s>0 and for all b∈(1/2,1),
‖uf‖Xs,b≲‖f‖H1,b‖u‖Xs,b. | (3.8) |
Proof. Since
‖uf‖Xs,b=‖^u∗f(τ,k)⟨k⟩s⟨τ−|k|2⟩b‖L2tℓ2k |
we can prove the result by duality on L2τℓ2k. We then take v∈L2τℓ2k such that ‖v‖L2τℓ2k=1, and have
∑k∈Z2∫R|^u∗f|(τ,k)⟨k⟩s⟨τ−|k|2⟩b|v|(τ,k)dτ |
≲∑k∈Z2∫R2⟨τ1+τ2−|k|2⟩b⟨k⟩s|ˆu|(τ1,k)|ˆf|(τ2)|v|(τ1+τ2,k)dτ1dτ2 |
≲∑k∈Z2∫R2⟨τ1−|k|2⟩b⟨k⟩s|ˆu|(τ1,k)|ˆf|(τ2)|v|(τ1+τ2,k)dτ1dτ2 |
+∑k∈Z2∫R2⟨τ2⟩b⟨k⟩s|ˆu|(τ1,k)|ˆf|(τ2)|v|(τ1+τ2,k)dτ1dτ2 |
=I+II. |
For the term I we get
I=∑k∈Z2∫R2⟨τ1−|k|2⟩b⟨k⟩s|ˆu|(τ1,k)|ˆf|(τ2)|v|(τ1+τ2,k)dτ1dτ2 |
≲∫Rτ2|ˆf(τ2)|(∫Rτ1‖⟨τ1−|k|2⟩b⟨k⟩sˆu(k)‖ℓ2k‖v(τ1+τ2,k)‖ℓ2kdτ1)dτ2 |
≲‖ˆf‖L1‖u‖Xs,b‖v‖L2τℓ2k. |
Similar computations on the term II give
II=∑k∈Z2∫R2⟨τ2⟩b⟨k⟩sˆu(τ1,k)ˆf(τ2)v(τ1+τ2,k)dτ1dτ2 |
≤‖f‖H1,b‖u‖Xs,b‖v‖L2τℓ2k. |
Finally, putting together the estimates for I and II the result follows.
Proposition 3.8. Let χI be a smooth cutoff function supported on [−2δ,2δ] such that χI≡1 on [−δ,δ], and β∈Hs+2b(T2). Then
‖χIβ‖Xs,b≲‖χI‖Hbt‖β‖Hs+2bx≲δ1/2−b‖β‖Hs+2bx. |
Proof. By definition of Xs,b spaces and by the properties of the functions χI and β we have that
‖χIβ‖Xs,b=(∑k∈Z2∫R⟨k⟩2s⟨τ−|k|2⟩2b|^χI(τ)|2|ˆβ(k)|2dτ)1/2≲(∑k∈Z2∫R⟨k⟩2(s+2b)⟨τ⟩2b|^χI(τ)|2|ˆβ(k)|2dτ)1/2≲‖χI‖Hbt‖β‖Hs+2bx. |
We now write χI as χI(t)=ψ(δ−1t), with ψ supported on [−2,2] and ψ≡1 on [−1,1], and have, since ^χI(τ)=δˆψ(δτ), that
‖χI‖Hbt=(δ2∫R⟨τ⟩2b|ˆψ(δτ)|2dτ)1/2=(δ∫R⟨δ−1τ⟩2b|ˆψ(τ)|2dτ)1/2≲δ1/2−b‖ψ‖Hbt. |
Finally, combining the two estimates above we obtain the desired inequality.
Proposition 3.9. Let h∈H1(T2), s>0, b′<b and b∈(1/2,1). Then, for p1>1/2 and s1>1,
‖χIhχIu1χIu2χIu3‖Xs,b−1≲‖χIh‖Hp1tHs1x‖χIu1‖Xs,b‖χIu2‖Xs,b′‖χIu3‖Xs,b′. |
Proof. The proof follows by using the standard technique employed to prove Xs,b-multilinear estimates (see [1]).
