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

Schrödinger equation in moving domains

Alessandro Duca Université Grenoble Alpes, CNRS, Institut Fourier, F-38000 Grenoble, France [email protected] Romain Joly Université Grenoble Alpes, CNRS, Institut Fourier, F-38000 Grenoble, France [email protected]
Abstract

We consider the Schrödinger equation

itu(t)=Δu(t) on Ω(t)i\partial_{t}u(t)=-\Delta u(t)\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \text{ on }\Omega(t) (\ast)

where Ω(t)N\Omega(t)\subset\mathbb{R}^{N} is a moving domain depending on the time t[0,T]t\in[0,T]. The aim of this work is to provide a meaning to the solutions of such an equation. We use the existence of a bounded reference domain Ω0\Omega_{0} and a specific family of unitary maps h(t):L2(Ω(t),)L2(Ω0,)h^{\sharp}(t):L^{2}(\Omega(t),\mathbb{C})\longrightarrow L^{2}(\Omega_{0},\mathbb{C}). We show that the conjugation by hh^{\sharp} provides a new equation of the form

itv=h(t)H(t)h(t)v on Ω0i\partial_{t}v=h^{\sharp}(t)H(t)h_{\sharp}(t)v\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \text{ on }\Omega_{0} (\ast\ast)

where h=(h)1h_{\sharp}=(h^{\sharp})^{-1}. The Hamiltonian H(t)H(t) is a magnetic Laplacian operator of the form

H(t)=(divx+iA)(x+iA)|A|2H(t)=-(\operatorname{div}_{x}+iA)\circ(\nabla_{x}+iA)-|A|^{2}

where AA is an explicit magnetic potential depending on the deformation of the domain Ω(t)\Omega(t). The formulation (\ast\astSchrödinger equation in moving domains) enables to ensure the existence of weak and strong solutions of the initial problem (\astSchrödinger equation in moving domains) on Ω(t)\Omega(t) endowed with Dirichlet boundary conditions. In addition, it also indicates that the correct Neumann type boundary conditions for (\astSchrödinger equation in moving domains) are not the homogeneous but the magnetic ones

νu(t)+iν|Au(t)=0,\partial_{\nu}u(t)+i\langle\nu|A\rangle u(t)=0,

even though (\astSchrödinger equation in moving domains) has no magnetic term. All the previous results are also studied in presence of diffusion coefficients as well as magnetic and electric potentials. Finally, we prove some associated byproducts as an adiabatic result for slow deformations of the domain and a time-dependent version of the so-called “Moser’s trick”. We use this outcome in order to simplify Equation (\ast\astSchrödinger equation in moving domains) and to guarantee the well-posedness for slightly less regular deformations of Ω(t)\Omega(t).

Keywords: Schrödinger equation, PDEs on moving domains, well-posedness, magnetic Laplacian operator, Moser’s trick, adiabatic result.

1 Main results

In this article, we study the well-posedness of the Schrödinger equation

itu(t,x)=Δu(t,x),tI,xΩ(t)i\partial_{t}u(t,x)=-\Delta u(t,x),\ \ \ \ \ \ \ t\in I\leavevmode\nobreak\ ,\leavevmode\nobreak\ \leavevmode\nobreak\ x\in\Omega(t) (1.1)

where II is an interval of times and tIΩ(t)Nt\in I\mapsto\Omega(t)\subset\mathbb{R}^{N} is a time-dependent family of bounded domains of N\mathbb{R}^{N} with d1d\geq 1. We consider the cases of Dirichlet boundary conditions and of suitable magnetic Neumann boundary conditions. This kind of problem is very natural when we consider a quantum particle confined in a structure which deforms in time.

The Schrödinger equation in moving domains has been widely studied in literature and an example is the classical article of Doescher and Rice [18]. For other references on the subject, we mention [3, 5, 8, 9, 17, 38, 40, 43, 47, 50]. In most of these references, (1.1) is studied in dimension d=1d=1 or in higher dimensions with symmetries as the radial case or the translating case. From this perspective, the purpose of this work is natural: we aim to study the well-posedness of (1.1) in a very general framework.

The difficulty of considering an equation in a moving domain as (1.1) is that the phase space L2(Ω(t),)L^{2}(\Omega(t),\mathbb{C}) and thus the operator Δ=Δ(t)\Delta=\Delta(t) depend on the time. The usual method adopted in these kinds of problems consists of transforming Ω(t)\Omega(t) in a bounded reference domain Ω0N\Omega_{0}\subset\mathbb{R}^{N}. Such transformation is then used in order to bring back the Schrödinger equation (1.1) in an equivalent equation in the phase space L2(Ω0,)L^{2}(\Omega_{0},\mathbb{C}), which does not depend on the time. To this purpose, one can introduce a family of diffeomorphisms (h(t,))tI(h(t,\cdot))_{t\in I} such that for each tIt\in I, h(t,)h(t,\cdot) is a 𝒞p\mathcal{C}^{p}-diffeomorphism from Ω¯0\overline{\Omega}_{0} onto Ω¯(t)\overline{\Omega}(t) with p1p\geq 1 (see Figure 1). Assume in addition that the function tIh(t,y)Nt\in I\mapsto h(t,y)\in\mathbb{R}^{N} is of class 𝒞q\mathcal{C}^{q} with respect to the time with q1q\geq 1.

