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

On the Cauchy problem for the Hartree approximation in quantum dynamics

Rémi Carles Univ Rennes, CNRS
IRMAR - UMR 6625
F-35000 Rennes, France
Remi.Carles@math.cnrs.fr
Clotilde Fermanian Kammerer Univ Paris Est Creteil, CNRS, LAMA, F-94010 Creteil, France
Univ Gustave Eiffel, LAMA, F-77447 Marne-la-Vallée, France
clotilde.fermanian@u-pec.fr
 and  Caroline Lasser Technische Universität München, Zentrum Mathematik, Deutschland classer@ma.tum.de
Abstract.

We prove existence and uniqueness results for the time-dependent Hartree approximation arising in quantum dynamics. The Hartree equations of motion form a coupled system of nonlinear Schrödinger equations for the evolution of product state approximations. They are a prominent example for dimension reduction in the context of the the time-dependent Dirac–Frenkel variational principle. We handle the case of Coulomb potentials thanks to Strichartz estimates. Our main result addresses a general setting where the nonlinear coupling cannot be considered as a perturbation. The proof uses a recursive construction that is inspired by the standard approach for the Cauchy problem associated to symmetric quasilinear hyperbolic equations.

RC was supported by Rennes Métropole through its AIS program, and by Centre Henri Lebesgue, program ANR-11-LABX-0020-0. CL was supported by the Centre for Advanced Study in Oslo, Norway, research project Attosecond Quantum Dynamics Beyond the Born-Oppenheimer Approximation.

1. Introduction

We consider the time-dependent Schrödinger equation

(1.1) itψ=Hψ,i{\partial}_{t}\psi=H\psi,

where the total Hamiltonian is given by

H=Hx+Hy+w(x,y),Hx=12Δx+V1(x),Hy=12Δy+V2(y)H=H_{x}+H_{y}+w(x,y),\quad H_{x}=-\frac{1}{2}\Delta_{x}+V_{1}(x),\quad H_{y}=-\frac{1}{2}\Delta_{y}+V_{2}(y)

with xd1x\in{\mathbb{R}}^{d_{1}} and yd2y\in{\mathbb{R}}^{d_{2}}, d1,d21d_{1},d_{2}\geq 1. The potentials V1,V2V_{1},V_{2} and ww are always real-valued, we will make extra regularity and decay assumptions later on . It is common wisdom that for dealing with quantum systems “a solution of the wave equation in many-dimensional space is far too complicated to be practicable” (Dirac 1930) and one aims at approximative methods that effectively reduce the space dimension. Here, we focus on initial data that decouple the space variables,

ψ(0,x,y)=ϕ0x(x)ϕ0y(y).\psi(0,x,y)=\phi^{x}_{0}(x)\phi^{y}_{0}(y).

Such a form of initial data indeed suggests a dimension reduction approach. Of course, if there is no coupling (w(x,y)=0w(x,y)=0), the full solution is itself a product state ψ(t,x,y)=ϕx(t,x)ϕy(t,y)\psi(t,x,y)=\phi_{x}(t,x)\phi_{y}(t,y), with

{itϕx=Hxϕx,ϕx(0,x)=ϕ0x(x),itϕy=Hyϕy,ϕy(0,y)=ϕ0y(y).\begin{cases}i\partial_{t}\phi^{x}=H_{x}\phi^{x},\quad\phi^{x}(0,x)=\phi^{x}_{0}(x),\\ i\partial_{t}\phi^{y}=H_{y}\phi^{y},\quad\phi^{y}(0,y)=\phi^{y}_{0}(y).\end{cases}

When the coupling is present, one seeks for approximate solutions u(t)ψ(t)u(t)\approx\psi(t) of product form u(t,x,y)=ϕx(t,x)ϕy(t,y)u(t,x,y)=\phi^{x}(t,x)\phi^{y}(t,y) in order to reduce the initial system (1.1) in d1+d2{\mathbb{R}}^{d_{1}+d_{2}} to two systems on spaces of smaller dimensions, d1{\mathbb{R}}^{d_{1}} and d2{\mathbb{R}}^{d_{2}}. In situations, where the overall configuration space has a natural decomposition of its dimension d1++dNd_{1}+\cdots+d_{N}, a corresponding product ansatz of NN factors is sought. Here we only investigate the case N=2N=2, mentioning that repeated application of the binary construction yields the more general case.  Applying the time-dependent Dirac–Frenkel variational principle to the manifold

={u=φxφyφxL2(d1),φyL2(d2)}\mathcal{M}=\left\{u=\varphi^{x}\otimes\varphi^{y}\mid\varphi^{x}\in L^{2}({\mathbb{R}}^{d_{1}}),\ \varphi^{y}\in L^{2}({\mathbb{R}}^{d_{2}})\right\}

yields the so-called time-dependent Hartree approximation,

ψ(t,x,y)ϕx(t,x)ϕy(t,y),\psi(t,x,y)\approx\phi^{x}(t,x)\phi^{y}(t,y)\in\mathcal{M},

where the pair (ϕx,ϕy)(\phi^{x},\phi^{y}) solves the nonlinearly coupled system

(1.2) {itϕx=Hxϕx+wyϕx,ϕx(0,x)=ϕ0x(x),itϕy=Hyϕy+wxϕy,ϕy(0,y)=ϕ0y(y).\begin{cases}i\partial_{t}\phi^{x}=H_{x}\phi^{x}+\langle w\rangle_{y}\phi^{x},\quad\phi^{x}(0,x)=\phi^{x}_{0}(x),\\ i\partial_{t}\phi^{y}=H_{y}\phi^{y}+\langle w\rangle_{x}\phi^{y},\quad\phi^{y}(0,y)=\phi^{y}_{0}(y).\end{cases}

The time-dependent potentials result from the averaging process

wy(t,x):=d2w(x,y)|ϕy(t,y)|2𝑑y,\displaystyle\langle w\rangle_{y}(t,x):=\int_{{\mathbb{R}}^{d_{2}}}w(x,y)|\phi^{y}(t,y)|^{2}dy,
wx(t,y):=d1w(x,y)|ϕx(t,x)|2𝑑x,\displaystyle\langle w\rangle_{x}(t,y):=\int_{{\mathbb{R}}^{d_{1}}}w(x,y)|\phi^{x}(t,x)|^{2}dx,

under the assumption, made throughout this paper, that

(1.3) ϕ0xL2(d1)=ϕ0yL2(d2)=1.\|\phi_{0}^{x}\|_{L^{2}({\mathbb{R}}^{d_{1}})}=\|\phi_{0}^{y}\|_{L^{2}({\mathbb{R}}^{d_{2}})}=1.

For any “reasonable” solution (at least with the regularity considered in this paper), the L2L^{2}-norms of ϕx(t,)\phi^{x}(t,\cdot) and ϕy(t,)\phi^{y}(t,\cdot), respectively, are independent of time, hence

ϕx(t)L2(d1)=ϕy(t)L2(d2)=1,\|\phi^{x}(t)\|_{L^{2}({\mathbb{R}}^{d_{1}})}=\|\phi^{y}(t)\|_{L^{2}({\mathbb{R}}^{d_{2}})}=1,

for all tt in the time interval where the solution to (1.2) is well-defined; see §6.3 for a proof.

Even though the time-dependent Hartree approximation is one of the most fundamental approximations in quantum dynamics, mathematical existence and uniqueness proofs are rather scarce.  Existence and uniqueness have been studied in the case where the interaction potential is of convolution type, i.e. for w(x,y)=W(xy)w(x,y)=W(x-y) and with one of the subsystems moving by classical mechanics (see [7, 2, 6] for example). A related investigation has targeted the time-dependent self-consistent field system [12] with coupling potentials of Schwartz class. However, our aim here is to discuss the existence and uniqueness of solutions for system (1.2) when the potentials wx\langle w\rangle_{x} or wy\langle w\rangle_{y} need not be bounded, and cannot be considered as a perturbation of V2V_{2} or V1V_{1}, respectively. This framework requires a different approach. In particular, our result provides the Cauchy theory for the systems discussed in the articles [3, 4] where the accuracy of the Hartree approximation is studied in the broader context of composite quantum dynamics and scale separation.

The steps for our existence and uniqueness proof are strongly inspired by the method which is classical in the study of quasilinear hyperbolic systems, see e.g. [1]. With nn\in{\mathbb{N}}, we associate the iterative scheme of recursive equations

(1.4) {itϕn+1x=Hxϕn+1x+wnyϕn+1x,ϕn+1x(0,x)=ϕ0x(x),itϕn+1y=Hyϕn+1y+wnxϕn+1y,ϕn+1y(0,y)=ϕ0y(y),\left\{\begin{aligned} i\partial_{t}\phi^{x}_{n+1}=H_{x}\phi^{x}_{n+1}+\langle w_{n}\rangle_{y}\phi^{x}_{n+1},\quad\phi^{x}_{n+1}(0,x)=\phi^{x}_{0}(x),\\ i\partial_{t}\phi^{y}_{n+1}=H_{y}\phi^{y}_{n+1}+\langle w_{n}\rangle_{x}\phi^{y}_{n+1},\quad\phi^{y}_{n+1}(0,y)=\phi^{y}_{0}(y),\end{aligned}\right.

with

(1.5) wny(t,x)=nw(x,y)|ϕny(t,y)|2𝑑y,wnx(t,y)=dw(x,y)|ϕnx(t,x)|2𝑑x.\langle w_{n}\rangle_{y}(t,x)=\int_{{\mathbb{R}}^{n}}w(x,y)|\phi^{y}_{n}(t,y)|^{2}dy,\quad\langle w_{n}\rangle_{x}(t,y)=\int_{{\mathbb{R}}^{d}}w(x,y)|\phi^{x}_{n}(t,x)|^{2}dx.

The main steps of the proof of our existence and uniqueness result are then:

  1. (1)

    The iterative scheme is well-defined and enjoys bounds in “large” norm, which control second order derivatives and polynomial growth of order two for some finite time horizon.

  2. (2)

    The solution of the scheme converges in “small” norm, which is the L2L^{2} norm.

  3. (3)

    It is possible to pass to the limit n+n\to+\infty in the equation, which leads to the construction of a solution that one then proves to be unique and global in time (provided the initial data is regular enough).

1.1. Outline

In the next Section 2, we recall elementary properties of the time-dependent variational principle and formally derive the Hartree equations (1.2). Then we discuss coupling potentials w(x,y)w(x,y) of Coulombic and of polynomial type in Section 3, where we also present our main result Theorem 3.11, which establishes existence and uniqueness of the solutions to the Hartree system for coupling with polynomial growth. The different steps of the proof of Theorem 3.11 are the subject of Section 4 (analysis of the iterative scheme), Section 5 (convergence in small norms) and Section 6 (passing to the limit). A sufficient condition for the growth of the coupling potential is verified in Section 7. The Appendices A and B summarize some technical arguments.

1.2. Notations

We write LTL^{\infty}_{T} for L([0,T])L^{\infty}([0,T]). The notations Lx2L^{2}_{x}, Ly2L^{2}_{y}, Lx,y2L^{2}_{x,y} stand for L2(xd1)L^{2}({\mathbb{R}}^{d_{1}}_{x}), L2(yd2)L^{2}({\mathbb{R}}^{d_{2}}_{y}), L2(x,yd1+d2)L^{2}({\mathbb{R}}^{d_{1}+d_{2}}_{x,y}), respectively. We denote by ,Lx2{\langle}\cdot,\cdot{\rangle}_{L^{2}_{x}}, ,Ly2{\langle}\cdot,\cdot{\rangle}_{L^{2}_{y}}, ,Lx,y2{\langle}\cdot,\cdot{\rangle}_{L^{2}_{x,y}} the corresponding inner products. For f,g0f,g\geq 0, we write fgf\lesssim g whenever there exists a “universal” constant (in the sense that it does not depend on time, space, or nn, typically) such that fCgf\leq Cg.

2. Variational principle

The time-dependent Hartree approximation results from the Dirac–Frenkel variational principle applied to the manifold

(2.1) ={u=φxφyφxLx2,φyLy2},\mathcal{M}=\left\{u=\varphi^{x}\otimes\varphi^{y}\mid\varphi^{x}\in L^{2}_{x},\ \varphi^{y}\in L^{2}_{y}\right\},

see also [15, §II.3.1] for the analogous discussion with Hartree products of NN orbitals in L2(3)L^{2}({\mathbb{R}}^{3}). The reader can also refer to [14]. The principle determines an approximate solution u(t)u(t)\in\mathcal{M} for the time-dependent Schrödinger equation

itψ=Hψi\partial_{t}\psi=H\psi

with initial data ψ(0)\psi(0)\in\mathcal{M} by requiring that for all times tt

(2.2) {tu(t)𝒯u(t),v,itu(t)Hu(t)=0for allv𝒯u(t),\left\{\begin{array}[]{l}\partial_{t}u(t)\in\mathcal{T}_{u(t)}\mathcal{M},\\[4.30554pt] \langle v,i\partial_{t}u(t)-Hu(t)\rangle=0\quad\text{for all}\quad v\in\mathcal{T}_{u(t)}\mathcal{M},\end{array}\right.

where 𝒯u(t)\mathcal{T}_{u(t)}\mathcal{M} denotes the tangent space of \mathcal{M} at u(t)u(t).  For deriving the Hartree equations, we first have to understand the manifold \mathcal{M} and its tangent space. Note that the representation of a Hartree function is non-unique, since φxφy=(aφx)(a1φy)\varphi^{x}\otimes\varphi^{y}=(a\varphi^{x})\otimes(a^{-1}\varphi^{y}) for any a{0}a\in{\mathbb{C}}\setminus\{0\}. However, we can have unique representations in the tangent space once appropriate gauge conditions are set.

Lemma 2.1 (Tangent space).

For any u=φxφyu=\varphi^{x}\otimes\varphi^{y}\in\mathcal{M}, u0u\neq 0,

𝒯u={vxφy+φxvyvxLx2,vyLy2}.\mathcal{T}_{u}\mathcal{M}=\left\{v^{x}\otimes\varphi^{y}+\varphi^{x}\otimes v^{y}\mid v^{x}\in L^{2}_{x},\ v^{y}\in L^{2}_{y}\right\}.

Any v𝒯uv\in\mathcal{T}_{u}\mathcal{M} has a unique representation of the form v=vxφy+φxvyv=v^{x}\otimes\varphi^{y}+\varphi^{x}\otimes v^{y}, if we impose the gauge condition φx,vx=0\langle\varphi^{x},v^{x}\rangle=0. The tangent spaces are complex linear subspaces of Lx,y2L^{2}_{x,y} such that u𝒯uu\in\mathcal{T}_{u}\mathcal{M} for all uu\in\mathcal{M}.

The lemma is proved in Appendix A. The following formal arguments show that, in case that the variational solution u(t)u(t) is well-defined and sufficiently regular, the L2L^{2} norm and the energy expectation value are conserved automatically. Indeed, we differentiate with respect to time tt and use the variational condition (2.2) for v=u(t)v=u(t),

ddtu(t)Lx,y22=2Reu(t),tu(t)Lx,y2=2Reu(t),1iHu(t)Lx,y2=0,\frac{d}{dt}\|u(t)\|^{2}_{L^{2}_{x,y}}=2\,\mathrm{Re}{\langle}u(t),\partial_{t}u(t){\rangle}_{L^{2}_{x,y}}=2\,\mathrm{Re}{\langle}u(t),\tfrac{1}{i}Hu(t){\rangle}_{L^{2}_{x,y}}=0,

due to self-adjointness of the Hamiltonian. Similarly, using self-adjointness and the variational condition (2.2) for v=tu(t)v=\partial_{t}u(t),

ddtu(t),Hu(t)Lx,y2=2Retu(t),Hu(t)Lx,y2=2Retu(t),itu(t)Lx,y2=0.\frac{d}{dt}\langle u(t),Hu(t)\rangle_{L^{2}_{x,y}}=2\,\mathrm{Re}\langle\partial_{t}u(t),Hu(t)\rangle_{L^{2}_{x,y}}=2\,\mathrm{Re}\langle\partial_{t}u(t),i\partial_{t}u(t)\rangle_{L^{2}_{x,y}}=0.

Let us now formally derive the Hartree system (1.2). We write

u(t)=φx(t)φy(t),u(t)=\varphi^{x}(t)\otimes\varphi^{y}(t),

with φx(t)Lx2=φy(t)Ly2=1\|\varphi^{x}(t)\|_{L^{2}_{x}}=\|\varphi^{y}(t)\|_{L^{2}_{y}}=1. We have

itu\displaystyle i\partial_{t}u =(itφx(t))φy(t)+φx(t)(itφy(t)),\displaystyle=(i\partial_{t}\varphi^{x}(t))\otimes\varphi^{y}(t)+\varphi^{x}(t)\otimes(i\partial_{t}\varphi^{y}(t)),
Hu\displaystyle Hu =Hxφx(t))φy(t)+φx(t)(Hyφy(t))+w(x,y)φx(t)φy(t).\displaystyle=H_{x}\varphi^{x}(t))\otimes\varphi^{y}(t)+\varphi^{x}(t)\otimes(H_{y}\varphi^{y}(t))+w(x,y)\varphi^{x}(t)\otimes\varphi^{y}(t).