We proceed by duality (on L2τℓ2k) observing that
‖χIhχIu1χIu2χIu3‖Xs,b−1=supv∈L2τℓ2k‖v‖L2τℓ2k=1|∫Rτ0∫Rτ1∫Rτ2∫Rτ3^χIh(τ0,k0)^χIu1(τ1,k1)^χIu2(τ2,k2) |
×^χIu3(τ3,k3)⟨τ0+τ1+τ2+τ3−|k0+k1+k2+k3|2⟩b−1⟨k0+k1+k2+k3⟩s |
×v(τ0+τ1+τ2+τ3,k0+k1+k2+k3)dτ0τ1dτ2dτ3|. | (3.9) |
We now take the Littlewood-Paley decomposition of h,u1,u2 and u3, that is we write
h=∑N0∈2ZPN0h,uj=∑Nj∈2ZPNjuj,j=1,2,3, |
where PN is the Fourier multiplier such that ^PNu(k) is supported in the region {k∈Z2;N/2≤|k|≤N}, and replace them in (3.9) to get
(3.9)=supv∈L2τℓ2k‖v‖L2τℓ2k=1∑N0,N1,N2,N3∑|k0|∼N0|k1|∼N1|k2|∼N2|k3|∼N3∫Rτ0∫Rτ1∫Rτ2∫Rτ3|^χIPN0h(τ0,k0)||^χIPN1u1(τ1,k1)| |
⟨τ0+τ1+τ2+τ3−|k0+k1+k2+k3|2⟩b−1|^χIPN2u2(τ2,k2)|^χIPN1u3(τ3,k3)| |
×⟨k0+k1+k2+k3⟩s|v(τ0+τ1+τ2+τ3,k0+k1+k2+k3)|dτ0dτ1dτ2dτ3. | (3.10) |
We then write
∑N0,N1,N2,N3∑|k0|∼N0|k1|∼N1|k2|∼N2|k3|∼N3=∑N0,N2,N3∑N1∼max{N0,N2,N3}∑|k0|∼N0|k1|∼N1|k2|∼N2|k3|∼N3+∑N0,N2,N3∑N1≫max{N0,N2,N3}∑α∑k1∈Qα|k0|∼N0|k2|∼N2|k3|∼N3, | (3.11) |
where Qα is a cube of side ∼max{N0,N2,N3} and where α ranges over the number of cubes of side ∼max{N0,N2,N3} covering a cube of side N1. From (3.10) and (3.11) we obtain
‖χIhχIu1χIu2χIu3‖Xs,b−1≲supv∈L2τℓ2k‖v‖L2τℓ2k=1(∑N0,N2,N3,∑N1≫max{N0,N2,N3}∑αNs1 |
∫R4|^χIPN0h(τ0,k0)||^χIPQαu1(τ1,k1)||^χIPN2u2(τ2,k2)||^χIPN3u3(τ3,k3)| |
⟨τ0+τ1+τ2+τ3−|(k0+k1+k2+k3)|2⟩b−1|^P˜Qαˇv(τ0+τ1+τ2+τ3,k0+k1+k2+k3)|dτ0dτ1dτ2dτ3 |
+∑N0,N2,N3,∑N1∼max{N0,N2,N3}Ns1∫R4|^χIPN0h(τ0,k0)||^χIPQαu1(τ1,k1)||^χIPN2u2(τ2,k2)||^χIPN3u3(τ3,k3)| |
⟨τ0+τ1+τ2+τ3−|(k0+k1+k2+k3)|2⟩b−1|^P˜Qαˇv|(τ0+τ1+τ2+τ3,k0+k1+k2+k3)|dτ0dτ1dτ2dτ3). |
We now denote by ˇv the inverse Fourier transform of v(τ,k) and by ˜Qα a suitable cube of side ∼N2. With these notations in mind, by using Plancherel's theorem and the property |ˆu∗ˆχI|≤|ˆu|∗ˆχI=(|ˆu|∨χI)∧ for χI such that ˆχI≥0 (for more details see [2]), we have
‖χIhχIu1χIu2χIu3‖Xs,b−1≲supv∈L2τℓ2k‖v‖L2τℓ2k=1(∑N0,N2,N3,∑N1≫max{N0,N2,N3}∑αNs1 |
∫R×T2χI(t)|^PN0h|∨(x)|^PQαu1|∨(t,x)×|^PN2u2|∨(t,x)|^PN3u3|∨(t,x) |
(⟨τ0+τ1+τ2+τ3−|(k0+k1+k2+k3)|2⟩b−1|^P˜Qαˇv|)∨(t,x)χ3I(t)dtdx |
+∑N0,N2,N3,∑N1∼max{N0,N2,N3}Ns1∫R×T2χI(t)|^PN0h|∨(x)||^PN1u1|∨(t,x) |
×|^PN2u2|∨(t,x)|^PN3u3|∨(t,x)(⟨τ0+τ1+τ2+τ3−|(k0+k1+k2+k3)|2⟩b−1|^P˜Qαˇv|)∨χ3I(t)dtdx) |
≅supv∈L2τℓ2k‖v‖L2τℓ2k=1(I+II). |
Now the delicate part to analyze is the one where the condition N1≫max{N0,N3,N3} appears, namely, the first term I on the right hand side of the inequality above. For such a term, on denoting by M:=max{N0,N3,N3}, we have
I≲∑N0,N2,N3,∑N1≫M∑αNs1‖χI(t)|^PN0h|∨(x)‖L∞t,x‖χI|^PQαu1|∨‖L4t,x‖χI|^PN2u2|∨‖L4t,x |
‖χI|^PN3u3|∨‖L4t,x‖(⟨τ0+τ1+τ2+τ3−|(k0+k1+k2+k3)|2⟩b−1|^P˜Qαˇv|)∨‖L4t,x |
≲∑N0,N2,N3,∑N1≫M∑αNs1‖χI(t)|^PN0h|∨(x)‖Hp1tHs1x‖χI|^PQαu1|∨‖L4t,x‖χI|^PN2u2|∨‖L4t,x |
‖χI|^PN3u3|∨‖L4t,x‖(⟨τ0+τ1+τ2+τ3−|(k0+k1+k2+k3)|2⟩b−1|^P˜Qαˇv|)∨‖L4t,x |
≲‖χIh‖Hp1tHs1x‖χIu1‖Xs,b‖χIu2‖Xs,b′‖χIu3‖Xs,b′‖ˇv‖L2t,x, |
where in the first line we used Sobolev embeddings and Hölder's inequality, while in the last one we applied
‖χIPNf‖L4t,x≤Nε‖χIf‖X0,b1,b1≥1/2(1−min{ε,1/2}), |
and choose b′>1/2(1−ε), with ε satisfying 3ε<s. Now, taking the supremum over the functions v∈L2tℓ2k such that ‖v‖L2ℓ2k=1, we get the desired estimate for the term I. Finally, by similar arguments, the required estimate can be proved for the term II as well, which concludes the proof.