Refer to captionΩ(t)\Omega(t)Ω0\Omega_{0}h(t,)h(t,\cdot)
Figure 1: The family of diffeomorphisms (h(t,))tI(h(t,\cdot))_{t\in I} which enables to go back to a fixed domain Ω0\Omega_{0}.

In order to bring back the Schrödinger equation (1.1) in an equivalent equation in L2(Ω0,)L^{2}(\Omega_{0},\mathbb{C}), one can introduce the pullback operator

h(t):ϕL2(Ω(t),)ϕh=ϕ(h(t,))L2(Ω0,)h^{*}(t)\leavevmode\nobreak\ :\leavevmode\nobreak\ \phi\in L^{2}(\Omega(t),\mathbb{C})\leavevmode\nobreak\ \longmapsto\leavevmode\nobreak\ \phi\circ h=\phi(h(t,\cdot))\in L^{2}(\Omega_{0},\mathbb{C}) (1.2)

and its inverse, the pushforward operator, defined by

h(t):ψL2(Ω0,)ψh1=ψ(h1(t,))L2(Ω(t),).h_{*}(t)\leavevmode\nobreak\ :\leavevmode\nobreak\ \psi\in L^{2}(\Omega_{0},\mathbb{C})\leavevmode\nobreak\ \longmapsto\leavevmode\nobreak\ \psi\circ h^{-1}=\psi(h^{-1}(t,\cdot))\in L^{2}(\Omega(t),\mathbb{C})\leavevmode\nobreak\ . (1.3)

If we compute the equation satisfied by w=huw=h^{*}u when uu is solution of (1.1) at least in a formal sense, then we find that (1.1) becomes

itw(t,y)=1|J|divy(|J|J1(J1)tyw(t,y))+iJ1th(t,y)|yw(t,y),tI,yΩ0,i\partial_{t}w(t,y)\leavevmode\nobreak\ =\leavevmode\nobreak\ -\frac{1}{|J|}\operatorname{div}_{y}\Big{(}|J|J^{-1}(J^{-1})^{t}\nabla_{y}w(t,y)\Big{)}+i\langle J^{-1}\partial_{t}h(t,y)|\nabla_{y}w(t,y)\rangle,\ \ \ \ \ \ \ \ t\in I\leavevmode\nobreak\ ,\leavevmode\nobreak\ \leavevmode\nobreak\ y\in\Omega_{0}, (1.4)

where J=J(t,y)=Dyh(t,y)J=J(t,y)=D_{y}h(t,y) is the Jacobian matrix of hh, |J||J| stands for |det(J)||\det(J)| and |\langle\cdot|\cdot\rangle corresponds to the scalar product in N\mathbb{C}^{N} (see Remark 2.4 in Section 2 for the computations).

A possible way to give a sense to the Schrödinger equation (1.1) consists in proving that (1.4), endowed with some boundary conditions, generates a well-posed flow in L2(Ω0,)L^{2}(\Omega_{0},\mathbb{C}). This can be locally done by exploiting some specific properties of the Schrödinger equation with perturbative terms of order one (see [34, 35, 36, 39]). Nevertheless this method presents possible disadvantages for our purposes. Indeed, such approach does not provide the natural conservation of the L2L^{2}-norm and the Hamiltonian structure of the equation is lost in the sense that the new differential operator is no longer self-adjoint with respect to a natural time-independent hermitian structure. This fact represents an obstruction, not only to the proof of global existence of solutions, but also to the use of different techniques such as the adiabatic theory.