Choosing elements v=vxφy+φxvy𝒯u(t)v=v^{x}\otimes\varphi^{y}+\varphi^{x}\otimes v^{y}\in\mathcal{T}_{u(t)}\mathcal{M} and evaluating (2.2), we obtain the following necessary and sufficient conditions:

  • (i)

    If vyv^{y}=0, we obtain that for all vxLx2v^{x}\in L^{2}_{x} such that vx,φx(t)Lx2=0\langle v^{x},\varphi^{x}(t)\rangle_{L^{2}_{x}}=0,

    vx,(itHx)φx(t)Lx2=d1+d2w(x,y)vx(x)φx(t,x)|φy(t,y)|2𝑑x𝑑y.\langle v^{x},(i\partial_{t}-H_{x})\varphi^{x}(t)\rangle_{L^{2}_{x}}=\int_{{\mathbb{R}}^{d_{1}+d_{2}}}w(x,y)v^{x}(x)\varphi^{x}(t,x)|\varphi^{y}(t,y)|^{2}dxdy.
  • (ii)

    If vxv^{x}=0, we obtain that for all vyLy2v^{y}\in L^{2}_{y} such that vy,φy(t)Ly2=0\langle v^{y},\varphi^{y}(t)\rangle_{L^{2}_{y}}=0,

    vy,(itHy)φy(t)Ly2=d1+d2w(x,y)vy(y)φy(t,y)|φx(t,x)|2𝑑x𝑑y.\langle v^{y},(i\partial_{t}-H_{y})\varphi^{y}(t)\rangle_{L^{2}_{y}}=\int_{{\mathbb{R}}^{d_{1}+d_{2}}}w(x,y)v^{y}(y)\varphi^{y}(t,y)|\varphi^{x}(t,x)|^{2}dxdy.

The choice of φx(t)\varphi^{x}(t) and φy(t)\varphi^{y}(t) satisfying the Hartree system (1.2) guarantees (i) and (ii).

3. Main result

We present existence and uniqueness results for the solution of the time-dependent Hartree system (1.2). In §3.1, we discuss how Strichartz estimates may be applied to Coulombic coupling. In §3.2, we give detailed assumptions on polynomial growth conditions. Then, in §3.3 we state our main result Theorem 3.11.

3.1. Coupling potentials of Coulombic form

The case of Coulomb singularities (in combination with classical nuclear dynamics) has already been addressed in [7, 2] by Schauder and Picard fixed point arguments, respectively. We briefly revisit the main result from [7], and show how it may be adapted thanks to Strichartz estimates. We suppose d1=d2=3d_{1}=d_{2}=3, and have in mind the case

(3.1) w(x,y)=ε|xy|,ε.w(x,y)=\frac{\varepsilon}{|x-y|},\quad\varepsilon\in{\mathbb{R}}.

We consider more generally the case w(x,y)=W(xy)w(x,y)=W(x-y), for a possibly singular WW. We assume that the potentials V1V_{1} and V2V_{2} are perturbations of smooth and at most quadratic potentials:

Vj=𝐕j+vj,V_{j}={\mathbf{V}}_{j}+v_{j},

where

𝐕j𝒬={VC(3;),αVL(3),|α|2},{\mathbf{V}}_{j}\in\mathcal{Q}=\left\{V\in C^{\infty}({\mathbb{R}}^{3};{\mathbb{R}}),\ {\partial}^{\alpha}V\in L^{\infty}({\mathbb{R}}^{3}),\ \forall|\alpha|\geq 2\right\},

and

vjLp(3)+L(3),for some p>3/2.v_{j}\in L^{p}({\mathbb{R}}^{3})+L^{\infty}({\mathbb{R}}^{3}),\quad\text{for some }p>3/2.

Typically, we may consider Coulomb potentials, as

1|x|=1|x|𝟏|x|<1+1|x|𝟏|x|1.\frac{1}{|x|}=\frac{1}{|x|}{\mathbf{1}}_{|x|<1}+\frac{1}{|x|}{\mathbf{1}}_{|x|\geq 1}.

The first term on the right hand side belongs to Lp(3)L^{p}({\mathbb{R}}^{3}) for any 1p<31\leq p<3, and the second term is obviously bounded. We then make the same assumption on WW. Denote

𝐇x=12Δx+𝐕1,𝐇y=12Δy+𝐕2.\mathbf{H}_{x}=-\frac{1}{2}\Delta_{x}+\mathbf{V}_{1},\quad\mathbf{H}_{y}=-\frac{1}{2}\Delta_{y}+\mathbf{V}_{2}.

Then eit𝐇xe^{-it\mathbf{H}_{x}} and eit𝐇ye^{-it\mathbf{H}_{y}} enjoy Strichartz estimates, and (1.2) can be solved at the L2L^{2} level, by a straightforward adaptation of [10, Corollary 4.6.5]:

Theorem 3.1.

Assume d1=d2=3d_{1}=d_{2}=3, 𝐕1,𝐕2𝒬\mathbf{V}_{1},\mathbf{V}_{2}\in\mathcal{Q}, v1,v2,WLp(3)+L(3)v_{1},v_{2},W\in L^{p}({\mathbb{R}}^{3})+L^{\infty}({\mathbb{R}}^{3}) for some p>3/2p>3/2, and ϕ0x,ϕ0yL2(3)\phi_{0}^{x},\phi_{0}^{y}\in L^{2}({\mathbb{R}}^{3}). Then (1.2) has a unique solution (ϕx,ϕy)C(;L2(3))2Llocq(;Lr(3))2(\phi^{x},\phi^{y})\in C({\mathbb{R}};L^{2}({\mathbb{R}}^{3}))^{2}\cap L^{q}_{\rm loc}({\mathbb{R}};L^{r}({\mathbb{R}}^{3}))^{2}, where 1=2/r+1/p1=2/r+1/p and qq is such that

2q=3(121r).\frac{2}{q}=3\left(\frac{1}{2}-\frac{1}{r}\right).

The L2L^{2}-norms of ϕx\phi^{x} and ϕy\phi^{y} are independent of tt\in{\mathbb{R}}, hence in view of (1.3),

ϕx(t)L2(3)=ϕy(t)L2(3)=1,t.\|\phi^{x}(t)\|_{L^{2}({\mathbb{R}}^{3})}=\|\phi^{y}(t)\|_{L^{2}({\mathbb{R}}^{3})}=1,\quad\forall t\in{\mathbb{R}}.

The proof is presented shortly in Appendix B.

Remark 3.2.

The sign of ε\varepsilon in (3.1) plays no role here. Indeed, the proof relies on local in time Strichartz estimates associated to 𝐇x\mathbf{H}_{x} and 𝐇y\mathbf{H}_{y}, respectively, and the potentials v1,v2v_{1},v_{2} and WW are treated as perturbations, whose sign is irrelevant in order to guarantee the above global existence result. On the other hand, Theorem 3.1 brings no information regarding the quality of the dynamics or the existence of a ground state.

Remark 3.3.

Under extra assumptions on the potentials v1v_{1} and v2v_{2} (no extra assumption is needed for WW, as it is associated to a convolution), it is possible to consider higher regularity properties. In particular, working at the level of H1H^{1}-regularity makes it possible to show the conservation of the energy

E(t)\displaystyle E(t) =Hxϕx(t),ϕx(t)Lx2+Hyϕy(t),ϕy(t)Ly2\displaystyle=\langle H_{x}\phi^{x}(t),\phi^{x}(t)\rangle_{L^{2}_{x}}+\langle H_{y}\phi^{y}(t),\phi^{y}(t)\rangle_{L^{2}_{y}}
+3×3W(xy)|ϕx(t,x)|2|ϕy(t,y)|2𝑑x𝑑y,\displaystyle\quad+\iint_{{\mathbb{R}}^{3}\times{\mathbb{R}}^{3}}W(x-y)|\phi^{x}(t,x)|^{2}|\phi^{y}(t,y)|^{2}dxdy,

provided that v1\nabla v_{1} and v2\nabla v_{2} also belong to Lp(3)+L(3)L^{p}({\mathbb{R}}^{3})+L^{\infty}({\mathbb{R}}^{3}) for some p>3/2p>3/2. We refer to Remark B.5 for more details.

Remark 3.4.

The role of the set 𝒬\mathcal{Q} is to guarantee that (local in time) Strichartz estimates are available for 𝐇x\mathbf{H}_{x} and 𝐇y\mathbf{H}_{y}. The same would still be true for a larger class of potentials, including for instance Kato potentials ([20]) or potentials decaying like an inverse square ([5]). The choice of this set 𝒬\mathcal{Q} is made in order to simplify the presentation, and because it is delicate to keep track of all the classes of potentials for which Strichartz estimates have been proved.

3.2. Coupling potentials with polynomial growth

The core of this paper addresses the case where the coupling potential ww may grow polynomially. To be more concrete, we recall the example addressed in [4].

Example 3.5.

Assume d1=d2=1d_{1}=d_{2}=1 and that the potentials are given by

V1(x)=12x2(x21)2,>0,V2(y)=ω22y2,w(x,y)=χ(x)y2,χC0().V_{1}(x)=\frac{1}{2}x^{2}\left(\frac{x}{2\ell}-1\right)^{2},\quad\ell>0,\quad V_{2}(y)=\frac{\omega^{2}}{2}y^{2},\quad w(x,y)=\chi(x)y^{2},\;\;\chi\in C_{0}^{\infty}({\mathbb{R}}).

Here, V1(x)V_{1}(x) corresponds to a double well and V2(y)V_{2}(y) to a harmonic bath. The coupling w(x,y)w(x,y) could be locally cubic when choosing χ(x)=x\chi(x)=x for xx in a neighborhood of zero.

We emphasize that in Example 3.5, the average wx{\langle}w{\rangle}_{x} grows quadratically in yy: in terms of growth, wx{\langle}w{\rangle}_{x} is comparable to V2V_{2} and cannot be considered as a perturbation as far as the Cauchy problem is concerned. This setting turns out to be very different from the one in [7, 2] (see also Theorem 3.1), and requires a different approach to be developed below.

3.2.1. Restriction to non-negative potentials

In the general case, we assume d1,d21d_{1},d_{2}\geq 1. First, the potentials V1V_{1} and V2V_{2} are smooth, real-valued, V1C(d1;)V_{1}\in C^{\infty}({\mathbb{R}}^{d_{1}};{\mathbb{R}}), V2C(d2;)V_{2}\in C^{\infty}({\mathbb{R}}^{d_{2}};{\mathbb{R}}), and bounded from below:

xd1,yd2,V1(x)C1andV2(y)C2,\forall x\in{\mathbb{R}}^{d_{1}},\forall y\in{\mathbb{R}}^{d_{2}},\;\;V_{1}(x)\geq-C_{1}\;\;\mbox{and}\;\;V_{2}(y)\geq-C_{2},

for some constants C1,C2>0C_{1},C_{2}>0. The operators HxH_{x} and HyH_{y} then are self-adjoint operators. Up to changing ϕx(t,x)\phi^{x}(t,x) to ϕx(t,x)eitC1\phi^{x}(t,x)e^{itC_{1}} (which amounts to replacing V1V_{1} by V1+C1V_{1}+C_{1} in (1.2)), and ϕy(t,y)\phi^{y}(t,y) to ϕy(t,y)eitC2\phi^{y}(t,y)e^{itC_{2}}, we may actually assume

(H1) V1(x)1,xd1,andV2(y)1,yd2,V_{1}(x)\geq 1,\ \forall x\in{\mathbb{R}}^{d_{1}},\quad\text{and}\quad V_{2}(y)\geq 1,\ \forall y\in{\mathbb{R}}^{d_{2}},

as we are only interested in existence results for the Cauchy problem (1.2). Thus HxH_{x} and HyH_{y} are sums of a nonnegative operator (Laplacian in xx and yy, respectively) and of a nonnegative potential. We use them to measure the regularity of the solutions of the system (1.2).

3.2.2. Functional setting

For kk\in{\mathbb{N}}, we define the Hilbert spaces

xk\displaystyle\mathcal{H}_{x}^{k} ={ϕL2(d1),Hxk/2ϕL2(d1)}andyk={ϕL2(d2),Hyk/2ϕL2(d2)},\displaystyle=\left\{\phi\in L^{2}({\mathbb{R}}^{d_{1}}),\;\;H^{k/2}_{x}\phi\in L^{2}({\mathbb{R}}^{d_{1}})\right\}\;\;\mbox{and}\;\;\mathcal{H}_{y}^{k}=\left\{\phi\in L^{2}({\mathbb{R}}^{d_{2}}),\;\;H^{k/2}_{y}\phi\in L^{2}({\mathbb{R}}^{d_{2}})\right\},

which are the natural analogues of Sobolev spaces HkH^{k} in the presence of (nonnegative) potentials (in view of (H1)), equipped with the norms given by

ϕxk2=ϕLx22+Hxk/2ϕLx22,ϕyk2=ϕLy22+Hyk/2ϕLy22.\|\phi\|_{\mathcal{H}_{x}^{k}}^{2}=\|\phi\|_{L^{2}_{x}}^{2}+\|H^{k/2}_{x}\phi\|_{L^{2}_{x}}^{2},\quad\|\phi\|_{\mathcal{H}_{y}^{k}}^{2}=\|\phi\|_{L^{2}_{y}}^{2}+\|H^{k/2}_{y}\phi\|_{L^{2}_{y}}^{2}.

For α,β\alpha,\beta\in{\mathbb{N}}, Φ=(ϕx,ϕy)xα×yβ\Phi=(\phi^{x},\phi^{y})\in\mathcal{H}_{x}^{\alpha}\times\mathcal{H}_{y}^{\beta}, we set

Φα,β2=ϕxxα2+ϕyyβ2=ϕxLx22+Hxα/2ϕxLx22+ϕyLy22+Hyβ/2ϕyLy22.\|\Phi\|_{\alpha,\beta}^{2}=\|\phi^{x}\|_{\mathcal{H}_{x}^{\alpha}}^{2}+\|\phi^{y}\|_{\mathcal{H}_{y}^{\beta}}^{2}=\|\phi^{x}\|_{L^{2}_{x}}^{2}+\|H_{x}^{\alpha/2}\phi^{x}\|_{L^{2}_{x}}^{2}+\|\phi^{y}\|_{L^{2}_{y}}^{2}+\|H_{y}^{\beta/2}\phi^{y}\|_{L^{2}_{y}}^{2}.

All along the paper, we use that in view of (H1), 0HxαHxα+10\leq H_{x}^{\alpha}\leq H_{x}^{\alpha+1} and 0HyβHyβ+10\leq H_{y}^{\beta}\leq H_{y}^{\beta+1}.

As (1.2) is reversible, from now on we consider positive time only. We shall work with the time-dependent functional spaces

XTα,β={Φ(t)=(ϕx(t),ϕy(t)),ϕxL([0,T],xα),ϕyL([0,T],yβ)}.X^{\alpha,\beta}_{T}=\left\{\Phi(t)=(\phi^{x}(t),\phi^{y}(t)),\;\phi^{x}\in L^{\infty}\left([0,T],\mathcal{H}_{x}^{\alpha}\right),\;\phi^{y}\in L^{\infty}\left([0,T],\mathcal{H}_{y}^{\beta}\right)\right\}.

If Φ=(ϕx,ϕy)XTα,β\Phi=(\phi^{x},\phi^{y})\in X^{\alpha,\beta}_{T}, we set

ΦXTα,β=supt[0,T]Φ(t)α,β.\|\Phi\|_{X_{T}^{\alpha,\beta}}=\sup_{t\in[0,T]}\|\Phi(t)\|_{\alpha,\beta}.

We choose to consider integer exponents α\alpha and β\beta for the sake of simplicity. We emphasize however that our approach requires α,β2\alpha,\beta\geq 2; see Section 4 for a more precise discussion on this aspect. We note that Theorem 3.11 allows α=β=2\alpha=\beta=2.

3.2.3. Main assumptions

We assume that the coupling potential wC(d1+d2;)w\in C^{\infty}({\mathbb{R}}^{d_{1}+d_{2}};{\mathbb{R}}) satisfies

(H2) There exist c0,C>0c_{0},C>0 with c0<1c_{0}<1 such that for all (x,y)d1×d2(x,y)\in{\mathbb{R}}^{d_{1}}\times{\mathbb{R}}^{d_{2}},

|w(x,y)|c0(V1(x)+V2(y)+C).|w(x,y)|\leq c_{0}(V_{1}(x)+V_{2}(y)+C).