As a corollary of Proposition 3.9 we get the inequality below.
Corollary 3.10. Let s>0 and b∈(1/2,1) Then, for all s>0, b′<b and b∈(1/2,1), we have
‖χIu1χIu2‖Xs,b−1≲‖χIu1‖Xs,b‖χIu2‖Xs,b′. |
Proof. The proof follows form Proposition 3.9 by taking h=u3=1.
From Proposition 3.6–3.9 and Corollary 3.10 we derive the suitable (weighted) formulation of the multilinear estimates holding in the time-degenerate case.
Proposition 3.11. Let s>0, b∈(1/2,1), b′<b, and Hp,bg(R) as in Definition 2.5. Then, for h∈H1(T2) and β∈Hs+2b(T2), we have
‖g′(t)f(t)u‖Xs,bg≲‖g′f‖H1,bg‖g′u‖Xs,bg, | (3.12) |
‖g′(t)χIβ‖Xs,bg≲‖g′χI‖H2,bg‖β‖Hs+2bx, | (3.13) |
‖g′(t)χIu1χIu2‖Xs,b−1g≲‖g′(t)χIu1‖Xs,bg‖g′(t)χIu2‖Xs,b′g, | (3.14) |
and, for p1>1/2,s1>1,
‖g′(t)χIhχIu1χIu2χIu3‖Xs,b−1g≲‖g′(t)χIh‖H2,p1gHs1x‖g′(t)χIu1‖Xs,bg |
×‖g′(t)χIu2‖Xs,bg‖g′(t)χIu2‖Xs,b′g. | (3.15) |
Proof. The proof follows immediately from Proposition 3.6–3.9 and Corollary 3.10 by using the identities ‖g′(t)v‖Xs,bg=‖v(g−1(⋅),⋅)‖Xs,b and ^v∘g−1(τ)=~g′v(τ).
Let us remark that we will not be using all the estimates in Proposition 3.11 in the time-degenerate setting. However, since the estimates in Proposition 3.8, 3.9 and in Corollary 3.10 involving standard Xs,b-spaces will be used in the next section, we translated them in the time-degenerate setting for completeness.
With the previous estimates at our disposal we can now focus on the proof of the local well-posedness of the IVP
{i∂tu+g′(t)Δxu=g′(t)|u|2u,u(0,x)=u0(x). |
The proof is centered on the contraction mapping theorem and by now the argument is standard [1,13]. Still we prefer to report the details below since we will be using the modified spaces ˜Xs,bg.
Proof Proof of Theorem 1.1. We start by defining the metric space X as
X:={u∈I×T2→C;‖u‖˜Xs,bg<∞}, |
where ˜Xs,bg is as in Definition 2.2.
Now, given u0∈Hs(T2), and I=[−2δ,2δ], with 0<δ<1 to be determined later, we define the operator
Φu0(u):=χIS(t)u0+χI(t)∫t0S(t,t′)g′(t′)|χI(t′)u(t′)|2χI(t′)u(t′)dt′, |
where χI is a smooth cutoff function such that χI≡1 on [−δ,δ], and prove that Φu0 is a contraction on BR:={u;‖u‖˜Xs,bg≤R}⊂˜Xs,bg, with R=2C′δ1/2−b‖u0‖Hs(T2).