We are interested in studying the Schrödinger equation (1.1) by preserving its Hamiltonian structure. From this perspective, it is natural to introduce the following unitary operator h(t)h^{\sharp}(t) defined by

h(t):ϕL2(Ω(t),)|J(,t)|(ϕh)(t)L2(Ω0,).h^{\sharp}(t)\leavevmode\nobreak\ :\leavevmode\nobreak\ \phi\in L^{2}(\Omega(t),\mathbb{C})\leavevmode\nobreak\ \longmapsto\leavevmode\nobreak\ \sqrt{|J(\cdot,t)|}\,(\phi\circ h)(t)\in L^{2}(\Omega_{0},\mathbb{C})\leavevmode\nobreak\ . (1.5)

We also denote by h(t)h_{\sharp}(t) its inverse

h(t)=(h(t))1:ψ(ψ/|J(,t)|)h1.h_{\sharp}(t)=(h^{\sharp}(t))^{-1}\leavevmode\nobreak\ :\leavevmode\nobreak\ \psi\mapsto\big{(}\psi/\sqrt{|J(\cdot,t)|}\big{)}\circ h^{-1}\leavevmode\nobreak\ . (1.6)

Notice that the relation h(t)uL2(Ω0)=uL2(Ω(t))\|h^{\sharp}(t)u\|_{L^{2}(\Omega_{0})}=\|u\|_{L^{2}(\Omega(t))} enables to transport the conservative structure through the change of variables. A direct computation, provided in Section 3.1, shows that uu solves (1.1) if and only if v=huv=h^{\sharp}u is solution of

itv(t,y)=\displaystyle i\partial_{t}v(t,y)= 1|J|divy(|J|J1(J1)ty(v(t,y)|J|))\displaystyle-\frac{1}{\sqrt{|J|}}\operatorname{div}_{y}\Bigg{(}|J|J^{-1}(J^{-1})^{t}\nabla_{y}\Big{(}\frac{v(t,y)}{\sqrt{|J|}}\Big{)}\Bigg{)}
+i2t(|J(t,y)|)|J|v+i|J|J1th(t,y)|yv(t,y)|J|,tI,yΩ0.\displaystyle+\frac{i}{2}\frac{\partial_{t}\big{(}|J(t,y)|\big{)}}{|J|}v\leavevmode\nobreak\ +\leavevmode\nobreak\ i{\sqrt{|J|}}\langle J^{-1}\partial_{t}h(t,y)|\nabla_{y}\frac{v(t,y)}{\sqrt{|J|}}\rangle,\ \ \ \ \ \ \ \ t\in I\leavevmode\nobreak\ ,\leavevmode\nobreak\ \leavevmode\nobreak\ y\in\Omega_{0}. (1.7)

Written as it stands, this equation is not easy to handle. For instance, it is unclear whether the equation is of Hamiltonian type and how to compute its spectrum. The central argument of this paper is to show that Equation (1.7) can be rewritten in the form

itv(t,y)=h[(divx+iAh)(x+iAh)+|Ah|2]hv(t,y),tI,yΩ0,i\partial_{t}v(t,y)=-h^{\sharp}\Big{[}\big{(}\operatorname{div}_{x}+iA_{h}\big{)}\circ\big{(}\nabla_{x}+iA_{h}\big{)}+|A_{h}|^{2}\Big{]}h_{\sharp}v(t,y),\ \ \ \ \ \ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ t\in I,\leavevmode\nobreak\ \leavevmode\nobreak\ y\in\Omega_{0}, (1.8)

with Ah(t,x)=12(hth)(t,x)A_{h}(t,x)=-\frac{1}{2}(h_{*}\partial_{t}h)(t,x). Now, the operator appearing in the last equation is the conjugate with respect to hh^{\sharp} and hh_{\sharp} of an explicit magnetic Laplacian. Thus, its Hamiltonian structure becomes obvious and some of its properties, as the spectrum, may be easier to study. We refer to [10, 20, 21, 29, 30, 46] for different spectral results on magnetic Laplacian operators.

Using the unitary operator hh^{\sharp} rather than hh^{*} is natural and it was already done in the literature in some very specific frameworks in [3, 5, 43]. The same idea was also adopted to study quantum waveguides in the time-independent framework, where the magnetic field AhA_{h} does not appear, see for instance [19, 26]. In [22, 23], the transformation hh^{*} is used on manifolds with time-varying metrics. The authors assume hh preserving the volumes which yields h=hh^{*}=h^{\sharp}. They obtain an operator involving a magnetic field similar to the one in (1.8). From this perspective, the relation between motion and magnetic field is not surprising. Physically, the momentum p=mvp=mv of a moving particle of mass mm, velocity vv and charge qq in a magnetic field AA must be replaced by p~=mv+qA\widetilde{p}=mv+qA. This corresponds to the magnetic field appearing in the equation (1.8) for the Galilean frames. An explicit example of the link between motion and magnetic field in our results can be seen in the boundary condition on a moving surface as in Figure 2 below. This is also related to the notion of “anti-convective derivative” of Henry, see [31].


The Dirichlet boundary condition
Once Equation (1.8) is obtained, the Cauchy problem is easy to handle by using classical results of existence of unitary flows generated by time-dependent Hamiltonians. In our work, we refer to Theorem A.1 presented in the appendix. In the case of the simple Laplacian operator with Dirichlet boundary condition, we obtain our first main result which states the following.