We emphasize that no condition is required concerning the above constant CC: for instance if V1V_{1} and V2V_{2} are bounded, then we may always pick c0<1c_{0}<1 so that (H2) is satisfied. For unbounded potentials, the requirement c0<1c_{0}<1 can be understood as some smallness property, in the sense that w(x,y)w(x,y) is a perturbation of V1(x)+V2(y)V_{1}(x)+V_{2}(y). This actually corresponds to the physical framework where the system (1.2) is introduced in order to approximate the exact solution ψ\psi of (1.1) through the formula ψϕxϕy\psi\approx\phi^{x}\otimes\phi^{y}; see [3] for a derivation of error estimates. We also note that the assumption c0<1c_{0}<1 implies that ww is (Hx+Hy)(H_{x}+H_{y})-bounded with relative bound c0<1c_{0}<1, hence by Kato–Rellich Theorem (see e.g. [18, Theorem X.12]), HH is self-adjoint.

Remark 3.6.

If ww is at most quadratic, in the sense that wC(d1+d2;)w\in C^{\infty}({\mathbb{R}}^{d_{1}+d_{2}};{\mathbb{R}})

x,yγwL(d1+d2),γd1+d2,|γ|2,{\partial}_{x,y}^{\gamma}w\in L^{\infty}({\mathbb{R}}^{d_{1}+d_{2}}),\quad\forall\gamma\in{\mathbb{N}}^{d_{1}+d_{2}},\ |\gamma|\geq 2,

then the assumption (H2) is not needed to guarantee that HH is self-adjoint (see the Faris–Lavine Theorem, [18, Theorem X.38]). Such a framework corresponds to the assumptions made for the error analysis in [3, 4]. For such potentials ww, the assumption c0<1c_{0}<1 in (H2) is needed only in order to ensure that the Hartree solutions are global in time.

We also assume some conditions on the regularity of commutators of the coupling potential with the operators HxH_{x} and HyH_{y}. For integers α,β1\alpha,\beta\geq 1, we consider the condition:

(H3)α,β. There exist c1,c2>0c_{1},c_{2}>0 such that for all k{1,,α}k\in\{1,\cdots,\alpha\}, {1,,β}\ell\in\{1,\cdots,\beta\}, for all fj=fj(x)f_{j}=f_{j}(x), gj=gj(y)g_{j}=g_{j}(y) in the Schwartz class (j{1,2}j\in\{1,2\}),

|Hxk1[w(,y),Hx]f1,f2Lx2|\displaystyle\left|{\langle}H_{x}^{k-1}[w(\cdot,y),H_{x}]f_{1},f_{2}{\rangle}_{L^{2}_{x}}\right| +|[w(,y),Hx]Hxk1f1,f2Lx2|\displaystyle+\left|{\langle}[w(\cdot,y),H_{x}]H_{x}^{k-1}f_{1},f_{2}{\rangle}_{L^{2}_{x}}\right|
c1(1+V2(y))f1xkf2xk,for a.a. yd2,\displaystyle\leq c_{1}\left(1+V_{2}(y)\right)\|f_{1}\|_{\mathcal{H}_{x}^{k}}\|f_{2}\|_{\mathcal{H}_{x}^{k}},\quad\text{for a.a. }y\in{\mathbb{R}}^{d_{2}},
|Hy1[w(x,),Hy]g1,g2Ly2|\displaystyle\left|{\langle}H_{y}^{\ell-1}[w(x,\cdot),H_{y}]g_{1},g_{2}{\rangle}_{L^{2}_{y}}\right| +|[w(x,),Hy]Hy1g1,g2Ly2|\displaystyle+\left|{\langle}[w(x,\cdot),H_{y}]H_{y}^{\ell-1}g_{1},g_{2}{\rangle}_{L^{2}_{y}}\right|
c2(1+V1(x))g1yg2y,for a.a. xd1.\displaystyle\leq c_{2}\left(1+V_{1}(x)\right)\|g_{1}\|_{\mathcal{H}_{y}^{\ell}}\|g_{2}\|_{\mathcal{H}_{y}^{\ell}},\quad\text{for a.a. }x\in{\mathbb{R}}^{d_{1}}.

Assumption (H3)α,β is made in order to generalize the framework of Example 3.5. The subsequent proofs do not use the special form of the Hamiltonians HxH_{x}, HyH_{y}, so that our result extends as soon as they are self-adjoint operators and assumptions (H1), (H2), (H3)α,β are satisfied. This applies in particular for magnetic Schrödinger operators.

Remark 3.7.

It is not necessary to assume that the potential ww is smooth, wC(d1+d2;)w\in C^{\infty}({\mathbb{R}}^{d_{1}+d_{2}};{\mathbb{R}}). We only need enough regularity in order to write assumption (𝐇𝟑)α,β{\bf(H3)}_{\alpha,\beta} for the α\alpha and β\beta that we consider (recalling that Hxk1H_{x}^{k-1} and Hy1H_{y}^{\ell-1} are self-adjoint). For example, if wC2(d1+d2;)w\in C^{2}({\mathbb{R}}^{d_{1}+d_{2}};{\mathbb{R}}), then Theorem 3.11 holds with α=β=2\alpha=\beta=2. We make this regularity assumption for simplicity, as the most important properties are those discussed in this subsection.

Remark 3.8.

Whenever w(x,y)w(x,y) is a Coulomb potential as in Section 3.1, the assumptions (H3)α,β are not satisfied. One then needs to take advantage of the convolution feature of the coupling and of the properties of HxH_{x} and HyH_{y} such as the Strichartz estimates in Proposition B.2.

We next present sufficient conditions on the potentials guaranteeing that assumptions (H2) and (H3)2,2 hold.

Lemma 3.9 (Sufficient conditions).

Let V1C(d1;)V_{1}\in C^{\infty}({\mathbb{R}}^{d_{1}};{\mathbb{R}}) and V2C(d2;)V_{2}\in C^{\infty}({\mathbb{R}}^{d_{2}};{\mathbb{R}}) such that V1,V20V_{1},V_{2}\geq 0. The above assumptions (H2) and (H3)2,2 are satisfied provided that the following estimates hold:

  • There exists C>0C>0 such that

    (3.2) |V1(x)|C(1+V1(x)),xd1;|V2(y)|C(1+V2(y)),yd2.|\nabla V_{1}(x)|\leq C\left(1+V_{1}(x)\right),\forall x\in{\mathbb{R}}^{d_{1}};\quad|\nabla V_{2}(y)|\leq C\left(1+V_{2}(y)\right),\forall y\in{\mathbb{R}}^{d_{2}}.
  • There exist 0<c0<10<c_{0}<1 and c>0c>0 independent of xd1x\in{\mathbb{R}}^{d_{1}} and yd2y\in{\mathbb{R}}^{d_{2}} such that

    (3.3) {|w(x,y)|c0(V1(x)+V2(y)+c),|xw(x,y)|c(V1(x)+V2(y)+1),|yw(x,y)|c(V1(x)+V2(y)+1),|Δxw(x,y)|+|Δyw(x,y)|c(V1(x)+V2(y)+1).\begin{cases}&|w(x,y)|\leq c_{0}(V_{1}(x)+V_{2}(y)+c),\\ &|\nabla_{x}w(x,y)|\leq c(\sqrt{V_{1}(x)}+V_{2}(y)+1),\\ &|\nabla_{y}w(x,y)|\leq c(V_{1}(x)+\sqrt{V_{2}(y)}+1),\\ &|\Delta_{x}w(x,y)|+|\Delta_{y}w(x,y)|\leq c(V_{1}(x)+V_{2}(y)+1).\end{cases}

The proof of this lemma is given in Section 7.

Remark 3.10.

We note that Example 3.5 meets the requirements stated in Lemma 3.9, provided that χL()<ω2/2\|\chi\|_{L^{\infty}({\mathbb{R}})}<\omega^{2}/2. Thus it satisfies the assumptions of Theorem 3.11 below for α=β=2\alpha=\beta=2.

3.3. Main result and comments

Before stating our main result we informally summarize the previous assumptions on the potentials for the case of polynomial coupling:

(𝐇𝟏):boundedness from below of the potentials V1(x) and V2(y);\displaystyle{\bf(H1)}:\text{boundedness from below of the potentials $V_{1}(x)$ and $V_{2}(y)$;}
(𝐇𝟐):control of w(x,y) in terms of V1(x)+V2(y);\displaystyle{\bf(H2)}:\text{control of $w(x,y)$ in terms of $V_{1}(x)+V_{2}(y)$;}
(𝐇𝟑)α,β:control of commutators involving w(x,y), in terms of Hx and Hy.\displaystyle{\bf(H3)}_{\alpha,\beta}:\text{control of commutators involving $w(x,y)$, in terms of $H_{x}$ and $H_{y}$.}

We have the following result on existence and uniqueness as well norm and energy conservation of the time-dependent Hartree approximation.

Theorem 3.11.

Let d1,d21d_{1},d_{2}\geq 1, α,β2\alpha,\beta\geq 2 and ϕ0xxα\phi_{0}^{x}\in\mathcal{H}_{x}^{\alpha}, ϕ0yyβ\phi_{0}^{y}\in\mathcal{H}_{y}^{\beta}. Suppose that (H1), (H2) and (H3)α,β are satisfied.

  • (1.2) possesses a unique, global solution in ΦC(+;L2×L2)T>0XTα,β\Phi\in C({\mathbb{R}}_{+};L^{2}\times L^{2})\cap\bigcap_{T>0}X^{\alpha,\beta}_{T}.

  • Conservations: the L2L^{2}-norms of ϕx\phi^{x} and ϕy\phi^{y} are independent of t0t\geq 0, hence in view of (1.3),

    ϕx(t)L2(d1)=ϕy(t)L2(d2)=1,t0.\|\phi^{x}(t)\|_{L^{2}({\mathbb{R}}^{d_{1}})}=\|\phi^{y}(t)\|_{L^{2}({\mathbb{R}}^{d_{2}})}=1,\quad\forall t\geq 0.

    In addition, the following total energy is also independent of t0t\geq 0:

    E(t)\displaystyle E(t) :=Hxϕx(t),ϕx(t)Lx2+Hyϕy(t),ϕy(t)Ly2\displaystyle:=\langle H_{x}\phi^{x}(t),\phi^{x}(t)\rangle_{L^{2}_{x}}+\langle H_{y}\phi^{y}(t),\phi^{y}(t)\rangle_{L^{2}_{y}}
    +d1×d2w(x,y)|ϕx(t,x)|2|ϕy(t,y)|2𝑑x𝑑y.\displaystyle\quad+\iint_{{\mathbb{R}}^{d_{1}}\times{\mathbb{R}}^{d_{2}}}w(x,y)|\phi^{x}(t,x)|^{2}|\phi^{y}(t,y)|^{2}dxdy.

We will see that the assumption c0<1c_{0}<1 in (H2) arises in two steps of the proof of Theorem 3.11. First, to make sure that the approximating scheme (1.4) introduced below is well-defined, we invoke Kato–Rellich Theorem, to show essentially that wy{\langle}w{\rangle}_{y} (or more precisely, wny{\langle}w_{n}{\rangle}_{y}) is HxH_{x}-bounded with relative bound smaller than one, and that the same holds when the roles of xx and yy are swapped. Second, the assumption c0<1c_{0}<1 guarantees that the conserved energy EE, defined in Theorem 3.11, is a coercive functional, so the conservation of EE provides uniform in times a priori estimates, which in turn allow to show that the local in time solutions are actually global in time solutions.

The property ΦC(+;L2×L2)T>0XTα,β\Phi\in C({\mathbb{R}}_{+};L^{2}\times L^{2})\cap\bigcap_{T>0}X^{\alpha,\beta}_{T} means that tΦ(t)α,βt\mapsto\|\Phi(t)\|_{\alpha,\beta} is locally bounded on +{\mathbb{R}}_{+}. The map tΦ(t)1,1t\mapsto\|\Phi(t)\|_{1,1} is bounded on +{\mathbb{R}}_{+} in view of the conservation of the coercive energy EE, but higher order norms may not be bounded as tt goes to infinity (recall that Theorem 3.11 requires α,β2\alpha,\beta\geq 2).

4. Analysis of the iterative scheme: existence and uniform bounds

This section is devoted to the analysis of the system (1.4). For nn\in{\mathbb{N}}, we denote by Φn=(ϕnx,ϕny)\Phi_{n}=(\phi_{n}^{x},\phi_{n}^{y}), the solution to the scheme (1.4) and we prove local in time uniform estimates. At this stage, we only need that ϕ0xxα\phi_{0}^{x}\in\mathcal{H}_{x}^{\alpha}, ϕ0yyβ\phi_{0}^{y}\in\mathcal{H}_{y}^{\beta} for integers α,β1\alpha,\beta\geq 1.

Lemma 4.1.

Let α,β1\alpha,\beta\geq 1. Assume that (H1), (H2) and (H3)α,β are satisfied. Assume ϕ0xxα\phi_{0}^{x}\in\mathcal{H}_{x}^{\alpha}, and ϕ0yyβ\phi_{0}^{y}\in\mathcal{H}_{y}^{\beta}. Then, the sequence (Φn)n(\Phi_{n})_{n\in{\mathbb{N}}} solution to (1.4) is well-defined and there exists T>0T>0 such that for all nn\in{\mathbb{N}}, the solution ΦnXTα,β\Phi_{n}\in X_{T}^{\alpha,\beta} of the scheme (1.4) satisfies

(4.1) ΦnXTα,β2Φ0α,β.\|\Phi_{n}\|_{X_{T}^{\alpha,\beta}}\leq 2\|\Phi_{0}\|_{\alpha,\beta}.

The proof of this lemma relies on the fact that the control of Φn+1\Phi_{n+1} involves terms which are linear in Φn+1\Phi_{n+1} and quadratic in Φn\Phi_{n}, and require the XT1,1X_{T}^{1,1}-norm of Φn\Phi_{n}. For this reason, the Lemma holds as soon as α,β1\alpha,\beta\geq 1. However in Theorem 3.11 , we require at least an XT2,2X^{2,2}_{T} regularity. The reason will appear in Section 5, as we do need uniform (in nn) estimates in XT2,2X^{2,2}_{T} to show that the sequence (Φn)n(\Phi_{n})_{n} converges in XT0,0=LTL2X^{0,0}_{T}=L^{\infty}_{T}L^{2}.

In Section 4.1, we address the construction of the family (Φn)n(\Phi_{n})_{n\in{\mathbb{N}}}, which relies on a commutation lemma that we prove in Section 4.3. Section 4.2 is devoted to the proof of the uniform bound stated in Lemma 4.1.

4.1. Well-posedness of the scheme

Before entering into the proof of Lemma 4.1, let us discuss why the scheme is indeed well-defined: as Φn+1\Phi_{n+1} solves a decoupled system of linear Schrödinger equations, it suffices to study the properties of the time-dependent potentials wny{\langle}w_{n}{\rangle}_{y} and wnx{\langle}w_{n}{\rangle}_{x}. We fix T>0T>0 arbitrary and take ϕ0xxα\phi_{0}^{x}\in\mathcal{H}_{x}^{\alpha}, and ϕ0yyβ\phi_{0}^{y}\in\mathcal{H}_{y}^{\beta}, with α,β1\alpha,\beta\geq 1. For n=0n=0, Φ0\Phi_{0} is obviously well-defined with Φ0XTα,β\Phi_{0}\in X_{T}^{\alpha,\beta}, and

(4.2) ϕnx(t)Lx2=ϕny(t)Ly2=1,t,\|\phi_{n}^{x}(t)\|_{L^{2}_{x}}=\|\phi_{n}^{y}(t)\|_{L^{2}_{y}}=1,\quad\forall t\in{\mathbb{R}},

holds for n=0n=0. We argue by induction. If ΦnXT1,1\Phi_{n}\in X_{T}^{1,1} satisfies (4.2), then in view of (H2), wny(t,x)\langle w_{n}\rangle_{y}(t,x) and wnx(t,y)\langle w_{n}\rangle_{x}(t,y) are well-defined. In addition, for t[0,T]t\in[0,T], (H2) yields

(4.3) |wny(t,x)|c0V1(x)ϕny(t)L22+Cϕny(t)y12, a.e. x,\displaystyle|\langle w_{n}\rangle_{y}(t,x)|\leq c_{0}V_{1}(x)\|\phi^{y}_{n}(t)\|_{L^{2}}^{2}+C\|\phi_{n}^{y}(t)\|^{2}_{\mathcal{H}_{y}^{1}},\quad\text{ a.e. }x,
|wnx(t,y)|c0V2(y)ϕnx(t)L22+Cϕnx(t)x12,a.e. y,\displaystyle|\langle w_{n}\rangle_{x}(t,y)|\leq c_{0}V_{2}(y)\|\phi^{x}_{n}(t)\|^{2}_{L^{2}}+C\|\phi_{n}^{x}(t)\|^{2}_{\mathcal{H}_{x}^{1}},\quad\text{a.e. }y,

for some constant CC whose value is irrelevant here, unlike the fact that we assume c0<1c_{0}<1. Indeed, together with (4.2), this implies that wny\langle w_{n}\rangle_{y} is HxH_{x}-bounded with relative bound at most c0c_{0}. By Kato–Rellich Theorem (see e.g. [18, Theorem X.12]), Φn+1XT0,0\Phi_{n+1}\in X_{T}^{0,0} is well-defined (see e.g. [19, Section VIII.4]), and (4.2) holds at level n+1n+1. Next, we prove that Φn+1XT1,1\Phi_{n+1}\in X_{T}^{1,1}. Applying the operator HxH_{x} to the first equation in (1.4), we find

(4.4) (itHx)(Hxϕn+1x)=wny(t)(Hxϕn+1x)+[Hx,wny(t)]ϕn+1x.(i{\partial}_{t}-H_{x})(H_{x}\phi_{n+1}^{x})=\langle w_{n}\rangle_{y}(t)(H_{x}\phi_{n+1}^{x})+[H_{x},\langle w_{n}\rangle_{y}(t)]\phi^{x}_{n+1}.