Note that, by using Proposition 3.5 and the fact that |g(I)|≲δ (which is due to the properties of g), we get, for u∈BR,
‖Φu0(u)‖˜Xs,bg=‖g′Φu0(u)‖Xs,bg |
=‖g′(t)χIS(t)u0‖Xs,bg+‖g′(t)χI(t)∫t0S(t,t′)g′(t′)|χI(t′)u(t′)|2χI(t′)u(t′)dt′‖Xs,bg |
≲|g(I)|1/2−b‖u0‖Hs(T2)+‖χIg′(t)u‖2Xs,b′g‖χIg′(t)u‖Xs,bg |
≤Cδ1/2−b‖u0‖Hs(T2)+C3δ1/2−bδ(b−b′)/4‖χIg′(t)u‖3Xs,bg |
≤R2+δ1/2−bδ(b−b′)/4R3. |
Since 14<b′<b, b∈(1/2,1), then, taking b=1/2+ε, with ε∈(0,1/16) (so b∈(12,916)), and b′=1/2−4ε>1/4, we have that 1/2−b+(b−b′)/4=ε/4>0. Therefore, by choosing δ sufficiently small and such that δ1/2−bδ(b−b′)/4R2≤1/2, from the previous inequality we obtain that Φu0 maps the ball BR into itself.
To conclude that Φu0 is a contraction we write
χIu|χIu|2−χIv|χIv|2=χI(u−v)|χIu|2+χIvχI¯uχI(u−v)+|χIv|2χI(¯u−¯v) |
and apply (3.7) to get
‖Φu0(u)−Φu0(v)‖˜Xs,bg≲(‖χIg′(t)u‖2Xs,b′g+‖χIg′(t)v‖2Xs,b′g)‖χIg′(t)(u−v)‖Xs,bg |
≲R3δ1/2−b+(b−b′)/4‖u−v‖˜Xs,bg<‖u−v‖˜Xs,bg, |
where the last inequality follows by choosing, eventually, δ smaller than before. This, finally, gives the result for t∈[−δ,δ], since, recall, χI≡1 on [−δ,δ], and
u(t,x)=S(t)u0+∫t0S(t,t′)g′(t′)|u(t′)|2(t′)u(t′)dt′,forallt∈[−δ,δ]. |
By using the previous approach we can also prove the local well-posedness of the IVP
{i∂tu+g′(t)Δxu=f(t)|u|2u,u(0,x)=u0(x), |
on I×T2, where I is a suitable finite interval of time and f is smooth enough, namely f∈H1,bg(R) for b∈(1/2,1). Note that assuming f∈H1,bg(R) in particular implies that f has to be zero at time zero of the same order of g′.
Proof Proof of Theorem 1.2. We start by defining the metric space X as
X:={u∈I×T2→C;‖u‖˜Xs,bg<∞}, |
where
˜Xs,bg:={u∈Xs,bg;g′u∈Xs,bg}, |
and
‖u‖˜Xs,bg=‖g′u‖Xs,bg. |
Now, given u0∈Hs(T2) such that ‖u0‖Hs=r, and I=[−2δ,2δ], with 0<δ<1, we define the operator
Φu0(u):=χIS(t)u0+χI(t)∫t0S(t,t′)f(t′)|χI(t′)u(t′)|2χI(t′)u(t′)dt′ |
where χI is a smooth cutoff function such that χ≡1 on [−δ,δ], and prove that Φu0 is a contraction on BR:={u;‖u‖˜Xs,bg≤R}⊂˜Xs,bg, with R=2C′δ1/2−b‖u0‖Hs(T2).
Note that, by using Proposition 3.5, Proposition 3.11, and the fact that |g(I)|\lesssim \delta (which is due to the properties of g ), we get, for u\in B_R ,
\| \Phi_{u_0}(u)\|_{\tilde{X}^{s, b}_g} = \| g'\Phi_{u_0}(u)\|_{X^{s, b}_g} |
= \| g'(t)\chi_I S(t)u_0\|_{X^{s, b}_g}+ \| g'(t) \int_0^t S(t, t') g'(t') \frac{f(t')}{g'(t')} |\chi_I(t')u(t')|^2 \chi_I(t')u(t') dt'\|_{X^{s, b}_g} |
\lesssim |g(I)|^{1/2-b}\| u_0\|_{H^s(\mathbb{T}^2)}+ \|\chi_I g'(t)u\|^2_{X^{s, b'}_g} \|g'(t')\chi_I u \, \frac{f(t')}{g'(t')}\|_{X^{s, b}_g} |
\lesssim |g(I)|^{1/2-b}\| u_0\|_{H^s(\mathbb{T}^2)}+ \|\chi_I g'(t)u\|^2_{X^{s, b'}_g} \|\chi_I g'(t) u \|_{X^{s, b}_g} \|g'(t)\frac{f(t)}{g'(t)}\|_{H^{1, b}_g} |
\leq C \delta^{1/2-b}\| u_0\|_{H^s(\mathbb{T}^2)}+ C^3 \delta^{1/2-b}\delta^{(b-b')/4}\| g'u\|_{X^{s, b}_g}^3 |
\le \frac R 2+ \delta^{1/2-b}\delta^{(b-b')/4} R^3\leq R |
for b, b' as before and \delta sufficiently small. To prove that \Phi_{u_0} is a contraction we follow the same steps as before. This completes the proof.