Theorem 1.1.

Let II\subset\mathbb{R} be an interval of times and let {Ω(t)}tIN\{\Omega(t)\}_{t\in I}\subset\mathbb{R}^{N} with NN\in\mathbb{N}^{*} be a family of domains. Assume that there exist a bounded reference domain Ω0\Omega_{0} in N\mathbb{R}^{N} and a family of diffeomorphisms (h(t,))tI𝒞2(I×Ω¯0,N)(h(t,\cdot))_{t\in I}\in\mathcal{C}^{2}(I\times\overline{\Omega}_{0},\mathbb{R}^{N}) such that h(t,Ω¯0)=Ω¯(t)h(t,\overline{\Omega}_{0})=\overline{\Omega}(t).

Then, Equation (1.8) endowed with Dirichlet boundary conditions generates a unitary flow U~(t,s)\tilde{U}(t,s) on L2(Ω0)L^{2}(\Omega_{0}) and we may define weak solutions of the Schrödinger equation

{itu(t,x)=Δxu(t,x),tI,xΩ(t)u|Ω(t)0\left\{\begin{array}[]{ll}i\partial_{t}u(t,x)=-\Delta_{x}u(t,x),&t\in I\leavevmode\nobreak\ ,\leavevmode\nobreak\ \leavevmode\nobreak\ x\in\Omega(t)\\ u_{|\partial\Omega(t)}\equiv 0&\end{array}\right. (1.9)

by transporting this flow via hh_{\sharp} to a unitary flow U(t,s):L2(Ω(s))L2(Ω(t))U(t,s):L^{2}(\Omega(s))\rightarrow L^{2}(\Omega(t)).

Assume in addition that the diffeomorphisms hh are of class 𝒞3\mathcal{C}^{3} with respect to the time and the space variable. Then, for any u0H2(Ω(t0))H01(Ω(t0))u_{0}\in H^{2}(\Omega(t_{0}))\cap H^{1}_{0}(\Omega(t_{0})), the above flow defines a solution u(t)=U(t,t0)u0u(t)=U(t,t_{0})u_{0} in 𝒞0(I,H2(Ω(t))H01(Ω(t)))𝒞1(I,L2(Ω(t)))\mathcal{C}^{0}(I,H^{2}(\Omega(t))\cap H^{1}_{0}(\Omega(t)))\cap\mathcal{C}^{1}(I,L^{2}(\Omega(t))) solving (1.9) in the L2L^{2}-sense.

Theorem 1.1 is consequence of the stronger result of Theorem 3.1 (Section 3.1) where we also include diffusion coefficients as well as magnetic and electric potentials. However, in this introduction, we consider the case of the free Laplacian to simplify the notations and to avoid further technicalities

Also notice that the above result does not require the reference domain Ω0\Omega_{0} to have any regularity. In particular, it may have corners such as a rectangle for example. Of course, all the domains Ω(t)\Omega(t) have to be diffeomorphic to Ω0\Omega_{0}. Hence, we cannot create singular perturbations such as adding or removing corners and holes. Nevertheless, Ω(t)\Omega(t) may typically be a family of rectangles or cylinders with different proportions.


The magnetic Neumann condition
At first sight, one may naturally think to associate to the Schrödinger Equation (1.1) with the homogenenous Neumann boundary conditions νu(t,x)=0\partial_{\nu}u(t,x)=0 where xΩ(t)x\in\partial\Omega(t) and ν\nu is the unit outward normal of Ω(t)\partial\Omega(t). However, these conditions cannot generate a unitary evolution, which is problematic for quantum dynamics. Simply consider the solution u1u\equiv 1, whose norm depends on the volume of Ω(t)\Omega(t). Even when the volume of Ω(t)\Omega(t) is constant, if uu solves (1.1) with homogeneous Neumann boundary conditions, then the computation (1.12) below shows that the evolution cannot be unitary, except when hthh_{*}\partial_{t}h is tangent to Ω(t)\partial\Omega(t) at all the boundary points, meaning that the shape of Ω(t)\Omega(t) is in fact unchanged.