Since HxH_{x} is self-adjoint, we deduce

Hx1/2ϕn+1x(t)Lx22\displaystyle\|H^{1/2}_{x}\phi^{x}_{n+1}(t)\|_{L^{2}_{x}}^{2} =ReHxϕn+1x(t),ϕn+1x(t)Lx2\displaystyle={\rm Re}{\langle}H_{x}\phi^{x}_{n+1}(t),\phi^{x}_{n+1}(t){\rangle}_{L^{2}_{x}}
=Hx1/2ϕ0xLx22+Re(0tddsHxϕn+1x(s),ϕn+1x(s)Lx2𝑑s)\displaystyle=\|H^{1/2}_{x}\phi^{x}_{0}\|_{L^{2}_{x}}^{2}+{\rm Re}\left(\int_{0}^{t}\frac{d}{ds}{\langle}H_{x}\phi^{x}_{n+1}(s),\phi^{x}_{n+1}(s){\rangle}_{L^{2}_{x}}ds\right)
=Hx1/2ϕ0xLx22Re(0ti[Hx,wny]ϕn+1x(s),ϕn+1x(s)Lx2𝑑s).\displaystyle=\|H^{1/2}_{x}\phi^{x}_{0}\|_{L^{2}_{x}}^{2}-{\rm Re}\left(\int_{0}^{t}{\langle}i[H_{x},\langle w_{n}\rangle_{y}]\phi^{x}_{n+1}(s),\phi^{x}_{n+1}(s){\rangle}_{L^{2}_{x}}ds\right).

Minkowski inequality yields, in view of (H3)1,1,

(4.5) |[Hx,wny(t)]f1,f2Lx2|ϕnyLTy12f1x1f2x1.\left|{\langle}[H_{x},\langle w_{n}\rangle_{y}(t)]f_{1},f_{2}{\rangle}_{L^{2}_{x}}\right|\lesssim\|\phi_{n}^{y}\|_{L^{\infty}_{T}\mathcal{H}^{1}_{y}}^{2}\|f_{1}\|_{\mathcal{H}_{x}^{1}}\|f_{2}\|_{\mathcal{H}_{x}^{1}}.

We infer the existence of a universal constant C>0C>0 such that

Hx1/2ϕn+1x(t)Lx22\displaystyle\|H^{1/2}_{x}\phi^{x}_{n+1}(t)\|_{L^{2}_{x}}^{2} Hx1/2ϕ0xLx22+CΦnXT1,120tϕn+1x(s)x12𝑑s.\displaystyle\leq\|H^{1/2}_{x}\phi^{x}_{0}\|_{L^{2}_{x}}^{2}+C\|\Phi_{n}\|_{X^{1,1}_{T}}^{2}\int_{0}^{t}\|\phi^{x}_{n+1}(s)\|^{2}_{\mathcal{H}_{x}^{1}}ds.

We deduce

(4.6) supt[0,T]ϕn+1x(t)x12ϕ0xx12+CΦnXT1,120tϕn+1x(s)x12𝑑s,\sup_{t\in[0,T]}\|\phi^{x}_{n+1}(t)\|^{2}_{\mathcal{H}_{x}^{1}}\lesssim\|\phi^{x}_{0}\|^{2}_{\mathcal{H}_{x}^{1}}+C\|\Phi_{n}\|_{X^{1,1}_{T}}^{2}\int_{0}^{t}\|\phi^{x}_{n+1}(s)\|^{2}_{\mathcal{H}_{x}^{1}}ds,

We have a similar estimate for Hy1/2ϕn+1y(t)Lx22\|H^{1/2}_{y}\phi^{y}_{n+1}(t)\|_{L^{2}_{x}}^{2}:

(4.7) supt[0,T]ϕn+1y(t)y12ϕ0yy12+CΦnXT1,120tϕn+1y(s)y12𝑑s,\sup_{t\in[0,T]}\|\phi^{y}_{n+1}(t)\|^{2}_{\mathcal{H}_{y}^{1}}\lesssim\|\phi^{y}_{0}\|^{2}_{\mathcal{H}_{y}^{1}}+C\|\Phi_{n}\|_{X^{1,1}_{T}}^{2}\int_{0}^{t}\|\phi^{y}_{n+1}(s)\|^{2}_{\mathcal{H}_{y}^{1}}ds,

and so Gronwall lemma and the inductive assumption yields Φn+1XT1,1\Phi_{n+1}\in X_{T}^{1,1} and completes the construction of the sequence (Φn)n(\Phi_{n})_{n\in{\mathbb{N}}}.

4.2. Uniform bounds

We conclude the proof of Lemma 4.1 in analyzing the regularity of the solutions.

Proof of Lemma 4.1.

In view of the definition of the scheme and of the conservations

ddtϕnxL22=ddtϕnyL22=0,\frac{d}{dt}\|\phi_{n}^{x}\|_{L^{2}}^{2}=\frac{d}{dt}\|\phi_{n}^{y}\|_{L^{2}}^{2}=0,

we need now consider HxαϕnxH_{x}^{\alpha}\phi^{x}_{n} and HyβϕnyH_{y}^{\beta}\phi^{y}_{n} for α,β1\alpha,\beta\geq 1. Let

R=2Φ0α,β,R=2\|\Phi_{0}\|_{\alpha,\beta},

and introduce

BR,T={ΦXTα,β,ΦXTα,βR}.B_{R,T}=\{\Phi\in X_{T}^{\alpha,\beta},\ \|\Phi\|_{X_{T}^{\alpha,\beta}}\leq R\}.

We distinguish two cases for the ease of presentation.

First case: α=β=1\alpha=\beta=1. In that case, if ΦnBR,T\Phi_{n}\in B_{R,T}, then estimates (4.6) and (4.7) imply

Φn+1(t)XT1,12Φ01,12+CTR2Φn+1(t)XT1,12.\|\Phi_{n+1}(t)\|_{X_{T}^{1,1}}^{2}\leq\|\Phi_{0}\|_{1,1}^{2}+CTR^{2}\|\Phi_{n+1}(t)\|_{X_{T}^{1,1}}^{2}.

We infer that choosing T>0T>0 sufficiently small in terms of RR, but independently of nn, ΦnBR,T\Phi_{n}\in B_{R,T} implies Φn+1BR,T\Phi_{n+1}\in B_{R,T}.

Higher regularity: The control of higher order regularity is obtained by a similar recursive argument which uses an iterated commutator estimate. Let α,β1\alpha,\beta\geq 1. We have

(4.8) {itHxkϕn+1x=(Hx+wny)Hxkϕn+1x+[Hxk,wny]ϕn+1x,Hxkϕn+1t=0x=Hxkϕ0x,itHyϕn+1y=(Hy+wnx)Hyβϕn+1y+[Hy,wnx]ϕn+1y,Hyϕn+1t=0y=Hyϕ0y.\left\{\begin{aligned} i\partial_{t}H_{x}^{k}\phi^{x}_{n+1}=(H_{x}+\langle w_{n}\rangle_{y})H_{x}^{k}\phi^{x}_{n+1}+[H^{k}_{x},\langle w_{n}\rangle_{y}]\phi^{x}_{n+1},\quad H_{x}^{k}\phi^{x}_{n+1\mid t=0}=H_{x}^{k}\phi^{x}_{0},\\ i\partial_{t}H_{y}^{\ell}\phi^{y}_{n+1}=(H_{y}+\langle w_{n}\rangle_{x})H_{y}^{\beta}\phi^{y}_{n+1}+[H_{y}^{\ell},\langle w_{n}\rangle_{x}]\phi^{y}_{n+1},\quad H_{y}^{\ell}\phi^{y}_{n+1\mid t=0}=H_{y}^{\ell}\phi^{y}_{0}.\end{aligned}\right.

The next lemma allows to control the commutators.

Lemma 4.2.

Let ΦnXT1,1\Phi_{n}\in X^{1,1}_{T} for some T>0T>0, and k,1k,\ell\geq 1 be integers. Suppose that (H3)k,ℓ is satisfied. For all t[0,T]t\in[0,T],

|[Hxk,wny(t)]f1,f2Lx2|\displaystyle\left|{\langle}[H_{x}^{k},\langle w_{n}\rangle_{y}(t)]f_{1},f_{2}{\rangle}_{L^{2}_{x}}\right| ϕnyLTy12f1xkf2xk,f1,f2xk\displaystyle\lesssim\|\phi_{n}^{y}\|^{2}_{L^{\infty}_{T}\mathcal{H}^{1}_{y}}\|f_{1}\|_{\mathcal{H}^{k}_{x}}\|f_{2}\|_{\mathcal{H}^{k}_{x}},\qquad\forall f_{1},f_{2}\in\mathcal{H}_{x}^{k}
|[Hy,wnx(t)]g1,g2Ly2|\displaystyle\left|{\langle}[H_{y}^{\ell},\langle w_{n}\rangle_{x}(t)]g_{1},g_{2}{\rangle}_{L^{2}_{y}}\right| ϕnxLTx12g1yg2y,g1,g2y.\displaystyle\lesssim\|\phi_{n}^{x}\|^{2}_{L^{\infty}_{T}\mathcal{H}^{1}_{x}}\|g_{1}\|_{\mathcal{H}^{\ell}_{y}}\|g_{2}\|_{\mathcal{H}^{\ell}_{y}},\qquad\forall g_{1},g_{2}\in\mathcal{H}_{y}^{\ell}.

Taking the lemma for granted, (4.8) implies, since HxH_{x} is self-adjoint,

ϕn+1x(t)xα2\displaystyle\|\phi_{n+1}^{x}(t)\|_{\mathcal{H}^{\alpha}_{x}}^{2} =ReHxαϕn+1x(t),ϕn+1x(t)Lx2\displaystyle={\rm Re}{\langle}H_{x}^{\alpha}\phi^{x}_{n+1}(t),\phi^{x}_{n+1}(t){\rangle}_{L^{2}_{x}}
=Hxα/2ϕ0xLx22+Re(0tddsHxαϕn+1x(s),ϕn+1x(s)Lx2𝑑s)\displaystyle=\|H^{\alpha/2}_{x}\phi^{x}_{0}\|_{L^{2}_{x}}^{2}+{\rm Re}\left(\int_{0}^{t}\frac{d}{ds}{\langle}H_{x}^{\alpha}\phi^{x}_{n+1}(s),\phi^{x}_{n+1}(s){\rangle}_{L^{2}_{x}}ds\right)
=Hxα/2ϕ0xLx22Re(0ti[Hxα,wny]ϕn+1x(s),ϕn+1x(s)Lx2𝑑s)\displaystyle=\|H^{\alpha/2}_{x}\phi^{x}_{0}\|_{L^{2}_{x}}^{2}-{\rm Re}\left(\int_{0}^{t}{\langle}i[H_{x}^{\alpha},\langle w_{n}\rangle_{y}]\phi^{x}_{n+1}(s),\phi^{x}_{n+1}(s){\rangle}_{L^{2}_{x}}ds\right)
Hxα/2ϕ0xLx22+CTϕnyLTy12supt[0,T]ϕn+1x(t)xα2.\displaystyle\leq\|H_{x}^{\alpha/2}\phi_{0}^{x}\|_{L^{2}_{x}}^{2}+CT\|\phi^{y}_{n}\|^{2}_{L^{\infty}_{T}\mathcal{H}^{1}_{y}}\sup_{t\in[0,T]}\|\phi_{n+1}^{x}(t)\|_{\mathcal{H}^{\alpha}_{x}}^{2}.

We have a similar estimate for Hyβ/2ϕn+1y(t)Lx22\|H^{\beta/2}_{y}\phi^{y}_{n+1}(t)\|_{L^{2}_{x}}^{2}, and so if ΦnBR,T\Phi_{n}\in B_{R,T}, then equations (4.6) and (4.7) imply

Φn+1(t)XTα,β2Φ0α,β2+CTR2Φn+1(t)XTα,β2.\|\Phi_{n+1}(t)\|_{X_{T}^{\alpha,\beta}}^{2}\leq\|\Phi_{0}\|_{\alpha,\beta}^{2}+CTR^{2}\|\Phi_{n+1}(t)\|_{X_{T}^{\alpha,\beta}}^{2}.

We infer that choosing T>0T>0 sufficiently small in terms of RR, but independently of nn, ΦnBR,T\Phi_{n}\in B_{R,T} implies Φn+1BR,T\Phi_{n+1}\in B_{R,T}. It thus remains to prove the lemma, which is the subject of the next subsection. ∎

Remark 4.3.

Lemma 4.1 holds as soon as α,β1\alpha,\beta\geq 1, but this is not enough in order to conclude that the sequence (Φn)n(\Phi_{n})_{n\in{\mathbb{N}}} converges to some solution of (1.2). Indeed, the mere boundedness in XT1,1X_{T}^{1,1} only implies the convergence of a subsequence in the weak-* topology: this is not enough to pass to the limit in (1.4), both because the subsequence need not retain consecutive indices, and because the topology considered is too large to pass to the limit in nonlinear terms. These issues are overcome by requiring α,β2\alpha,\beta\geq 2 in Sections 5 and 6.

4.3. Proof of Lemma 4.2

Of course, (4.5) implies the result when k=1k=1. Take k1k\geq 1 and assume that the result holds for all mkm\leq k. We write

[Hxk+1,wny(t)]\displaystyle[H_{x}^{k+1},\langle w_{n}\rangle_{y}(t)] =Hxk[Hx,wny(t)]+[Hxk,wny(t)]Hx\displaystyle=H_{x}^{k}[H_{x},\langle w_{n}\rangle_{y}(t)]+[H_{x}^{k},\langle w_{n}\rangle_{y}(t)]H_{x}
=Hx[Hk1,wny(t)]Hx+Hxk[Hx,wny(t)]+[Hx,wny(t)]Hxk.\displaystyle=H_{x}[H^{k-1},\langle w_{n}\rangle_{y}(t)]H_{x}+H_{x}^{k}[H_{x},\langle w_{n}\rangle_{y}(t)]+[H_{x},\langle w_{n}\rangle_{y}(t)]H_{x}^{k}.

We deduce from (4.5) and the recursive assumption that for f1,f2𝒮(d1)f_{1},f_{2}\in\mathcal{{\mathcal{S}}}({\mathbb{R}}^{d_{1}}), we have

|[Hxk+1,wny(t)]f1,f2Lx2|\displaystyle\left|{\langle}[H_{x}^{k+1},{\langle}w_{n}{\rangle}_{y}(t)]f_{1},f_{2}{\rangle}_{L^{2}_{x}}\right| |[Hxk1,wny(t)]Hxf1,Hxf2Lx2|\displaystyle\leq\left|{\langle}[H_{x}^{k-1},{\langle}w_{n}{\rangle}_{y}(t)]H_{x}f_{1},H_{x}f_{2}{\rangle}_{L^{2}_{x}}\right|
+|Hxk[Hx,wny(t)]f1,f2Lx2|+|[Hx,wny(t)]Hxkf,gLx2|.\displaystyle\quad+\left|{\langle}H_{x}^{k}[H_{x},{\langle}w_{n}{\rangle}_{y}(t)]f_{1},f_{2}{\rangle}_{L^{2}_{x}}\right|+\left|{\langle}[H_{x},{\langle}w_{n}{\rangle}_{y}(t)]H_{x}^{k}f,g{\rangle}_{L^{2}_{x}}\right|.

By the recursive assumption

|[Hxk1,wny(t)]Hxf1,Hxf2Lx2|\displaystyle\left|{\langle}[H_{x}^{k-1},{\langle}w_{n}{\rangle}_{y}(t)]H_{x}f_{1},H_{x}f_{2}{\rangle}_{L^{2}_{x}}\right| ϕnyLTy12Hxf1xk1Hxf2xk1\displaystyle\lesssim\|\phi^{y}_{n}\|_{L^{\infty}_{T}\mathcal{H}_{y}^{1}}^{2}\|H_{x}f_{1}\|_{\mathcal{H}^{k-1}_{x}}\|H_{x}f_{2}\|_{\mathcal{H}^{k-1}_{x}}
ϕnyLTy12f1xk+1f2xk+1.\displaystyle\lesssim\|\phi^{y}_{n}\|_{L^{\infty}_{T}\mathcal{H}_{y}^{1}}^{2}\|f_{1}\|_{\mathcal{H}^{k+1}_{x}}\|f_{2}\|_{\mathcal{H}^{k+1}_{x}}.