The one dimensional case. The previous techniques apply with few suitable modifications in the case when the problem is set on \mathbb{R}\times \mathbb{T} . Here to prove that the quintic semilinear problem (1.10) is locally well-posed with solution (multiplied by a suitable cutoff function in time) belonging to the X^{s, b}_g(\mathbb{R}\times \mathbb{T}) spaces, one translates, as in the previous case, the well-known standard Strichartz and multilinear estimates into suitable weighted Strichartz and multilinear estimates. We omit the details which are left to the interested reader.
Proof Proof of Theorem 1.5. The proof is straightforward after applying the technique used in the two-dimensional case based on the use of X^{s, b}_g -spaces.
This section is devoted to the study of the IVP
\begin{equation} \left\{ \begin{array}{l} i\partial_t u+a_1(x_1)\partial_{x_1}^2 u+a_2(x_2)\partial_{x_2}^2 u = u|u|^2, \\ u(0, x) = u_0(x), \end{array}\right. \end{equation} | (4.1) |
with a_1, a_2\in C^\infty (\mathbb{T}) real valued and strictly positive. As before, we treat in detail the two dimensional case above, since, with few modifications, the result in Theorem 1.7 about the quintic one-dimensional case can be proved.
Our strategy to treat this problem consists in combining a change of variables with a gauge transformation to reduce the problem to one where the linear part of the equation has constant coefficients.
We recall that a gauge transformation is a multiplication operator of the form
Tf(t, x): = e^{\Phi(t, x)}f(t, x) |
where the function \Phi (in this periodic setting) is periodic in x .
We start by applying in (4.1) the change of variables
(x_1, x_2) = (\alpha_1(y_1), \alpha_2(y_2)): = \alpha(y), |
so that, on denoting by v(t, y): = u(t, \alpha(y)) and by v_0(y): = u_0(\alpha(y)) , and assuming that u solves (4.1), we have
\begin{eqnarray} i\partial_t v(t, y)+\Delta_y v(t, y)& = & i\partial_tu(t, \alpha(y))+(\partial^2_{x_1}u)(t, \alpha(y))(\partial_{y_1}\alpha_1(y_1))^2 +(\partial^2_{x_2}u)(t, \alpha(y))(\partial_{y_2}\alpha_2(y_2))^2\\ &+&(\partial_{x_1}u)(t, \alpha(y))\partial^2_{y_1}\alpha_1(y_1) +(\partial_{x_2}u)(t, \alpha(y))\partial^2_{y_2}\alpha_2(y_2). \end{eqnarray} |
Then, by choosing \alpha(y) = (\alpha_1(y_1), \alpha_2(y_2)) such that \partial_{y_1}\alpha_1(y_1) = \sqrt{a_1(\alpha_1(y_1))} and \partial_{y_2}\alpha_2(y_2) = \sqrt{a_2(\alpha_2(y_2))} , and using the fact that (\partial_{x_j}u)(t, \alpha(y)) = (\partial_{y_j}v(t, y))\partial_{y_j}\alpha_j(y_j) for all j = 1, 2 , we get that v solves
\begin{equation} \left\{\begin{array}{l} i\partial_t v(t, y)+\Delta_yv(t, y)-(\partial_{y_1}v(t, y))\frac{\partial^2_{y_1}\alpha_1(y_1)}{\partial_{y_1}\alpha_1(y_1)} -(\partial_{y_2}v(t, y))\frac{\partial^2_{y_2}\alpha_2(y_2)}{\partial_{y_2}\alpha_2(y_2)} = v|v|^2, \\ v(0, y) = v_0(y). \end{array}\right. \end{equation} | (4.2) |
We now apply the gauge transform
Tf(t, y): = e^{\Phi(y)}v(t, y) = \exp\Big\{-\frac 1 2\int_{0}^{y_1}\frac{\alpha_1''(s_1)}{\alpha_1'(s_1)}ds_1-\frac 1 2 \int_{0}^{y_2}\frac{\alpha_2''(s_2)}{\alpha_2'(s_2)}ds_2 \Big\}f(t, y), |
(where, note, \Phi is periodic since \partial_{y_j}\alpha_j(y_j) = \sqrt{a_j(\alpha_j(y_j))} ) on the left and right hand side of (4.2) and have, on denoting by w(t, y): = e^{\Phi(y)} v(t, y) ,
i\partial_t w+\Delta_y w- \left( \partial_{y_1}^2\Phi+ \partial_{y_2}^2\Phi+(\partial_{y_1}\Phi)^2+(\partial_{y_2}\Phi)^2\right)w = e^{-2\Phi}w|w|^2. |
Finally, we reduced the study of (4.1) to the study of
\begin{equation} \left\{\begin{array}{l} i\partial_t w+\Delta_y w = e^{-2\Phi}w|w|^2-\beta w\\ w(0, y) = w_0(y) \end{array}\right. \end{equation} | (4.3) |
with \beta = \beta(y) = \partial_{y_1}^2\Phi+ \partial_{y_2}^2\Phi+(\partial_{y_1}\Phi)^2+(\partial_{y_2}\Phi)^2 , and w_0(y) = e^{\Phi(y)} u_0(\alpha(y)) .