From this perspective, another interesting aspect of our result appears. The expression of Equation (1.8) indicates that the correct boundary conditions to consider are the ones associated with Neumann realization of the magnetic Laplacian operator, that are

ν(hv)+iν|Ah(hv)=0 on Ω(t).\partial_{\nu}(h_{\sharp}v)+i\langle\nu|A_{h}\rangle(h_{\sharp}v)=0\leavevmode\nobreak\ \leavevmode\nobreak\ \text{ on }\partial\Omega(t). (1.10)

If we denote by ν0\nu_{0} the unit outward normal of Ω0\partial\Omega_{0}, then the last identity can be transposed in

(J1)tν0|(J1)t|J|y(v|J|)i2(th)v=0 on Ω0\Big{\langle}(J^{-1})^{t}\nu_{0}\Big{|}(J^{-1})^{t}\sqrt{|J|}\nabla_{y}\big{(}\frac{v}{\sqrt{|J|}}\big{)}-\frac{i}{2}(\partial_{t}h)v\Big{\rangle}=0\leavevmode\nobreak\ \leavevmode\nobreak\ \text{ on }\partial\Omega_{0}\leavevmode\nobreak\ (1.11)

(see Remark 3.3 for further details on the computations). Even though the conditions seems complicated on Ω0\Omega_{0}, they simply write as the classical magnetic Neumann boundary conditions for the original problem in Ω(t)\Omega(t), see (1.13) below. In particular, they exactly correspond to the ones of a planar wave bouncing off the moving surface, as it is clear in the example of Figure 2. We also notice that they perfectly match with the preservation of the energy since if u(t)u(t) solves (1.1) at least formally, then

tΩ(t)|u(t,x)|2dx\displaystyle\partial_{t}\int_{\Omega(t)}|u(t,x)|^{2}\,{\text{\rm d}}x =Ω(t)ν|hth|u(t,x)|2dx+2Ω(t)tu(t,x)u¯(t,x)dx\displaystyle=\int_{\partial\Omega(t)}\langle\nu|h_{*}\partial_{t}h\rangle|u(t,x)|^{2}\,{\text{\rm d}}x+2\Re\int_{\Omega(t)}\partial_{t}u(t,x)\overline{u}(t,x)\,{\text{\rm d}}x
=Ω(t)ν|hth|u(t,x)|2dx+2(iΩ(t)νu(t,x)u¯(t,x)dx).\displaystyle=\int_{\partial\Omega(t)}\langle\nu|h_{*}\partial_{t}h\rangle|u(t,x)|^{2}\,{\text{\rm d}}x+2\Re\left(i\int_{\partial\Omega(t)}\partial_{\nu}u(t,x)\overline{u}(t,x)\,{\text{\rm d}}x\right). (1.12)

Once the correct boundary condition is inferred, we obtain the following result in the same way as the Dirichlet case.

Theorem 1.2.

Let II\subset\mathbb{R} be an interval of times and let {Ω(t)}tIN\{\Omega(t)\}_{t\in I}\subset\mathbb{R}^{N} with NN\in\mathbb{N}^{*} be a family of domains. Assume that there exist a bounded reference domain Ω0\Omega_{0} in N\mathbb{R}^{N} of class 𝒞1\mathcal{C}^{1} and a family of diffeomorphisms (h(t,))tI𝒞2(I×Ω¯0,N)(h(t,\cdot))_{t\in I}\in\mathcal{C}^{2}(I\times\overline{\Omega}_{0},\mathbb{R}^{N}) such that h(t,Ω¯0)=Ω¯(t)h(t,\overline{\Omega}_{0})=\overline{\Omega}(t).

Then, Equation (1.8) endowed with the magnetic Neumann boundary conditions (1.10) (or equivalently (1.11)) generates a unitary flow U~(t,s)\tilde{U}(t,s) on L2(Ω0)L^{2}(\Omega_{0}) and we may define weak solutions of the Schrödinger equation

{itu(t,x)=Δxu(t,x),tI,xΩ(t)νu(t,x)i2ν|hth(t,x)u(t,x)=0,tI,xΩ(t)\left\{\begin{array}[]{ll}i\partial_{t}u(t,x)=-\Delta_{x}u(t,x),&t\in I\leavevmode\nobreak\ ,\leavevmode\nobreak\ \leavevmode\nobreak\ x\in\Omega(t)\\ \partial_{\nu}u(t,x)-\frac{i}{2}\langle\nu|h_{*}\partial_{t}h(t,x)\rangle u(t,x)=0,&t\in I\leavevmode\nobreak\ ,\leavevmode\nobreak\ \leavevmode\nobreak\ x\in\partial\Omega(t)\end{array}\right. (1.13)

by transporting this flow via hh_{\sharp} to a unitary flow U(t,s):L2(Ω(s))L2(Ω(t))U(t,s):L^{2}(\Omega(s))\rightarrow L^{2}(\Omega(t)).

Assume in addition that the diffeomorphisms hh are of class 𝒞3\mathcal{C}^{3} with respect to the time and the space variable. Then, for any u0H2(Ω(t0))u_{0}\in H^{2}(\Omega(t_{0})) satisfying the magnetic Neumann boundary condition of (1.13), the above flow defines a solution u(t)=U(t,t0)u0u(t)=U(t,t_{0})u_{0} in 𝒞0(I,H2(Ω(t)))𝒞1(I,L2(Ω(t)))\mathcal{C}^{0}(I,H^{2}(\Omega(t)))\cap\mathcal{C}^{1}(I,L^{2}(\Omega(t))) solving (1.13) in the L2L^{2}-sense and satisfying the magnetic Neumann boundary condition.