Finally, in view of (H3)k,ℓ and Minkowski inequality, we have

|Hxk[Hx,wny(t)]f1,f2Lx2|\displaystyle\left|{\langle}H_{x}^{k}[H_{x},{\langle}w_{n}{\rangle}_{y}(t)]f_{1},f_{2}{\rangle}_{L^{2}_{x}}\right| ϕnyLTy12f1xk+1f2xk+1,\displaystyle\lesssim\|\phi_{n}^{y}\|_{L^{\infty}_{T}\mathcal{H}^{1}_{y}}^{2}\|f_{1}\|_{\mathcal{H}_{x}^{k+1}}\|f_{2}\|_{\mathcal{H}_{x}^{k+1}},
|[Hx,wny(t)]Hxkf1,f2Lx2|\displaystyle\left|{\langle}[H_{x},{\langle}w_{n}{\rangle}_{y}(t)]H_{x}^{k}f_{1},f_{2}{\rangle}_{L^{2}_{x}}\right| ϕnyLTy12f1xk+1f2xk+1,\displaystyle\lesssim\|\phi_{n}^{y}\|_{L^{\infty}_{T}\mathcal{H}^{1}_{y}}^{2}\|f_{1}\|_{\mathcal{H}_{x}^{k+1}}\|f_{2}\|_{\mathcal{H}_{x}^{k+1}},

which concludes the proof, after arguing similarly with HyH_{y}.

5. Convergence in small norms

The second step of the proof of Theorem 3.11 consists in passing to the limit n+n\rightarrow+\infty and prove the existence of a limit to the sequence (Φn)n(\Phi_{n})_{n\in{\mathbb{N}}} of solutions to (1.4). The main result in this section is:

Lemma 5.1.

Assume that there exist T>0T>0 and R>0R>0 such that

supnΦnXT2,2R.\sup_{n\in{\mathbb{N}}}\|\Phi_{n}\|_{X_{T}^{2,2}}\leq R.

Then there exist T1]0,T]T_{1}\in]0,T] and ΦXT12,2\Phi\in X_{T_{1}}^{2,2} such that

(5.1) sup0tT1Φn(t)Φ(t)Lx2×Ly2=ΦnΦXT10,0n0.\sup_{0\leq t\leq T_{1}}\|\Phi_{n}(t)-\Phi(t)\|_{L^{2}_{x}\times L^{2}_{y}}=\|\Phi_{n}-\Phi\|_{X^{0,0}_{T_{1}}}\mathop{\longrightarrow}\limits_{n\rightarrow\infty}0.

If in addition (Φn)n(\Phi_{n})_{n} is bounded in XTα,βX_{T}^{\alpha,\beta} for some integers α,β2\alpha,\beta\geq 2, then ΦXT1α,β\Phi\in X_{T_{1}}^{\alpha,\beta}.

Proof.

Consider (1.4) at steps n+1n+1 and nn, respectively, and subtract the corresponding equations. We find, for n1n\geq 1,

(itHx)(ϕn+1xϕnx)\displaystyle\left(i{\partial}_{t}-H_{x}\right)\left(\phi_{n+1}^{x}-\phi_{n}^{x}\right) =wnyϕn+1xwn1yϕnx\displaystyle={\langle}w_{n}{\rangle}_{y}\phi_{n+1}^{x}-{\langle}w_{n-1}{\rangle}_{y}\phi_{n}^{x}
=wny(ϕn+1xϕnx)+(wnywn1y)ϕnx,\displaystyle={\langle}w_{n}{\rangle}_{y}\left(\phi_{n+1}^{x}-\phi_{n}^{x}\right)+\left({\langle}w_{n}{\rangle}_{y}-{\langle}w_{n-1}{\rangle}_{y}\right)\phi_{n}^{x},

and energy estimates yield, for T1]0,T]T_{1}\in]0,T], since Φn+1t=0=Φnt=0\Phi_{n+1\mid t=0}=\Phi_{n\mid t=0},

(5.2) ϕn+1xϕnxLT1Lx20T1(wn(s)ywn1(s)y)ϕnx(s)Lx2𝑑s.\|\phi_{n+1}^{x}-\phi_{n}^{x}\|_{L^{\infty}_{T_{1}}L^{2}_{x}}\leq\int_{0}^{T_{1}}\left\|\left({\langle}w_{n}(s){\rangle}_{y}-{\langle}w_{n-1}(s){\rangle}_{y}\right)\phi_{n}^{x}(s)\right\|_{L^{2}_{x}}ds.

In view of (H2), the key term is estimated by

|wn(t)ywn1(t)y|d2(V1(x)+V2(y)+1)||ϕny(t,y)|2|ϕn1y(t,y)|2|𝑑y.\left|{\langle}w_{n}(t){\rangle}_{y}-{\langle}w_{n-1}(t){\rangle}_{y}\right|\lesssim\int_{{\mathbb{R}}^{d_{2}}}(V_{1}(x)+V_{2}(y)+1)\left||\phi_{n}^{y}(t,y)|^{2}-|\phi^{y}_{n-1}(t,y)|^{2}\right|dy.

Writing |ϕny|2|ϕn1y|2=Re((ϕnyϕn1y)(ϕny¯+ϕn1y¯))|\phi_{n}^{y}|^{2}-|\phi^{y}_{n-1}|^{2}=\text{\rm Re}\left((\phi_{n}^{y}-\phi^{y}_{n-1})(\overline{\phi_{n}^{y}}+\overline{\phi^{y}_{n-1}})\right), and using Cauchy-Schwarz inequality,

|wn(t)ywn1(t)y|\displaystyle\left|{\langle}w_{n}(t){\rangle}_{y}-{\langle}w_{n-1}(t){\rangle}_{y}\right| (V1(x)+1)(ϕnyLy2+ϕn1yLy2)ϕnyϕn1yLy2\displaystyle\lesssim(V_{1}(x)+1)\left(\|\phi_{n}^{y}\|_{L^{2}_{y}}+\|\phi_{n-1}^{y}\|_{L^{2}_{y}}\right)\|\phi_{n}^{y}-\phi_{n-1}^{y}\|_{L^{2}_{y}}
+(V2ϕnyLy2+V2ϕn1yLy2)ϕnyϕn1yLy2\displaystyle\quad+\left(\|V_{2}\phi_{n}^{y}\|_{L^{2}_{y}}+\|V_{2}\phi_{n-1}^{y}\|_{L^{2}_{y}}\right)\|\phi_{n}^{y}-\phi_{n-1}^{y}\|_{L^{2}_{y}}
(5.3) (V1(x)+1))supkΦkXT2,2ϕnyϕn1yLy2.\displaystyle\lesssim\left(V_{1}(x)+1)\right)\sup_{k\in{\mathbb{N}}}\|\Phi_{k}\|_{X^{2,2}_{T}}\|\phi_{n}^{y}-\phi_{n-1}^{y}\|_{L^{2}_{y}}.

Plugging this estimate into (5.2), we infer, thanks to Minkowski inequality,

ϕn+1xϕnxLT1Lx2\displaystyle\|\phi_{n+1}^{x}-\phi_{n}^{x}\|_{L^{\infty}_{T_{1}}L^{2}_{x}} supkΦkXT2,20T1(V1+1))ϕnx(s)Lx2ϕny(s)ϕn1y(s)Ly2ds\displaystyle\lesssim\sup_{k\in{\mathbb{N}}}\|\Phi_{k}\|_{X^{2,2}_{T}}\int_{0}^{T_{1}}\left\|\left(V_{1}+1)\right)\phi_{n}^{x}(s)\right\|_{L^{2}_{x}}\|\phi_{n}^{y}(s)-\phi_{n-1}^{y}(s)\|_{L^{2}_{y}}ds
supkΦkXT2,220T1ϕny(s)ϕn1y(s)Ly2𝑑s\displaystyle\lesssim\sup_{k\in{\mathbb{N}}}\|\Phi_{k}\|_{X^{2,2}_{T}}^{2}\int_{0}^{T_{1}}\|\phi_{n}^{y}(s)-\phi_{n-1}^{y}(s)\|_{L^{2}_{y}}ds
R2T1supt[0,T1]ϕny(t)ϕn1y(t)Ly2.\displaystyle\lesssim R^{2}T_{1}\sup_{t\in[0,T_{1}]}\|\phi_{n}^{y}(t)-\phi_{n-1}^{y}(t)\|_{L^{2}_{y}}.

We obtain a similar estimate by exchanging the roles of xx and yy, and so

(5.4) Φn+1ΦnXT10,0R2T1ΦnΦn1XT10,0.\|\Phi_{n+1}-\Phi_{n}\|_{X^{0,0}_{T_{1}}}\lesssim R^{2}T_{1}\|\Phi_{n}-\Phi_{n-1}\|_{X^{0,0}_{T_{1}}}.

Fixing T1]0,T]T_{1}\in]0,T] sufficiently small, the series

nΦn+1ΦnXT10,0\sum_{n\in{\mathbb{N}}}\|\Phi_{n+1}-\Phi_{n}\|_{X^{0,0}_{T_{1}}}

converges geometrically, and Φn\Phi_{n} converges in XT10,0X_{T_{1}}^{0,0}, to some ΦXT10,0\Phi\in X_{T_{1}}^{0,0}.

On the other hand, the boundedness of (Φn)n(\Phi_{n})_{n} in XT2,2X_{T}^{2,2} implies that a subsequence is converging in the weak-* topology of XT2,2X_{T}^{2,2}. By uniqueness of limits in the sense of distributions, we infer ΦXT12,2\Phi\in X_{T_{1}}^{2,2}. The same holds when XT2,2X_{T}^{2,2} is replaced by XTα,βX_{T}^{\alpha,\beta} for α,β2\alpha,\beta\geq 2. ∎

6. Passing to the limit in the equation

We now have all the elements in hands for proving Theorem 3.11 by showing that the limit function Φ\Phi constructed in Lemma 5.1 is a solution to equation (1.2) with the properties stated in Theorem 3.11.

6.1. Existence of a local solution

Combining Lemmas 4.1 and 5.1, we infer that under the assumptions of Theorem 3.11, there exists T1>0T_{1}>0 such that ΦnΦ\Phi_{n}\to\Phi in XT10,0X_{T_{1}}^{0,0}. By uniqueness of the limit, we also have ΦnΦ\Phi_{n}\rightharpoonup\Phi in XT1α,βX_{T_{1}}^{\alpha,\beta} (and no extraction of a subsequence is needed). Resuming the estimates from the proof of Lemma 5.1, we observe that for n,mn,m\in{\mathbb{N}}, t[0,T1]t\in[0,T_{1}] and xd1x\in{\mathbb{R}}^{d_{1}},

|wn(t)ywm(t)y|\displaystyle|{\langle}w_{n}(t){\rangle}_{y}-{\langle}w_{m}(t){\rangle}_{y}| =|d2w(x,y)(|ϕny(t,y)|2|ϕmy(t,y)|2)𝑑y|\displaystyle=\left|\int_{{\mathbb{R}}^{d_{2}}}w(x,y)\left(|\phi_{n}^{y}(t,y)|^{2}-|\phi_{m}^{y}(t,y)|^{2}\right)dy\right|
(V1(x)+1)supkΦkXT2,2ϕny(t)ϕmy(t)Ly2.\displaystyle\lesssim\left(V_{1}(x)+1\right)\sup_{k\in{\mathbb{N}}}\|\Phi_{k}\|_{X^{2,2}_{T}}\|\phi_{n}^{y}(t)-\phi_{m}^{y}(t)\|_{L^{2}_{y}}.

Passing to the limit m+m\rightarrow+\infty, we obtain that for nn\in{\mathbb{N}}, t[0,T1]t\in[0,T_{1}] and xd1x\in{\mathbb{R}}^{d_{1}},

|wn(t)yw(t)y|\displaystyle|{\langle}w_{n}(t){\rangle}_{y}-{\langle}w(t){\rangle}_{y}| (V1(x)+1)supkΦkXT2,2ϕny(t)ϕy(t)Ly2.\displaystyle\lesssim\left(V_{1}(x)+1\right)\sup_{k\in{\mathbb{N}}}\|\Phi_{k}\|_{X^{2,2}_{T}}\|\phi_{n}^{y}(t)-\phi^{y}(t)\|_{L^{2}_{y}}.

Therefore, keeping the same notation RR as from Lemma 5.1,

wn(t)yϕn+1x(t)w(t)yϕx(t)Lx2\displaystyle\left\|{\langle}w_{n}(t){\rangle}_{y}\phi_{n+1}^{x}(t)-{\langle}w(t){\rangle}_{y}\phi^{x}(t)\right\|_{L^{2}_{x}} (wn(t)yw(t)y)ϕn+1x(t)Lx2\displaystyle\lesssim\left\|\left({\langle}w_{n}(t){\rangle}_{y}-{\langle}w(t){\rangle}_{y}\right)\phi_{n+1}^{x}(t)\right\|_{L^{2}_{x}}
+w(t)y(ϕn+1x(t)ϕx(t))Lx2\displaystyle\quad+\left\|{\langle}w(t){\rangle}_{y}\left(\phi_{n+1}^{x}(t)-\phi^{x}(t)\right)\right\|_{L^{2}_{x}}
Rϕny(t)ϕy(t)Ly2(V1+1)ϕn+1x(t)Lx2\displaystyle\lesssim R\|\phi_{n}^{y}(t)-\phi^{y}(t)\|_{L^{2}_{y}}\left\|(V_{1}+1)\phi_{n+1}^{x}(t)\right\|_{L^{2}_{x}}
+(V1+1+ϕyLT1y12)(ϕn+1x(t)ϕx(t))Lx2,\displaystyle\quad+\left\|\left(V_{1}+1+\|\phi^{y}\|^{2}_{L^{\infty}_{T_{1}}\mathcal{H}^{1}_{y}}\right)\left(\phi_{n+1}^{x}(t)-\phi^{x}(t)\right)\right\|_{L^{2}_{x}},

where we have used (4.3) and the normalization (1.3). The first term on the right hand side goes to zero as nn\to\infty, uniformly in t[0,T1]t\in[0,T_{1}]. So does the last one in the case α,β3\alpha,\beta\geq 3, since by interpolation Φn\Phi_{n} then converges to Φ\Phi strongly in XT2,2X^{2,2}_{T}. In the case where α\alpha or β\beta is equal to 22, we can only claim a weak convergence,

wnyϕn+1xnwyϕxin L([0,T1];Lx2) weak-.{\langle}w_{n}{\rangle}_{y}\phi_{n+1}^{x}\mathop{\rightharpoonup}\limits_{n\rightarrow\infty}{\langle}w{\rangle}_{y}\phi^{x}\quad\text{in }L^{\infty}([0,T_{1}];L^{2}_{x})\text{ weak-$*$}.

Similarly,

wnxϕn+1ynwxϕyin L([0,T1];Ly2) weak-,{\langle}w_{n}{\rangle}_{x}\phi_{n+1}^{y}\mathop{\rightharpoonup}\limits_{n\rightarrow\infty}{\langle}w{\rangle}_{x}\phi^{y}\quad\text{in }L^{\infty}([0,T_{1}];L^{2}_{y})\text{ weak-$*$},

and Φ\Phi solves (1.2) for t[0,T1]t\in[0,T_{1}], in the sense of distributions. In view of the regularity ΦXT1α,β\Phi\in X_{T_{1}}^{\alpha,\beta}, Duhamel’s formula,

ϕx(t)=eitHxϕ0xi0tei(ts)Hx(wyϕx)(s)𝑑s,\displaystyle\phi^{x}(t)=e^{-itH_{x}}\phi_{0}^{x}-i\int_{0}^{t}e^{-i(t-s)H_{x}}\left({\langle}w{\rangle}_{y}\phi_{x}\right)(s)ds,
ϕy(t)=eitHyϕ0yi0tei(ts)Hy(wxϕy)(s)𝑑s,\displaystyle\phi^{y}(t)=e^{-itH_{y}}\phi_{0}^{y}-i\int_{0}^{t}e^{-i(t-s)H_{y}}\left({\langle}w{\rangle}_{x}\phi_{y}\right)(s)ds,

then shows the continuity in time ΦC([0,T1];Lx2×Ly2)\Phi\in C([0,T_{1}];L^{2}_{x}\times L^{2}_{y}).

6.2. Uniqueness

At this stage, it is rather clear that uniqueness holds in XT2,2X^{2,2}_{T}, no matter how large α\alpha and β\beta are. Suppose that Φ~XT2,2\tilde{\Phi}\in X_{T}^{2,2} is another solution to (1.2) for T>0T>0: the system satisfied by ΦΦ~\Phi-\tilde{\Phi} is similar to the one satisfied by Φn+1Φn\Phi_{n+1}-\Phi_{n}, and considered in the proof of Lemma 5.1. Since Φ,Φ~XT2,2\Phi,\tilde{\Phi}\in X_{T}^{2,2}, there exists R>0R>0 such that

ΦXT2,2+Φ~XT2,2R,\|\Phi\|_{X_{T}^{2,2}}+\|\tilde{\Phi}\|_{X_{T}^{2,2}}\leq R,

and repeating the computations presented in the proof of Lemma 5.1, we obtain, for any T1]0,T]T_{1}\in]0,T],

ΦΦ~XT10,0CT1RΦΦ~XT10,0.\|\Phi-\tilde{\Phi}\|_{X_{T_{1}}^{0,0}}\leq CT_{1}R\|\Phi-\tilde{\Phi}\|_{X_{T_{1}}^{0,0}}.