Theorem 4.1. Let s > 0 and \beta\in H^{s+2b'}(\mathbb{T}^2) with b' > 1/2-\varepsilon for some \varepsilon > 0 . Then, given \Phi such that e^{-2\Phi}\in H^1(\mathbb{T}^2) , for every w_0\in H^s(\mathbb{T}^2) there exists a unique solution of the IVP (4.3) in the time interval [-T, T] for a suitable T = T(\|w_0\|_{H^s(\mathbb{T}^2)}) . Moreover, the solution w satisfies
w\in C([-T, T];H^s(\mathbb{T}^2)) |
and
\chi_Iw\in X^{s, b}(\mathbb{R}\times\mathbb{T}^2), |
with b\in (1/2, 1) such that b > b' , I a closed neighborhood of [-T, T] , and \chi_I a smooth cutoff fuction such that \chi_I\equiv 1 on [-T, T] .
Proof. In this proof we consider the metric space X = X^{s, b}(\mathbb{R}\times\mathbb{T}^2) and take w_0\in H^s(\mathbb{T}^2) . We then consider I = [-2\delta, 2\delta] , with 0 < \delta < 1 to be determined later, and define the operator
\Psi_{w_0}(w): = \chi_I S(t)w_0+\chi_I(t)\int_0^tS(t, t')\Big( \chi_I^2e^{-2\Phi}w(t')|\chi_Iw(t')|^2-\chi_I \beta(y) \chi_I w(t') \Big)dt', |
where \chi_I is a smooth cutoff function such that \chi_I\equiv 1 on [-\delta, \delta] . Now, given B_R: = \{w; \|w\|_{X^{s, b}}\leq R\} , with R = 2C\delta^{1/2-b}\|w_0\|_{H^s(\mathbb{T}^2)} and C suitable (see the rest of the proof), we prove that \Psi_{w_0} is a contraction. By using Proposition 3.7, 3.9 and Corollary 3.10, we get, for w\in B_R , p_1 > 1/2, s_1 > 1 (here we take p_1 = 1/2+\varepsilon' , with 0 < \varepsilon' < 1 small enough),
\begin{eqnarray} \|\Psi_{w_0}(w)\|_{X^{s, b}}&\leq& C_1\delta^{1/2-b}\|w_0\|_{H^s(\mathbb{T}^2)}+C_2\|\chi_Ie^{-2\Phi}\|_{H^{p_1}_tH^{s_1}_x}\delta^{1/2-b+(b-b')/4}\|\chi_I w\|^3_{X^{s, b}}\\ &+& \|\chi_I \beta\|_{X^{s, b'}}\|\chi_Iw\|_{X^{s, b}} , \\ &\leq&C_1R+C_2'R+C'_3\delta^{1/2-b+(b-b')/4}R^3\leq R \end{eqnarray} | (4.4) |
for b, b' as above (see, for instance, the proof of Theorem 1.1), \delta small enough and C suitable, for instance C = 3\max\{C_1, C_2', C_3'\} . Note that in (4.4) we used the estimate in Proposition 3.8, that is, \|\chi_I \beta\|_{X^{s, b'}}\lesssim \|\chi_I\|_{H^{b'}_t}\|\beta\|_{H^{s+2b'}_y(\mathbb{T}^2)}\lesssim \delta^{1/2-b'} C\lesssim C for b' = 1/2-\varepsilon , since 1/2-\varepsilon < b' < b, b\in (1/2, 1) . By similar arguments, eventually by choosing a different suitable constant C , we can conclude that \Phi_{w_0} is a contraction on B_R . This, finally, concludes the proof (with T = \delta ).