Refer to captionx1=(t)x_{1}=\ell(t)νu=0\partial_{\nu}u=0itu=Δui\partial_{t}u=-\Delta ux1=0x_{1}=0νu=i2(t)u\partial_{\nu}u=\frac{i}{2}\ell^{\prime}(t)u
Figure 2: The correct Neumann boundary conditions for the Schrödinger equation in a cylinder with a moving end. Notice the magnetic Neumann boundary condition at the moving surface, even though the equation has no magnetic term. See Section 5 for further details on the computations.

Gauge transformation
As it is well known, the magnetic potential has a gauge invariance. In particular, for any ϕ\phi of class 𝒞1\mathcal{C}^{1} in space, we have

eiϕ(x)[(x+iAh)2]eiϕ(x)=(x+i(Ah+xϕ))2.e^{-i\phi(x)}\big{[}(\nabla_{x}+iA_{h})^{2}\big{]}e^{i\phi(x)}\leavevmode\nobreak\ =\leavevmode\nobreak\ \big{(}\nabla_{x}+i(A_{h}+\nabla_{x}\phi)\big{)}^{2}\leavevmode\nobreak\ . (1.14)

Thus, it is possible to delete the magnetic term Ah=12hthA_{h}=-\frac{1}{2}h_{*}\partial_{t}h by the change of gauge when there exists ϕ\phi of class 𝒞1\mathcal{C}^{1} such that

tI,xΩ(t),(hth)(t,x)=2xϕ(t,x).\forall t\in I\leavevmode\nobreak\ ,\leavevmode\nobreak\ \leavevmode\nobreak\ \forall x\in\Omega(t)\leavevmode\nobreak\ ,\leavevmode\nobreak\ \leavevmode\nobreak\ (h_{*}\partial_{t}h)(t,x)=2\nabla_{x}\phi(t,x)\leavevmode\nobreak\ .

In such context, the well-posedness of the equations (1.9) and (1.13) can be investigated by considering

w(t,y)=heiϕ(t,x)u:=|J(t,y)|eiϕ(t,h(t,y))u(t,h(t,y)).w(t,y)=h^{\sharp}e^{-i\phi(t,x)}u:=\sqrt{|J(t,y)|}e^{-i\phi(t,h(t,y))}u(t,h(t,y)).

and by studying the solution of the following equation endowed with the corresponding boundary conditions

itw(t,y)=(hΔxh+14|th(t,y)|2t(hϕ)(t,y))w(t,y),tI,yΩ0.i\partial_{t}w(t,y)=-\Big{(}h^{\sharp}\Delta_{x}h_{\sharp}+\frac{1}{4}|\partial_{t}h(t,y)|^{2}-\partial_{t}(h^{*}\phi)(t,y)\Big{)}w(t,y),\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ t\in I\leavevmode\nobreak\ ,\leavevmode\nobreak\ \leavevmode\nobreak\ y\in\Omega_{0}. (1.15)

The gauge transformation, not only simplifies Equation (1.8), but also yields a gain of regularity in the hypotheses on hh adopted in the theorems 1.1 and 1.2. This fact follows as the main part of the new Hamiltonian in (1.15) does not contain th\partial_{t}h anymore. In details, if we consider h𝒞t1(I,𝒞x2(Ω0,N))h\in\mathcal{C}^{1}_{t}(I,\mathcal{C}^{2}_{x}(\Omega_{0},\mathbb{R}^{N})), then the existence of weak solutions of (1.9) and (1.13) can be guaranteed when ϕ\phi is of class 𝒞3\mathcal{C}^{3} in space and W1,W^{1,\infty} in time. The existence of strong solutions, instead, holds when ϕ\phi is of class 𝒞4\mathcal{C}^{4} in space and W1,W^{1,\infty} in time.

Also remark that the gauge transformation is not always possible to use. For example, if Ω(t)\Omega(t) is a rotation of a square, th\partial_{t}h is not curl-free and cannot be rectified due to the presence of corners at which h(t,y)h(t,y) is imposed (corners have to be send onto corners). Finally, we may also use the simpler gauge of the electric potential if some of the terms of (1.15) are constant, see Section 5.