Picking T1>0T_{1}>0 such that CT1R<1CT_{1}R<1 shows that ΦΦ~\Phi\equiv\tilde{\Phi} for t[0,T1]t\in[0,T_{1}], and we infer that ΦΦ~\Phi\equiv\tilde{\Phi} on [0,T][0,T] by covering [0,T][0,T] by finitely many intervals of length at most T1T_{1}.

6.3. Conservations

We now address the second point in Theorem 3.11: we assume that (1.2) has a unique solution ΦXT2,2\Phi\in X^{2,2}_{T} for some T>0T>0. This implies in particular, in view of (1.2), that tϕxL([0,T];Lx2){\partial}_{t}\phi^{x}\in L^{\infty}([0,T];L^{2}_{x}) and tϕyL([0,T];Ly2){\partial}_{t}\phi^{y}\in L^{\infty}([0,T];L^{2}_{y}), and the multiplier techniques evoked below are justified without using regularizing argument as in e.g. [10].

For the conservation of the L2L^{2}-norms, multiply the first equation in (1.2) by ϕx¯\overline{\phi^{x}}, integrate in space on d1{\mathbb{R}}^{d_{1}}, and consider the imaginary part: we readily obtain

ddtϕx(t)Lx22=0.\frac{d}{dt}\|\phi^{x}(t)\|_{L^{2}_{x}}^{2}=0.

We proceed similarly for ϕy\phi^{y}, and the conservation of the L2L^{2}-norms follows.

For the energy, consider the multiplier tϕx¯{\partial}_{t}\overline{\phi^{x}} in the equation for ϕx\phi^{x}: as evoked above, all the products are well-defined, in the worst possible case as products of two L2L^{2} functions. Integrate in space and consider the real part: we obtain

ddtE(t)=0.\frac{d}{dt}E(t)=0.

6.4. Globalization

In view of Lemmas 4.1 and 5.1, it suffices to prove a priori estimates on ΦXT2,2\|\Phi\|_{X_{T}^{2,2}}, showing that this quantity is locally bounded in TT, to infer that ΦXT2,2\Phi\in X_{T}^{2,2} for all T>0T>0, and then globalize the solution by the standard ODE alternative.

We use the conservation of the total energy, whose expression we develop:

E(t)\displaystyle E(t) =(Hxϕx(t),ϕx(t))Lx2+(Hyϕy(t),ϕy(t))Ly2\displaystyle=\left(H_{x}\phi^{x}(t),\phi^{x}(t)\right)_{L^{2}_{x}}+\left(H_{y}\phi^{y}(t),\phi^{y}(t)\right)_{L^{2}_{y}}
+d1×d2w(x,y)|ϕx(t,x)|2|ϕy(t,y)|2𝑑x𝑑y\displaystyle\quad+\iint_{{\mathbb{R}}^{d_{1}}\times{\mathbb{R}}^{d_{2}}}w(x,y)|\phi^{x}(t,x)|^{2}|\phi^{y}(t,y)|^{2}dxdy
=12xϕx(t)L2(d1)2+d1V1(x)|ϕx(t,x)|2𝑑x+12yϕy(t)L2(d2)2\displaystyle=\frac{1}{2}\|\nabla_{x}\phi^{x}(t)\|_{L^{2}({\mathbb{R}}^{d_{1}})}^{2}+\int_{{\mathbb{R}}^{d_{1}}}V_{1}(x)|\phi^{x}(t,x)|^{2}dx+\frac{1}{2}\|\nabla_{y}\phi^{y}(t)\|_{L^{2}({\mathbb{R}}^{d_{2}})}^{2}
+d2V2(y)|ϕy(t,y)|2𝑑y+d1×d2w(x,y)|ϕx(t,x)|2|ϕy(t,y)|2𝑑x𝑑y.\displaystyle\quad+\int_{{\mathbb{R}}^{d_{2}}}V_{2}(y)|\phi^{y}(t,y)|^{2}dy+\iint_{{\mathbb{R}}^{d_{1}}\times{\mathbb{R}}^{d_{2}}}w(x,y)|\phi^{x}(t,x)|^{2}|\phi^{y}(t,y)|^{2}dxdy.

Since c0<1c_{0}<1 in (H2), we infer

E(t)\displaystyle E(t) 12xϕx(t)L2(d1)2+12yϕy(t)L2(d2)2+(1c0)d1V1(x)|ϕx(t,x)|2𝑑x\displaystyle\geq\frac{1}{2}\|\nabla_{x}\phi^{x}(t)\|_{L^{2}({\mathbb{R}}^{d_{1}})}^{2}+\frac{1}{2}\|\nabla_{y}\phi^{y}(t)\|_{L^{2}({\mathbb{R}}^{d_{2}})}^{2}+(1-c_{0})\int_{{\mathbb{R}}^{d_{1}}}V_{1}(x)|\phi^{x}(t,x)|^{2}dx
+(1c0)d2V2(y)|ϕy(t,y)|2𝑑yc0Cd1|ϕx(t,x)|2𝑑xc0Cd2|ϕy(t,y)|2𝑑y.\displaystyle\quad+(1-c_{0})\int_{{\mathbb{R}}^{d_{2}}}V_{2}(y)|\phi^{y}(t,y)|^{2}dy-c_{0}C\int_{{\mathbb{R}}^{d_{1}}}|\phi^{x}(t,x)|^{2}dx-c_{0}C\int_{{\mathbb{R}}^{d_{2}}}|\phi^{y}(t,y)|^{2}dy.

The conservations established above yield

12xϕx(t)L2(d1)2\displaystyle\frac{1}{2}\|\nabla_{x}\phi^{x}(t)\|_{L^{2}({\mathbb{R}}^{d_{1}})}^{2} +12yϕy(t)L2(d2)2+(1c0)d1V1(x)|ϕx(t,x)|2𝑑x\displaystyle+\frac{1}{2}\|\nabla_{y}\phi^{y}(t)\|_{L^{2}({\mathbb{R}}^{d_{2}})}^{2}+(1-c_{0})\int_{{\mathbb{R}}^{d_{1}}}V_{1}(x)|\phi^{x}(t,x)|^{2}dx
+(1c0)d2V2(y)|ϕy(t,y)|2𝑑yE(0)+2c0C.\displaystyle+(1-c_{0})\int_{{\mathbb{R}}^{d_{2}}}V_{2}(y)|\phi^{y}(t,y)|^{2}dy\leq E(0)+2c_{0}C.

This is the coercivity property announced in the introduction, showing that there exists MM depending only on Φ01,1\|\Phi_{0}\|_{1,1} such that

ΦXT1,1M,\|\Phi\|_{X_{T}^{1,1}}\leq M,

for any interval [0,T][0,T] on which the solution is well-defined. Proceeding like in the proof of Lemma 4.1, we have

(6.1) supt[0,T]Hxϕx(t)Lx22Hxϕ0xLx22+20T|([Hx,wy(t)]ϕx(t),Hxϕx(t))Lx2|𝑑t.\sup_{t\in[0,T]}\|H_{x}\phi^{x}(t)\|_{L^{2}_{x}}^{2}\leq\|H_{x}\phi_{0}^{x}\|_{L^{2}_{x}}^{2}+2\int_{0}^{T}\left|\left([H_{x},\langle w\rangle_{y}(t)]\phi^{x}(t),H_{x}\phi^{x}(t)\right)_{L^{2}_{x}}\right|dt.

In view of Lemma 4.2 with k=1k=1, f=ϕxf=\phi^{x} and g=Hxϕxg=H_{x}\phi^{x}, we infer

supt[0,T]Hxϕx(t)Lx22\displaystyle\sup_{t\in[0,T]}\|H_{x}\phi^{x}(t)\|_{L^{2}_{x}}^{2} Hxϕ0xLx22+CϕyLTy120Tϕx(t)x2Hxϕx(t)Lx2𝑑t\displaystyle\leq\|H_{x}\phi_{0}^{x}\|_{L^{2}_{x}}^{2}+C\|\phi^{y}\|^{2}_{L^{\infty}_{T}\mathcal{H}^{1}_{y}}\int_{0}^{T}\|\phi^{x}(t)\|_{\mathcal{H}^{2}_{x}}\|H_{x}\phi^{x}(t)\|_{L^{2}_{x}}dt
Hxϕ0xLx22+CM20Tϕx(t)x22𝑑t.\displaystyle\leq\|H_{x}\phi_{0}^{x}\|_{L^{2}_{x}}^{2}+CM^{2}\int_{0}^{T}\|\phi^{x}(t)\|_{\mathcal{H}^{2}_{x}}^{2}dt.

The conservation of the L2L^{2}-norm of ϕx\phi^{x} implies

supt[0,T]ϕx(t)x22ϕxx22+CM20Tϕx(t)x22𝑑t,\sup_{t\in[0,T]}\|\phi^{x}(t)\|_{\mathcal{H}^{2}_{x}}^{2}\leq\|\phi^{x}\|_{\mathcal{H}^{2}_{x}}^{2}+CM^{2}\int_{0}^{T}\|\phi^{x}(t)\|_{\mathcal{H}^{2}_{x}}^{2}dt,

hence an exponential a priori control of the x2\mathcal{H}^{2}_{x}-norm of ϕx(t)\phi^{x}(t) by Gronwall lemma. The same holds for ϕy(t)\phi^{y}(t), hence the conclusion of Theorem 3.11.

7. Proof of Lemma 3.9

We briefly explain why (3.3) implies (H3)2,2, thanks to an integration by parts, in view of (3.2). Typically, for f1,f2𝒮(d1)f_{1},f_{2}\in{\mathcal{S}}({\mathbb{R}}^{d_{1}}),

[w(x,y),Hx]f1,f2Lx2=12Δxw(,y)f1,f2Lx2+xw(,y)f1,f2Lx2.{\langle}[w(x,y),H_{x}]f_{1},f_{2}{\rangle}_{L^{2}_{x}}=\frac{1}{2}{\langle}\Delta_{x}w(\cdot,y)f_{1},f_{2}{\rangle}_{L^{2}_{x}}+{\langle}\nabla_{x}w(\cdot,y)\cdot\nabla f_{1},f_{2}{\rangle}_{L^{2}_{x}}.

Therefore, for almost all yd2y\in{\mathbb{R}}^{d_{2}}, Cauchy-Schwarz inequality yields

|[w(x,y),Hx]f1,f2Lx2|\displaystyle\left|{\langle}[w(x,y),H_{x}]f_{1},f_{2}{\rangle}_{L^{2}_{x}}\right| 12|Δxw(,y)|1/2f1Lx2|Δxw(,y)|1/2f2Lx2\displaystyle\leq\frac{1}{2}\||\Delta_{x}w(\cdot,y)|^{1/2}f_{1}\|_{L^{2}_{x}}\||\Delta_{x}w(\cdot,y)|^{1/2}f_{2}\|_{L^{2}_{x}}
+xf1Lx2xw(,y)f2Lx2.\displaystyle\quad+\|\nabla_{x}f_{1}\|_{L^{2}_{x}}\|\nabla_{x}w(\cdot,y)f_{2}\|_{L^{2}_{x}}.

Using (3.3),

|Δxw(,y)|1/2fLx22\displaystyle\||\Delta_{x}w(\cdot,y)|^{1/2}f\|^{2}_{L^{2}_{x}} V1fLx22+(1+V2(y))fLx22fx12+(1+V2(y))fLx22,\displaystyle\lesssim\|\sqrt{V_{1}}f\|^{2}_{L^{2}_{x}}+(1+V_{2}(y))\|f\|^{2}_{L^{2}_{x}}\lesssim\|f\|^{2}_{\mathcal{H}^{1}_{x}}+(1+V_{2}(y))\|f\|^{2}_{L^{2}_{x}},
xw(,y)fLx2\displaystyle\|\nabla_{x}w(\cdot,y)f\|_{L^{2}_{x}} (V1+V2(y)+1)fLx2fx1+(1+V2(y))fLx2.\displaystyle\lesssim\|(\sqrt{V_{1}}+V_{2}(y)+1)f\|_{L^{2}_{x}}\lesssim\|f\|_{\mathcal{H}^{1}_{x}}+(1+V_{2}(y))\|f\|_{L^{2}_{x}}.

We deduce the expected relation for k==1k=\ell=1:

|[w(x,y),Hx]f1,f2Lx2|(1+V2(y))f1x1f2x1.\left|{\langle}[w(x,y),H_{x}]f_{1},f_{2}{\rangle}_{L^{2}_{x}}\right|\lesssim(1+V_{2}(y))\|f_{1}\|_{\mathcal{H}^{1}_{x}}\|f_{2}\|_{\mathcal{H}^{1}_{x}}.

For k=2k=2, write

|Hx[w(x,y),Hx]f1,f2Lx2|\displaystyle\left|{\langle}H_{x}[w(x,y),H_{x}]f_{1},f_{2}{\rangle}_{L^{2}_{x}}\right| =|[w(x,y),Hx]f1,Hxf2Lx2|\displaystyle=\left|{\langle}[w(x,y),H_{x}]f_{1},H_{x}f_{2}{\rangle}_{L^{2}_{x}}\right|
12|Δxw(,y)f1,Hxf2Lx2|+|xw(,y)f1,Hxf2Lx2|\displaystyle\leq\frac{1}{2}\left|{\langle}\Delta_{x}w(\cdot,y)f_{1},H_{x}f_{2}{\rangle}_{L^{2}_{x}}\right|+\left|{\langle}\nabla_{x}w(\cdot,y)\cdot\nabla f_{1},H_{x}f_{2}{\rangle}_{L^{2}_{x}}\right|
(1+V1+V2(y))f1Lx2Hxf2Lx2\displaystyle\lesssim\|\left(1+V_{1}+V_{2}(y)\right)f_{1}\|_{L^{2}_{x}}\|H_{x}f_{2}\|_{L^{2}_{x}}
+xw(,y)f1Lx2Hxf2Lx2\displaystyle\quad+\|\nabla_{x}w(\cdot,y)\cdot\nabla f_{1}\|_{L^{2}_{x}}\|H_{x}f_{2}\|_{L^{2}_{x}}
f1x2f2x2+V2(y)f1Lx2f2x2\displaystyle\lesssim\|f_{1}\|_{\mathcal{H}^{2}_{x}}\|f_{2}\|_{\mathcal{H}^{2}_{x}}+V_{2}(y)\|f_{1}\|_{L^{2}_{x}}\|f_{2}\|_{\mathcal{H}^{2}_{x}}
+xw(,y)f1Lx2f2x2,\displaystyle\quad+\|\nabla_{x}w(\cdot,y)\cdot\nabla f_{1}\|_{L^{2}_{x}}\|f_{2}\|_{\mathcal{H}^{2}_{x}},

where we have used the estimate HxfLx2fx2\|H_{x}f\|_{L^{2}_{x}}\leq\|f\|_{\mathcal{H}_{x}^{2}}. For the last term, (3.3) yields

xw(,y)f1Lx2\displaystyle\|\nabla_{x}w(\cdot,y)\cdot\nabla f_{1}\|_{L^{2}_{x}} (V1+V2(y)+1)f1Lx2\displaystyle\lesssim\left\|\left(\sqrt{V_{1}}+V_{2}(y)+1\right)\nabla f_{1}\right\|_{L^{2}_{x}}
V1f1Lx2+(V2(y)+1)f1Lx2\displaystyle\lesssim\left\|\sqrt{V_{1}}\nabla f_{1}\right\|_{L^{2}_{x}}+\left(V_{2}(y)+1\right)\|\nabla f_{1}\|_{L^{2}_{x}}
V1f1Lx2+(V2(y)+1)fLx21/2ΔfLx21/2\displaystyle\lesssim\left\|\sqrt{V_{1}}\nabla f_{1}\right\|_{L^{2}_{x}}+\left(V_{2}(y)+1\right)\|f\|_{L^{2}_{x}}^{1/2}\|\Delta f\|_{L^{2}_{x}}^{1/2}
V1f1Lx2+(V2(y)+1)f1x2.\displaystyle\lesssim\left\|\sqrt{V_{1}}\nabla f_{1}\right\|_{L^{2}_{x}}+\left(V_{2}(y)+1\right)\|f_{1}\|_{\mathcal{H}^{2}_{x}}.

For the first term on the last right hand side, we use an integration by parts:

V1f1Lx22\displaystyle\left\|\sqrt{V_{1}}\nabla f_{1}\right\|_{L^{2}_{x}}^{2} =d1V1(x)f1(x)f1(x)𝑑x\displaystyle=\int_{{\mathbb{R}}^{d_{1}}}V_{1}(x)\nabla f_{1}(x)\cdot\nabla f_{1}(x)dx
=d1V1(x)f1(x)Δf1(x)𝑑xd1f1(x)V1(x)f1(x)𝑑x.\displaystyle=-\int_{{\mathbb{R}}^{d_{1}}}V_{1}(x)f_{1}(x)\Delta f_{1}(x)dx-\int_{{\mathbb{R}}^{d_{1}}}f_{1}(x)\nabla V_{1}(x)\cdot\nabla f_{1}(x)dx.