Proof Proof of Theorem 1.3. In order to prove Theorem 1.3 it is enough to apply Theorem 4.1 to the IVP (4.3). To do that we need that \beta\in H^{s+b'}(\mathbb{T}^2) and that e^{-2\Phi}\in H^1(\mathbb{T}^2) . Since we have that a_1, \, a_2 \in C^\infty(\mathbb{T}^2) , a_1, \, a_2 are strictly positive, \alpha_1'(y_1) = \sqrt{a_1(\alpha_1(y_1))} > 0 , and \alpha_2'(y_2) = \sqrt{a_2(\alpha_2(y_2))} > 0 , then we get that \Phi\in C^\infty(\mathbb{T}^2) , and, consequently, that \beta and e^{-2\Phi} have the required properties. Therefore, by Theorem 4.1, the IVP (4.3) is locally-well posed in H^s for all s > 0 , and, going back to the IVP (4.2) with the inverse gauge transform, we can conclude that the latter is also locally well posed in H^s , for all s > 0 , and that the solution v satisfies \chi_I v\in X^{s, b}_\Phi(\mathbb{R}\times \mathbb{T}^2) , where, recall,
X^{s, b}_\Phi(\mathbb{R}\times \mathbb{T}^2): = \{f:\mathbb{R}\times \mathbb{T}^2\rightarrow \mathbb{C}; e^{\Phi}f\in X^{s, b}(\mathbb{R}\times \mathbb{T}^2)\}. |
Finally, the solution u of (4.1), will belong to the space X^{s, b}_{\Phi, \tilde{\alpha}}: = \{f:\mathbb{R}\times \mathbb{T}^2\rightarrow \mathbb{C}; (e^{\Phi}\, f)\circ \tilde{\alpha}\in X^{s, b}(\mathbb{R}\times \mathbb{T}^2)\} , where \tilde{\alpha}: \mathbb{R}\times \mathbb{T}^2\rightarrow \mathbb{R}\times \mathbb{T}^2 is the diffeomorphism given by \tilde{\alpha}(t, y): = (t, \alpha(y)) . This concludes the proof.
Remark 4.2. Going back to the proof that we just concluded it is easy to see that one does not need to require C^\infty (\mathbb{T}) regularity for the coefficients a_1, \, a_2 , in fact H^2 is enough. Here we did not want to concentrate on the regularity of the coefficients, but rather on the required regularity of the initial data, and show that it is the same as the one required for a completely flat torus, that is H^s, \, s > 0 .
Remark 4.3. By combining the arguments applied in this and in the previous sections one can also consider a problem of the form
\begin{equation} \left\{\begin{array}{l} i\partial_tu+ g'(t)\sum\limits_{j = 1}^2a_j(x_j)\partial_{x_j} u = f(t)u|u|^2, \\ u(0, x) = u_0(x), \end{array}\right. \end{equation} | (4.5) |
where g, f and a_j , j = 1, 2 , satisfy the same properties as above.
In this case, performing a change of variables in space first, and, afterwards, applying a gauge transform with a suitable \Phi = \Phi(y) , one finally reduces the problem (4.5) to one of the form (3.2) (eventually with a nonlinearity of the form f(t)e^{-2\Phi}w|w|^2-g'(t)\beta(x)w ) to which Theorem 1.2 applies with few modifications. Finally one goes back to the solution of the original initial value problem by using the inverse gauge transform and changing variables again. The solution obtained this way will belong to a suitable weighted space \tilde{X}^{s, b}_{g, \Phi, \tilde{ \alpha}}(\mathbb{R}\times \mathbb{T}^2): = \{f:\mathbb{R}\times \mathbb{T}^2\rightarrow \mathbb{C}; (e^{\Phi}f)\circ \tilde{\alpha}\in \tilde{X}^{s, b}_g(\mathbb{R}\times \mathbb{T}^2)\} with \tilde{ \alpha} as before.
The one dimensional case. Is it easy to see that in the one dimensional case, that is when the problem is considered on \mathbb{R}\times\mathbb{T} , the presence of a single space variable allows us to apply the same scheme applied in the two-dimensional case.
Proof Proof of Theorem 1.7. The proof is straightforward and follows from the two-dimensional strategy. Therefore, after a change of variable in space followed by the application of a gauge transform with a space-dependent function \Phi , the problem is translated in a constant coefficients one. The solution of the resulting constant coefficients nonlinear problem will, finally, provide the solution of the original one via the use of the inverse gauge transform and a change of variables.
Remark 4.4. We remark again that we expect that all the results of this paper can be generalized to the higher dimensional setting with few modification. We also expect the results in the time-degenerate case to be true on general closed manifolds combining our strategy with the known results mentioned in the introduction.
Serena Federico has received funding from the European Unions Horizon 2020 research and innovation programme under the Marie Skłodowska-Curie grant agreement No 838661 and by the FWO Odysseus 1 grant G.0H94.18N: Analysis and Partial Differential Equations. Gigliola Staffilani is funded in part by DMS-1764403 and by the Simons Foundation.
The authors declare no conflict of interest.