Moser’s trick
Another way to simplify Equation (1.8) is to use a family of diffeomorphisms h~(t)\tilde{h}(t) such that the determinant of the Jacobian is independent of yy. In other words, when J~=Dyh~\tilde{J}=D_{y}\tilde{h} satisfies the following identity

tI,yΩ0,|J~(t,y)|:=a(t).\forall t\in I\leavevmode\nobreak\ ,\leavevmode\nobreak\ \leavevmode\nobreak\ \forall y\in\Omega_{0}\leavevmode\nobreak\ ,\leavevmode\nobreak\ \leavevmode\nobreak\ |\tilde{J}(t,y)|:=a(t)\leavevmode\nobreak\ . (1.16)

In this case, the multiplication for the Jacobian JJ of commutes with the spatial derivatives and then, Equations (1.7) and (1.8) can be simplified in the following expression, for Ah~=12h~th~A_{\tilde{h}}=-\frac{1}{2}\tilde{h}_{*}\partial_{t}\tilde{h},

itv(t,y)\displaystyle i\partial_{t}v(t,y) =divy(J~1(J~1)tyv)+i2a(t)a(t)v+iJ~1th~|yv\displaystyle=-\operatorname{div}_{y}(\tilde{J}^{-1}(\tilde{J}^{-1})^{t}\nabla_{y}v)+\frac{i}{2}\frac{a^{\prime}(t)}{a(t)}v+i\langle\tilde{J}^{-1}\partial_{t}\tilde{h}|\nabla_{y}v\rangle
=h~[(divx+iAh~)(x+iAh~)+|Ah~|2]h~v(t,y).\displaystyle=-\tilde{h}^{*}\big{[}(\operatorname{div}_{x}+iA_{\tilde{h}})\circ(\nabla_{x}+iA_{\tilde{h}})+|A_{\tilde{h}}|^{2}\big{]}\tilde{h}_{*}v(t,y). (1.17)

This strategy can be used, not only to simplify the equations, but also to gain some regularity since |J~||\tilde{J}| is now constant and thus smooth. Therefore, h~\tilde{h}^{\sharp} maps Hk(Ω(t))H^{k}(\Omega(t)) into Hk(Ω0)H^{k}(\Omega_{0}) as soon as h~\tilde{h} is of class 𝒞k\mathcal{C}^{k} in space.

For tt fixed, finding a diffeomorphism h~\tilde{h} satisfying the identity (1.16) follows from a very famous work of Moser [42]. This kind of result called “Moser’s trick” was widely studied in literature even in the case of moving domains (see Section 4.1). Nevertheless, most of these outcomes are not interested in studying the optimal time and space regularity as well as their proofs are sometimes simply outlined. For this purpose, in Section 4, we prove the following result by following the arguments of [16].

Theorem 1.3.

Let k1k\geq 1, and rr\in\mathbb{N} with kr0k\geq r\geq 0. Let α(0,1)\alpha\in(0,1) and let Ω0N\Omega_{0}\subset\mathbb{R}^{N} be a connected bounded domain of class 𝒞k+2,α\mathcal{C}^{k+2,\alpha}. Let II\subset\mathbb{R} an interval of times and assume that there exists a family (Ω(t))tI(\Omega(t))_{t\in I} of domains such that there exists a family (h(t))tI(h(t))_{t\in I} of diffeomorphisms

h:yΩ¯0h(t,y)Ω(t)¯h\leavevmode\nobreak\ :\leavevmode\nobreak\ y\in\overline{\Omega}_{0}\leavevmode\nobreak\ \longrightarrow\leavevmode\nobreak\ h(t,y)\in\overline{\Omega(t)}

which are of class 𝒞k,α(Ω¯0,Ω(t)¯)\mathcal{C}^{k,\alpha}(\overline{\Omega}_{0},\overline{\Omega(t)}) with respect to yy and of class 𝒞r(Ω¯0,N)\mathcal{C}^{r}(\overline{\Omega}_{0},\mathbb{R}^{N}) with respect to tt.

Then, there exists a family (h~(t))tI(\tilde{h}(t))_{t\in I} of diffeomorphisms from Ω¯0\overline{\Omega}_{0} onto Ω(t)¯\overline{\Omega(t)}, with the same regularity as hh, and such that det(Dyh~(t))\det(D_{y}\tilde{h}(t)) is constant with respect to yy, that is that

yΩ0,det(Dyh~(t,y))=meas(Ω(t))meas(Ω0).\forall y\in\Omega_{0}\leavevmode\nobreak\ ,\leavevmode\nobreak\ \leavevmode\nobreak\ \det(D_{y}\tilde{h}(t,y))=\frac{{\text{\rm meas}}(\Omega(t))}{{\text{\rm meas}}(\Omega_{0})}\leavevmode\nobreak\ .