By Cauchy-Schwarz inequality, the first term on the right hand side is estimated by

V1f1Lx2Δf1Lx22Hxf1Lx22.\|V_{1}f_{1}\|_{L^{2}_{x}}\|\Delta f_{1}\|_{L^{2}_{x}}\leq 2\|H_{x}f_{1}\|_{L^{2}_{x}}^{2}.

Invoking (3.2), and using Cauchy-Schwarz inequality again,

|d1f(x)V1(x)f(x)𝑑x|\displaystyle\left|\int_{{\mathbb{R}}^{d_{1}}}f(x)\nabla V_{1}(x)\cdot\nabla f(x)dx\right| d1(1+V1(x))|f(x)||f(x)|𝑑x(1+V1)fLx2fLx2\displaystyle\lesssim\int_{{\mathbb{R}}^{d_{1}}}(1+V_{1}(x))|f(x)||\nabla f(x)|dx\lesssim\|(1+V_{1})f\|_{L^{2}_{x}}\|\nabla f\|_{L^{2}_{x}}
(fLx2+HxfLx2)fLx21/2ΔfLx21/2\displaystyle\lesssim\left(\|f\|_{L^{2}_{x}}+\|H_{x}f\|_{L^{2}_{x}}\right)\|f\|_{L^{2}_{x}}^{1/2}\|\Delta f\|_{L^{2}_{x}}^{1/2}
HxfLx21/2fLx23/2+HxfLx23/2fLx21/2fLx22+HxfLx22,\displaystyle\lesssim\|H_{x}f\|_{L^{2}_{x}}^{1/2}\|f\|_{L^{2}_{x}}^{3/2}+\|H_{x}f\|_{L^{2}_{x}}^{3/2}\|f\|_{L^{2}_{x}}^{1/2}\lesssim\|f\|_{L^{2}_{x}}^{2}+\|H_{x}f\|_{L^{2}_{x}}^{2},

where we have used Young inequality for the last estimate.

|[w(x,y),Hx]Hxf1,f2|\displaystyle\left|{\langle}[w(x,y),H_{x}]H_{x}f_{1},f_{2}{\rangle}\right| =|Hxf1,[w(x,y),Hx]f2|Hxf1Lx2Δxw(,y)f2Lx2.\displaystyle=\left|{\langle}H_{x}f_{1},[w(x,y),H_{x}]f_{2}{\rangle}\right|\leq\|H_{x}f_{1}\|_{L^{2}_{x}}\|\Delta_{x}w(\cdot,y)f_{2}\|_{L^{2}_{x}}.

To estimate [w(,y),Hx]Hxf1,f2Lx2{\langle}[w(\cdot,y),H_{x}]H_{x}f_{1},f_{2}{\rangle}_{L^{2}_{x}}, we use the self-adjointness of HxH_{x} and write

[w(,y),Hx]Hxf1,f2Lx2=Hxf1,[w(,y),Hx]f2Lx2.{\langle}[w(\cdot,y),H_{x}]H_{x}f_{1},f_{2}{\rangle}_{L^{2}_{x}}={\langle}H_{x}f_{1},[w(\cdot,y),H_{x}]f_{2}{\rangle}_{L^{2}_{x}}.

We use the above estimate, where the roles of f1f_{1} and f2f_{2} have been swapped, to conclude that the first inequality in (H3)2,2 holds. The proof of the second one is similar.

Appendix A Tangent space

For completeness, we give the elementary considerations for determining the tangent spaces of the Hartree manifold, Lemma 2.1.

Proof.

We consider a curve Γ(s)=φx(s)φy(s)\Gamma(s)=\varphi^{x}(s)\otimes\varphi^{y}(s)\in\mathcal{M} with Γ(0)=u\Gamma(0)=u. Then,

Γ˙(0)=φ˙x(0)φy+φxφ˙y(0),\dot{\Gamma}(0)=\dot{\varphi}^{x}(0)\otimes\varphi^{y}+\varphi^{x}\otimes\dot{\varphi}^{y}(0),

which verifies the claimed representation of any tangent function as

v=vxφy+φxvy.v=v^{x}\otimes\varphi^{y}+\varphi^{x}\otimes v^{y}.

Let us consider a=(ax,ay)2a=(a^{x},a^{y})\in{\mathbb{C}}^{2} with ax+ay=0a^{x}+a^{y}=0. We set wx=vx+axφxw^{x}=v^{x}+a^{x}\varphi^{x} and wy=vy+ayφyw^{y}=v^{y}+a^{y}\varphi^{y}. Then, w=wxφy+φxwyw=w^{x}\otimes\varphi^{y}+\varphi^{x}\otimes w^{y} satisfies

w=vxφy+φxvy+(ax+ay)φxφy=v.w=v^{x}\otimes\varphi^{y}+\varphi^{x}\otimes v^{y}+(a^{x}+a^{y})\varphi^{x}\otimes\varphi^{y}=v.

Choosing ax=φx,vx/φx,φxa^{x}=-\langle\varphi^{x},v^{x}\rangle/\langle\varphi^{x},\varphi^{x}\rangle and ay=axa^{y}=-a^{x}, we obtain a representation of vv satisfying the claimed gauge condition. We verify that this condition implies uniqueness. We assume that v=vxφy+φxvy=v~xφy+φxv~yv=v^{x}\otimes\varphi^{y}+\varphi^{x}\otimes v^{y}=\tilde{v}^{x}\otimes\varphi^{y}+\varphi^{x}\otimes\tilde{v}^{y} with φx,vx=φx,v~x=0\langle\varphi^{x},v^{x}\rangle=\langle\varphi^{x},\tilde{v}^{x}\rangle=0. Then, for any ϑyLy2\vartheta^{y}\in L^{2}_{y},

φxϑy,v=φx,φxϑy,vy=φx,φxϑy,v~y,\langle\varphi^{x}\otimes\vartheta^{y},v\rangle=\langle\varphi^{x},\varphi^{x}\rangle\langle\vartheta^{y},v^{y}\rangle=\langle\varphi^{x},\varphi^{x}\rangle\langle\vartheta^{y},\tilde{v}^{y}\rangle,

which implies vy=v~yv^{y}=\tilde{v}^{y}. Then, for any ϑxLy2\vartheta^{x}\in L^{2}_{y},

ϑxφy,vLx,y2\displaystyle\langle\vartheta^{x}\otimes\varphi^{y},v\rangle_{L^{2}_{x,y}} =ϑx,vxLx2φy,φyLy2+ϑx,φxLx2φy,vyLy2\displaystyle=\langle\vartheta^{x},v^{x}\rangle_{L^{2}_{x}}\langle\varphi^{y},\varphi^{y}\rangle_{L^{2}_{y}}+\langle\vartheta^{x},\varphi^{x}\rangle_{L^{2}_{x}}\langle\varphi^{y},v^{y}\rangle_{L^{2}_{y}}
=ϑx,v~xLx2φy,φyLy2+ϑx,φxLx2φy,v~yLy2,\displaystyle=\langle\vartheta^{x},\tilde{v}^{x}\rangle_{L^{2}_{x}}\langle\varphi^{y},\varphi^{y}\rangle_{L^{2}_{y}}+\langle\vartheta^{x},\varphi^{x}\rangle_{L^{2}_{x}}\langle\varphi^{y},\tilde{v}^{y}\rangle_{L^{2}_{y}},

which implies vx=v~xv^{x}=\tilde{v}^{x}. Choosing vx=0v^{x}=0 and vy=φyv^{y}=\varphi^{y}, we have v=uv=u so that u𝒯uu\in\mathcal{T}_{u}\mathcal{M}. ∎

Appendix B Coulombic type coupling

We recall standard definition and results.

Definition B.1 (Admissible pairs in 3{\mathbb{R}}^{3}).

A pair (q,r)(q,r) is admissible if q,r2q,r\geq 2, and

2q=3(121r).\frac{2}{q}=3\left(\frac{1}{2}-\frac{1}{r}\right).

As the range allowed for (q,r)(q,r) is compact, we set, for II\subset{\mathbb{R}} a time interval,

uS(I)=sup(q,r) admissibleuLq(I;Lr(3)).\|u\|_{S(I)}=\sup_{(q,r)\text{ admissible}}\|u\|_{L^{q}(I;L^{r}({\mathbb{R}}^{3}))}.

In view of [11] and [13], we have:

Proposition B.2.

Let d=3d=3 and 𝐕𝒬\mathbf{V}\in\mathcal{Q}. Denote 𝐇=12Δ+𝐕\mathbf{H}=-\frac{1}{2}\Delta+\mathbf{V}.
(1)(1) There exists ChomC_{\rm hom} such that for all interval II such that |I|1|I|\leq 1,

eit𝐇φS(I)ChomφL2,φL2(3).\|e^{-it\mathbf{H}}\varphi\|_{S(I)}\leq C_{\rm hom}\|\varphi\|_{L^{2}},\quad\forall\varphi\in L^{2}({\mathbb{R}}^{3}).

(2)(2) Denote

D(F)(t,x)=0tei(tτ)𝐇F(τ,x)dτ.D(F)(t,x)=\int_{0}^{t}e^{-i(t-\tau)\mathbf{H}}F(\tau,x)\mathrm{d}\tau.

There exists CinhomC_{\rm inhom} such that for all interval I0I\ni 0 such that |I|1|I|\leq 1,

D(F)S(I)CinhomFS(I).\left\lVert D(F)\right\rVert_{S(I)}\leq C_{\rm inhom}\left\lVert F\right\rVert_{S(I)^{*}}.

The existence of (local in time) Strichartz estimates of Proposition B.2 is the main ingredient of the proof of Theorem 3.1. Actually, as soon as such estimates are available for the operators 𝐇x\mathbf{H}_{x} and 𝐇y\mathbf{H}_{y}, then Theorem 3.1 remains valid. As mentioned in the introduction, such cases can be found in e.g. [20] or [5]. On the other hand, we emphasize that for superquadratic potentials, like V1V_{1} in Example 3.5, Strichartz estimates suffer a loss of regularity; see [16, 21].

Remark B.3.

The case of a harmonic potential, 𝐕(x)=|x|2\mathbf{V}(x)=|x|^{2}, shows that 𝐇\mathbf{H} may have eigenvalues, and explains why the above time intervals II are required to have finite length.

Remark B.4.

The potential 𝐕\mathbf{V} may also be time dependent, in view the original framework of [11]: 𝐕Lloc(t×x3)\mathbf{V}\in L^{\infty}_{\rm loc}({\mathbb{R}}_{t}\times{\mathbb{R}}_{x}^{3}) is real-valued, and smooth with respect to the space variable: for (almost) all tt\in{\mathbb{R}}, x𝐕(t,x)x\mapsto\mathbf{V}(t,x) is a CC^{\infty} map. Moreover, it is at most quadratic in space:

T>0,αd,|α|2,xα𝐕L([T,T]×x3).\forall T>0,\quad\forall\alpha\in{\mathbb{N}}^{d},\ |\alpha|\geq 2,\quad{\partial}_{x}^{\alpha}\mathbf{V}\in L^{\infty}([-T,T]\times{\mathbb{R}}_{x}^{3}).

Under these assumptions, suitable modifications of Proposition B.2 are needed, but they do not alter the conclusion of Theorem 3.1 (see [8]). See also [20] for another class of time dependent potentials.

Proof of Theorem 3.1.

We give the main technical steps of the proof, and refer to [10] for details. By Duhamel’s formula, we write (1.2) as

ϕx(t)\displaystyle\phi^{x}(t) =eit𝐇xϕ0xi0tei(tτ)𝐇x(v1ϕx+(W|ϕy|2)ϕx)(τ)dτ=:F1(ϕx,ϕy),\displaystyle=e^{-it\mathbf{H}_{x}}\phi_{0}^{x}-i\int_{0}^{t}e^{-i(t-\tau)\mathbf{H}_{x}}\left(v_{1}\phi^{x}+\left(W\ast|\phi^{y}|^{2}\right)\phi^{x}\right)(\tau)d\tau=:F_{1}(\phi^{x},\phi^{y}),
ϕy(t)\displaystyle\phi^{y}(t) =eit𝐇yϕ0yi0tei(tτ)𝐇y(v2ϕy+(W|ϕx|2)ϕy)(τ)dτ=:F2(ϕx,ϕy).\displaystyle=e^{-it\mathbf{H}_{y}}\phi_{0}^{y}-i\int_{0}^{t}e^{-i(t-\tau)\mathbf{H}_{y}}\left(v_{2}\phi^{y}+\left(W\ast|\phi^{x}|^{2}\right)\phi^{y}\right)(\tau)d\tau=:F_{2}(\phi^{x},\phi^{y}).

Theorem 3.1 follows from a standard fixed point argument based on Strichartz estimates. For 0<T10<T\leq 1, we introduce

Y(T)\displaystyle Y(T) ={(ϕx,ϕy)C([0,T];L2(3))2:ϕxS([0,T])2Chomϕ0xL2,\displaystyle=\{(\phi^{x},\phi^{y})\in C([0,T];L^{2}({\mathbb{R}}^{3}))^{2}:\quad\|\phi^{x}\|_{S([0,T])}\leq 2C_{\rm hom}\|\phi_{0}^{x}\|_{L^{2}},
ϕyS([0,T])2Chomϕ0yL2},\displaystyle\qquad\|\phi^{y}\|_{S([0,T])}\leq 2C_{\rm hom}\|\phi_{0}^{y}\|_{L^{2}}\},

and the distance

d(ϕ1,ϕ2)=ϕ1ϕ2S([0,T],d(\phi_{1},\phi_{2})=\|\phi_{1}-\phi_{2}\|_{S([0,T]},

where ChomC_{\rm hom} stems from Proposition B.2. Then (Y(T),d)(Y(T),d) is a complete metric space.

By using Strichartz estimates and Hölder inequality, we have:

F1(ϕx,ϕy)S([0,T])\displaystyle\|F_{1}(\phi^{x},\phi^{y})\|_{S([0,T])} Chomϕ0xL2+Cinhom(v1ϕxS([0,T])+(W|ϕy|2)ϕxS([0,T])),\displaystyle\leq C_{\rm hom}\|\phi_{0}^{x}\|_{L^{2}}+C_{\rm inhom}\left(\|v_{1}\phi^{x}\|_{S([0,T])^{*}}+\left\|\left(W\ast|\phi^{y}|^{2}\right)\phi^{x}\right\|_{S([0,T])^{*}}\right),

for any (ϕx,ϕy)Y(T)(\phi^{x},\phi^{y})\in Y(T). By assumption (see Theorem 3.1), we may write

v1=v1p+v1,v2=v2p+v2,W=Wp+W,v1q,v2q,WqLq(3),v_{1}=v_{1}^{p}+v_{1}^{\infty},\quad v_{2}=v_{2}^{p}+v_{2}^{\infty},\quad W=W^{p}+W^{\infty},\quad v_{1}^{q},v_{2}^{q},W^{q}\in L^{q}({\mathbb{R}}^{3}),

and the value pp can obviously be the same for the three potentials, by taking the minimum between the three pp’s if needed. Regarding v1ϕxS([0,T])\|v_{1}\phi^{x}\|_{S([0,T])^{*}}, we write

v1ϕxS([0,T])\displaystyle\|v_{1}^{\infty}\phi^{x}\|_{S([0,T])^{*}} v1ϕxL1([0,T];L2)v1LϕxL1([0,T];L2)\displaystyle\leq\|v_{1}^{\infty}\phi^{x}\|_{L^{1}([0,T];L^{2})}\leq\|v_{1}^{\infty}\|_{L^{\infty}}\|\phi^{x}\|_{L^{1}([0,T];L^{2})}
Tv1LϕxL([0,T];L2)Tv1LϕxS([0,T]).\displaystyle\leq T\|v_{1}^{\infty}\|_{L^{\infty}}\|\phi^{x}\|_{L^{\infty}\leq([0,T];L^{2})}\leq T\|v_{1}^{\infty}\|_{L^{\infty}}\|\phi^{x}\|_{S([0,T])}.

Let rr be such that

1r=1r+1p1=2r+1p.\frac{1}{r^{\prime}}=\frac{1}{r}+\frac{1}{p}\Longleftrightarrow 1=\frac{2}{r}+\frac{1}{p}.