[1] |
J. Bourgain, Fourier transform restriction phenomena for certain lattice subsets and applications to nonlinear evolution equations I. Schrödinger equations, Geom. Funct. Anal., 3 (1993), 107–156. doi: 10.1007/BF01896020
![]() |
[2] | J. Bourgain, Problems in Hamiltonian PDE's, In: Visions in Mathematics, Birkhäuser Basel, 2000, 32–56. |
[3] | J. Bourgain, C. Demeter, The proof of the \ell^2 decoupling conjecture, Ann. Math., 182 (2015), 351–389. |
[4] |
N. Burq, P. Gérard, N. Tzvetkov, Strichartz inequalities and the nonlinear Schrödinger equation on compact manifolds, Am. J. Math., 126 (2004), 569–605. doi: 10.1353/ajm.2004.0016
![]() |
[5] |
Y. M. Chen, S. H. Ma, Z. Y. Ma, Solitons for the cubic-quintic nonlinear Schrödinger equation with varying coefficients, Chinese Phys. B, 21 (2012), 050510. doi: 10.1088/1674-1056/21/5/050510
![]() |
[6] |
M. Cicognani, M. Reissig, Well-posedness for degenerate Schrödinger equations, Evol. Equ. Control The., 3 (2014), 15–33. doi: 10.3934/eect.2014.3.15
![]() |
[7] | C. Fan, Y. Ou, G. Staffilani, H. Wang, 2D-defocusing nonlinear Schrödinger equation with random data on irrational tori, Stoch. Partial Differ., 9 (2021), 142–206. |
[8] | S. Federico, M. Ruzhansky, Smoothing and strichartz estimates for degenerate Schrödinger-type equations, 2020, arXiv: 2005.01622. |
[9] |
S. Federico, G. Staffilani, Smoothing effect for time-degenerate Schrödinger operators, J. Differ. Equations, 298 (2021), 205–247. doi: 10.1016/j.jde.2021.07.006
![]() |
[10] |
Z. Hani, A bilinear oscillatory integral estimate and bilinear refinements to Strichartz estimates on closed manifolds, Anal. PDE, 5 (2012), 339–363. doi: 10.2140/apde.2012.5.339
![]() |
[11] |
C. E. Kenig, G. Ponce, C. Rolvung, L. Vega, Variable coefficient Schrödinger flows for ultrahyperbolic operators, Adv. Math., 196 (2005), 373–486. doi: 10.1016/j.aim.2004.02.002
![]() |
[12] | C. E. Kenig, G. Ponce, L. Vega, The Cauchy problem for the Korteweg-de Vries equation in Sobolev spaces of negative indices, Duke Math. J., 71 (1993), 1–21. |
[13] |
C. E. Kenig, G. Ponce, L. Vega, A bilinear estimate with applications to the KdV equation, J. Am. Math. Soc., 9 (1996), 573–603. doi: 10.1090/S0894-0347-96-00200-7
![]() |
[14] |
B. Li, X. F. Zhang, Y. Q. Li, W. M. Liu, Propagation and interaction of matter-wave solitons in Bose-Einstein condensates with time-dependent scattering length and varying potentials, J. Phys. B At. Mol. Opt., 44 (2011), 175301. doi: 10.1088/0953-4075/44/17/175301
![]() |
[15] |
J. Marzuola, J. Metcalfe, D. Tataru, Strichartz estimates and local smoothing estimates for asymptotically flat Schrödinger equations, J. Funct. Anal., 255 (2008), 1497–1553. doi: 10.1016/j.jfa.2008.05.022
![]() |
[16] |
H. Mizutani, N. Tzvetkov, Strichartz estimates for non-elliptic Schrödinger equations on compact manifolds, Commun. Part. Diff. Eq., 40 (2015), 1182–1195. doi: 10.1080/03605302.2015.1010211
![]() |
[17] |
D. Salort, The Schrödinger equation type with a nonelliptic operator, Commun. Part. Diff. Eq., 32 (2007), 209–228. doi: 10.1080/03605300601128074
![]() |
[18] |
G. Staffilani, D. Tataru, Strichartz estimates for a Schrödinger operator with nonsmooth coefficients, Commun. Part. Diff. Eq., 27 (2002), 1337–1372. doi: 10.1081/PDE-120005841
![]() |
[19] |
H. Wang, B. Li, Solitons for a generalized variable-coefficient nonlinear Schrödinger equation, Chinese Phys. B, 20 (2011), 040203. doi: 10.1088/1674-1056/20/4/040203
![]() |
[20] |
C. L. Zheng, Y. Li, Exact projective solutions of a generalized nonlinear Schrödinger system with variable parameters, Chinese Phys. B, 21 (2012), 70305. doi: 10.1088/1674-1056/21/7/070305
![]() |
1. | Roberta Bianchini, Chiara Saffirio, Fluid instabilities, waves and non-equilibrium dynamics of interacting particles: a short overview, 2022, 5, 2640-3501, 1, 10.3934/mine.2023033 | |
2. | Serena Federico, 2022, Chapter 2, 978-3-031-24310-3, 19, 10.1007/978-3-031-24311-0_2 | |
3. | Serena Federico, Carleman estimates for third order operators of KdV and non KdV-type and applications, 2024, 203, 0373-3114, 2801, 10.1007/s10231-024-01467-7 | |
4. | Serena Federico, Michael Ruzhansky, Smoothing and Strichartz estimates for degenerate Schrödinger-type equations, 2024, 242, 0362546X, 113500, 10.1016/j.na.2024.113500 |