Note that this exponent is the one introduced in the statement of Theorem 3.1. The assumption p>3/2p>3/2 implies 2r<62\leq r<6. Let qq be such that (q,r)(q,r) is admissible: r<6r<6 implies q>2q>2. Hölder inequality yields

v1pϕxS([0,T])\displaystyle\|v_{1}^{p}\phi^{x}\|_{S([0,T])^{*}} v1pϕxLq([0,T];Lr)v1pLpϕxLq([0,T];Lr)\displaystyle\leq\|v_{1}^{p}\phi^{x}\|_{L^{q^{\prime}}([0,T];L^{r^{\prime}})}\leq\|v_{1}^{p}\|_{L^{p}}\|\phi^{x}\|_{L^{q^{\prime}}([0,T];L^{r})}
T1/θv1pLpϕxLq([0,T];Lr)T1/θv1pLpϕxS([0,T]),\displaystyle\leq T^{1/\theta}\|v_{1}^{p}\|_{L^{p}}\|\phi^{x}\|_{L^{q}([0,T];L^{r})}\leq T^{1/\theta}\|v_{1}^{p}\|_{L^{p}}\|\phi^{x}\|_{S([0,T])},

where θ\theta is such that

1q=1q+1θ.\frac{1}{q^{\prime}}=\frac{1}{q}+\frac{1}{\theta}.

Note that θ\theta is finite, as q>2q>2

For the convolution term, first write

(W|ϕy|2)ϕxS([0,T])\displaystyle\left\|\left(W^{\infty}\ast|\phi^{y}|^{2}\right)\phi^{x}\right\|_{S([0,T])^{*}} (W|ϕy|2)ϕxL1([0,T];L2)\displaystyle\leq\left\|\left(W^{\infty}\ast|\phi^{y}|^{2}\right)\phi^{x}\right\|_{L^{1}([0,T];L^{2})}
W|ϕy|2L1([0,T];L)ϕxL([0,T];L2)\displaystyle\leq\left\|W^{\infty}\ast|\phi^{y}|^{2}\right\|_{L^{1}([0,T];L^{\infty})}\|\phi^{x}\|_{L^{\infty}([0,T];L^{2})}
TWLϕyL([0,T];L2)2ϕxL([0,T];L2)\displaystyle\leq T\|W^{\infty}\|_{L^{\infty}}\|\phi^{y}\|_{L^{\infty}([0,T];L^{2})}^{2}\|\phi^{x}\|_{L^{\infty}([0,T];L^{2})}
TWLϕyS([0,T])2ϕxS([0,T]).\displaystyle\leq T\|W^{\infty}\|_{L^{\infty}}\|\phi^{y}\|_{S([0,T])}^{2}\|\phi^{x}\|_{S([0,T])}.

Introduce r1r_{1} such that

(B.1) 1r1=1r1+12p2=4r1+1p1+12p=1p+2r1.\frac{1}{r_{1}^{\prime}}=\frac{1}{r_{1}}+\frac{1}{2p}\Longleftrightarrow 2=\frac{4}{r_{1}}+\frac{1}{p}\Longleftrightarrow 1+\frac{1}{2p}=\frac{1}{p}+\frac{2}{r_{1}}.

The assumption p>3/2p>3/2 implies 2r1<32\leq r_{1}<3. Let q1q_{1} be such that (q1,r1)(q_{1},r_{1}) is admissible: q1>4q_{1}>4. Hölder inequality yields

(Wp|ϕy|2)ϕxS([0,T])\displaystyle\left\|\left(W^{p}\ast|\phi^{y}|^{2}\right)\phi^{x}\right\|_{S([0,T])^{*}} (Wp|ϕy|2)ϕxLq1([0,T];Lr1)\displaystyle\leq\left\|\left(W^{p}\ast|\phi^{y}|^{2}\right)\phi^{x}\right\|_{L^{q_{1}^{\prime}}([0,T];L^{r_{1}^{\prime}})}
Wp|ϕy|2Lk([0,T];L2p)ϕxLq1([0,T];Lr1)\displaystyle\leq\left\|W^{p}\ast|\phi^{y}|^{2}\right\|_{L^{k}([0,T];L^{2p})}\|\phi^{x}\|_{L^{q_{1}}([0,T];L^{r_{1}})}
Wp|ϕy|2Lk([0,T];L2p)ϕxS([0,T]),\displaystyle\leq\left\|W^{p}\ast|\phi^{y}|^{2}\right\|_{L^{k}([0,T];L^{2p})}\|\phi^{x}\|_{S([0,T])},

where kk is such that

1q1=1q1+1k1=2q1+1k.\frac{1}{q_{1}^{\prime}}=\frac{1}{q_{1}}+\frac{1}{k}\Longleftrightarrow 1=\frac{2}{q_{1}}+\frac{1}{k}.

Note that since q1>4q_{1}>4, we have q1>2kq_{1}>2k. In view of (B.1), Young inequality yields

Wp|ϕy|2Lk([0,T];L2p)\displaystyle\left\|W^{p}\ast|\phi^{y}|^{2}\right\|_{L^{k}([0,T];L^{2p})} WpLp|ϕy|2Lk([0,T];Lr1/2)=WpLpϕyL2k([0,T];Lr1)2\displaystyle\leq\|W^{p}\|_{L^{p}}\left\||\phi^{y}|^{2}\right\|_{L^{k}([0,T];L^{r_{1}/2})}=\|W^{p}\|_{L^{p}}\|\phi^{y}\|_{L^{2k}([0,T];L^{r_{1}})}^{2}
TηWpLpϕyLq1([0,T];Lr1)2TηWpLpϕyS([0,T])2,\displaystyle\leq T^{\eta}\|W^{p}\|_{L^{p}}\|\phi^{y}\|_{L^{q_{1}}([0,T];L^{r_{1}})}^{2}\leq T^{\eta}\|W^{p}\|_{L^{p}}\|\phi^{y}\|_{S([0,T])}^{2},

where η>0\eta>0 is given by η=1/(2k)1/q1\eta=1/(2k)-1/q_{1}.

The same inequalities obviously holds by switching xx and yy, and so for T>0T>0 sufficiently small, Φ:=(ϕx,ϕy)(F1(ϕx,ϕy),F2(ϕx,ϕy))=:𝐅(Φ)\Phi:=(\phi^{x},\phi^{y})\mapsto(F_{1}(\phi^{x},\phi^{y}),F_{2}(\phi^{x},\phi^{y}))=:\mathbf{F}(\Phi) leaves Y(T)Y(T) invariant.

Using similar estimates, again relying on Strichartz and Hölder inequalities involving the same Lebesgue exponents (𝐅\mathbf{F} is the sum of a linear and a trilinear term in Φ\Phi), we infer that up to decreasing T>0T>0, 𝐅\mathbf{F} is a contraction on Y(T)Y(T), and so there exists a unique ΦY(T)\Phi\in Y(T) solving (1.2). The global existence of the solution for (1.2) follows from the conservation of the L2L^{2}-norms of ϕx\phi^{x} and ϕy\phi^{y}, respectively.

Uniqueness of such solutions follows once again from Strichartz and Hölder inequalities involving the same Lebesgue exponents as above, like for the contraction part of the argument. The main remark consists in noticing that the above Lebesgue indices satisfy r>r1r>r_{1}, hence q<q1q<q_{1}, and so Llocq1Lr1LlocqLrLL2L_{\rm loc}^{q_{1}}L^{r_{1}}\subset L_{\rm loc}^{q}L^{r}\cap L^{\infty}L^{2}. ∎

Remark B.5 (H1H^{1}-regularity).

If in Theorem 3.1, we assume in addition that

v1,v2Lp(3)+L(3)form some p>3/2,\nabla v_{1},\nabla v_{2}\in L^{p}({\mathbb{R}}^{3})+L^{\infty}({\mathbb{R}}^{3})\quad\text{form some }p>3/2,

then for ϕ0x,ϕ0yH1(3)\phi_{0}^{x},\phi_{0}^{y}\in H^{1}({\mathbb{R}}^{3}) and xϕ0x,yϕ0yL2(3)x\phi_{0}^{x},y\phi_{0}^{y}\in L^{2}({\mathbb{R}}^{3}) (this last assumption may be removed when V1,V2L(3)\nabla V_{1},\nabla V_{2}\in L^{\infty}({\mathbb{R}}^{3}) – the minimal assumption to work at the H1H^{1}-level with V1,V2𝒬V_{1},V_{2}\in\mathcal{Q} is ϕ0xV1,ϕ0yV2L2(3)\phi_{0}^{x}\nabla V_{1},\phi_{0}^{y}\nabla V_{2}\in L^{2}({\mathbb{R}}^{3}), see [9]), the global solution constructed in Theorem 3.1 satisfies

(ϕx,ϕy)C(;H1(3))2Llocq(;W1,r(3))2,(xϕx,yϕy)C(;L2(3)3)2.(\phi^{x},\phi^{y})\in C({\mathbb{R}};H^{1}({\mathbb{R}}^{3}))^{2}\cap L^{q}_{\rm loc}({\mathbb{R}};W^{1,r}({\mathbb{R}}^{3}))^{2},\quad(x\phi^{x},y\phi^{y})\in C({\mathbb{R}};L^{2}({\mathbb{R}}^{3})^{3})^{2}.

To see this, it suffices to resume the above proof, and check that xF1(ϕx,ϕy)\nabla_{x}F_{1}(\phi^{x},\phi^{y}) and yF2(ϕx,ϕy)\nabla_{y}F_{2}(\phi^{x},\phi^{y}) satisfy essentially the same estimates as F1(ϕx,ϕy),F2(ϕx,ϕy)F_{1}(\phi^{x},\phi^{y}),F_{2}(\phi^{x},\phi^{y}) in S([0,T])S([0,T]). One first has to commute the gradient with eit𝐇xe^{-it\mathbf{H}_{x}} or eit𝐇ye^{-it\mathbf{H}_{y}}. Typically,

xF1(ϕx,ϕy)\displaystyle\nabla_{x}F_{1}(\phi^{x},\phi^{y}) =eit𝐇xxϕ0xi0tei(tτ)𝐇xx(v1ϕx+(W|ϕy|2)ϕx)(τ)𝑑τ\displaystyle=e^{-it\mathbf{H}_{x}}\nabla_{x}\phi_{0}^{x}-i\int_{0}^{t}e^{-i(t-\tau)\mathbf{H}_{x}}\nabla_{x}\left(v_{1}\phi^{x}+\left(W\ast|\phi^{y}|^{2}\right)\phi^{x}\right)(\tau)d\tau
i0tei(tτ)𝐇xF1(ϕx,ϕy)(τ)xV1dτ,\displaystyle\quad-i\int_{0}^{t}e^{-i(t-\tau)\mathbf{H}_{x}}F_{1}(\phi^{x},\phi^{y})(\tau)\nabla_{x}V_{1}d\tau,

where the last factor accounts for the possible lack of commutation between 𝐇x\mathbf{H}_{x} and x\nabla_{x}, [it𝐇x,x]=xV1[-i{\partial}_{t}-\mathbf{H}_{x},\nabla_{x}]=\nabla_{x}V_{1}. Since V1V_{1} is at most quadratic, V1\nabla V_{1} is at most linear, and we obtain a closed system of estimates by considering

xF1(ϕx,ϕy)\displaystyle xF_{1}(\phi^{x},\phi^{y}) =eit𝐇x(xϕ0x)i0tei(tτ)𝐇x(x(v1ϕx+(W|ϕy|2)ϕx))(τ)𝑑τ\displaystyle=e^{-it\mathbf{H}_{x}}(x\phi_{0}^{x})-i\int_{0}^{t}e^{-i(t-\tau)\mathbf{H}_{x}}\left(x\left(v_{1}\phi^{x}+\left(W\ast|\phi^{y}|^{2}\right)\phi^{x}\right)\right)(\tau)d\tau
+i0tei(tτ)𝐇xxF1(ϕx,ϕy)(τ)𝑑τ,\displaystyle\quad+i\int_{0}^{t}e^{-i(t-\tau)\mathbf{H}_{x}}\nabla_{x}F_{1}(\phi^{x},\phi^{y})(\tau)d\tau,

where we have used [it𝐇x,x]=x[-i{\partial}_{t}-\mathbf{H}_{x},x]=-\nabla_{x}. We omit the details, and refer to [10] (see also [8]). As pointed in Remark 3.3, the energy

E(t)\displaystyle E(t) =(Hxϕx(t),ϕx(t))Lx2+(Hyϕy(t),ϕy(t))Ly2\displaystyle=\left(H_{x}\phi^{x}(t),\phi^{x}(t)\right)_{L^{2}_{x}}+\left(H_{y}\phi^{y}(t),\phi^{y}(t)\right)_{L^{2}_{y}}
+3×3W(xy)|ϕx(t,x)|2|ϕy(t,y)|2𝑑x𝑑y,\displaystyle\quad+\iint_{{\mathbb{R}}^{3}\times{\mathbb{R}}^{3}}W(x-y)|\phi^{x}(t,x)|^{2}|\phi^{y}(t,y)|^{2}dxdy,

which is well defined with the above regularity, is independent of time. Formally, this can be seen by multiplying the first equation in (1.2) by tϕx{\partial}_{t}\phi^{x}, the second by tϕy{\partial}_{t}\phi^{y}, integrating in space, considering the real part, and summing the two identities. To make the argument rigorous (we may not have enough regularity to be allowed to proceed as described), one may use a regularization procedure as in [10], or rely on a clever use of the regularity provided by Strichartz estimates, as in [17].

References

  • [1] S. Alinhac and P. Gérard. Pseudo-differential operators and the Nash-Moser theorem, volume 82 of Graduate Studies in Mathematics. American Mathematical Society, Providence, RI, 2007. Translated from the 1991 French original by Stephen S. Wilson.
  • [2] L. Baudouin. Existence and regularity of the solution of a time dependent Hartree-Fock equation coupled with a classical nuclear dynamics. Rev. Mat. Complut., 18(2):285–314, 2005.
  • [3] I. Burghardt, R. Carles, C. Fermanian Kammerer, B. Lasorne, and C. Lasser. Separation of scales: Dynamical approximations for composite quantum systems. J. Phys. A, 54(41):414002, 2021.
  • [4] I. Burghardt, R. Carles, C. Fermanian Kammerer, B. Lasorne, and C. Lasser. Dynamical approximations for composite quantum systems: assessment of error estimates for a separable ansatz. J. Phys. A, 55(22):224010, 2022.
  • [5] N. Burq, F. Planchon, J. G. Stalker, and A. S. Tahvildar-Zadeh. Strichartz estimates for the wave and Schrödinger equations with potentials of critical decay. Indiana Univ. Math. J., 53(6):1665–1680, 2004.
  • [6] F. Cacciafesta, A.-S. de Suzzoni, and D. Noja. A Dirac field interacting with point nuclear dynamics. Math. Ann., 376(3-4):1261–1301, 2020.
  • [7] E. Cancès and C. Le Bris. On the time-dependent Hartree-Fock equations coupled with a classical nuclear dynamics. Math. Models Methods Appl. Sci., 9(7):963–990, 1999.
  • [8] R. Carles. Nonlinear Schrödinger equation with time dependent potential. Commun. Math. Sci., 9(4):937–964, 2011.
  • [9] R. Carles. Sharp weights in the Cauchy problem for nonlinear Schrödinger equations with potential. Z. Angew. Math. Phys., 66(4):2087–2094, 2015.
  • [10] T. Cazenave. Semilinear Schrödinger equations, volume 10 of Courant Lecture Notes in Mathematics. New York University Courant Institute of Mathematical Sciences, New York, 2003.
  • [11] D. Fujiwara. Remarks on the convergence of the Feynman path integrals. Duke Math. J., 47(3):559–600, 1980.
  • [12] S. Jin, C. Sparber, and Z. Zhou. On the classical limit of a time-dependent self-consistent field system: analysis and computation. Kinet. Relat. Models, 10(1):263–298, 2017.
  • [13] M. Keel and T. Tao. Endpoint Strichartz estimates. Amer. J. Math., 120(5):955–980, 1998.
  • [14] C. Lasser and C. Su. Various variational approximations of quantum dynamics. Journal of Mathematical Physics, 63(7):072107, 2022.
  • [15] C. Lubich. From quantum to classical molecular dynamics: reduced models and numerical analysis. European Mathematical Society (EMS), Zürich, 2008.
  • [16] H. Mizutani. Strichartz estimates for Schrödinger equations with variable coefficients and unbounded potentials II. Superquadratic potentials. Commun. Pure Appl. Anal., 13(6):2177–2210, 2014.
  • [17] T. Ozawa. Remarks on proofs of conservation laws for nonlinear Schrödinger equations. Calc. Var. Partial Differ. Equ., 25(3):403–408, 2006.
  • [18] M. Reed and B. Simon. Methods of modern mathematical physics. II. Fourier analysis, self-adjointness. Academic Press [Harcourt Brace Jovanovich Publishers], New York, 1975.
  • [19] M. Reed and B. Simon. Methods of modern mathematical physics. I Functional analysis. Academic Press, Inc. [Harcourt Brace Jovanovich, Publishers], New York, second edition, 1980.
  • [20] I. Rodnianski and W. Schlag. Time decay for solutions of Schrödinger equations with rough and time-dependent potentials. Invent. Math., 155(3):451–513, 2004.
  • [21] K. Yajima and G. Zhang. Local smoothing property and Strichartz inequality for Schrödinger equations with potentials superquadratic at infinity. J. Differential Equations, 202(1):81–110, 2004.