Even though results similar to Theorem 1.3 have already been stated, the fact that h(t)h(t) may simply be continuous with respect to the time and that h~(t)\tilde{h}(t) has the same regularity of h(t)h(t) seems to be new. There are some simple cases where we can define explicit h~\tilde{h} as in dimension N=1N=1 or in the examples of Section 5, but this is not aways possible. In these last situations, Equation (1.17) may be difficult to use as h~\tilde{h} is not explicit. However, Equation (1.17) yields a gain of regularity in the Cauchy problem which we resume in the following corollary.

Corollary 1.4.

Let II\subset\mathbb{R} be an interval of times and let {Ω(t)}tIN\{\Omega(t)\}_{t\in I}\subset\mathbb{R}^{N} with NN\in\mathbb{N}^{*} be a family of domains. Assume that there exist a bounded reference domain Ω0\Omega_{0} in N\mathbb{R}^{N} of class 𝒞4,α\mathcal{C}^{4,\alpha} with α(0,1)\alpha\in(0,1) and a family of diffeomorphisms (h(t,))tI𝒞t2(I,𝒞x1(Ω0,N))(h(t,\cdot))_{t\in I}\in\mathcal{C}^{2}_{t}(I,\mathcal{C}^{1}_{x}(\Omega_{0},\mathbb{R}^{N})) such that h(t,Ω¯0)=Ω¯(t)h(t,\overline{\Omega}_{0})=\overline{\Omega}(t).

Then, we may define as in Theorems 1.1 and 1.2 weak solutions of the above equations (1.9) and (1.13) by considering v=h~uv=\tilde{h}_{\sharp}u, with h~\tilde{h} as in Theorem 1.3, which solves the Schrödinger equation (1.17) with the corresponding boundary conditions and by transporting the flow of (1.17) via the above change of variable.

Assume in addition that hh belongs to 𝒞t3(I,𝒞y2(Ω0,N))\mathcal{C}^{3}_{t}(I,\mathcal{C}^{2}_{y}(\Omega_{0},\mathbb{R}^{N})). Then, for any u0H2(Ω(t0))H01(Ω(t0))u_{0}\in H^{2}(\Omega(t_{0}))\cap H^{1}_{0}(\Omega(t_{0})), the above flow defines a solution u(t)u(t) in the space 𝒞0(I,H2(Ω(t))H01(Ω(t)))𝒞1(I,L2(Ω(t)))\mathcal{C}^{0}(I,H^{2}(\Omega(t))\cap H^{1}_{0}(\Omega(t)))\cap\mathcal{C}^{1}(I,L^{2}(\Omega(t))) solving (1.9) or (1.13) in the L2L^{2}-sense.

Corollary 1.4 follows from the same arguments leading to the theorems 1.1 or 1.2 (see Section 3). The only difference is the gain of regularity in space. Indeed, the term |J(t,y)||J(t,y)| appearing in hh^{\sharp} or hh_{\sharp} is replaced by |J~(t,y)|=meas(Ω(t))meas(Ω0)=a(t)|\tilde{J}(t,y)|=\frac{{\text{\rm meas}}(\Omega(t))}{{\text{\rm meas}}(\Omega_{0})}=a(t) which is constant in space and then smooth. To this end, we simply have to replace the first family of diffeomorphisms h(t)h(t) by the one given by Theorem 1.3.

Notice that, in Corollary 1.4, we have to assume that the reference domain Ω0\Omega_{0} is smooth. If Ω(t)\Omega(t) are simply of class 𝒞1\mathcal{C}^{1} or 𝒞2\mathcal{C}^{2}, this is not a real restriction since we may choose a smooth reference domain and a not so smooth diffeomorphism hh. In the case where Ω(t)\Omega(t) has corner, as rectangles for example, then Corollary 1.4 do not formally apply. However, in the case of moving rectangles, finding a family of explicit diffeomorphisms h(t)h(t) satisfying (1.16) is easy and the arguments behind Corollary 1.4 can be directly used, see the computations of Section 5.


An example of application: an adiabatic result
Consider a family of domains {Ω(τ)}τ[0,1]\{\Omega(\tau)\}_{\tau\in[0,1]} of N\mathbb{R}^{N} such that there exist a bounded reference domain Ω0N\Omega_{0}\subset\mathbb{R}^{N} and a family of diffeomorphisms (h(τ,))τ[0,1]𝒞3([0,1]×Ω¯0,N)(h(\tau,\cdot))_{\tau\in[0,1]}\in\mathcal{C}^{3}([0,1]\times\overline{\Omega}_{0},\mathbb{R}^{N}) such that h(τ,Ω¯0)=Ω¯(τ)h(\tau,\overline{\Omega}_{0})=\overline{\Omega}(\tau). Assume that for each τ[0,1]\tau\in[0,1], the Dirichlet Laplacian operator Δ\Delta on Ω(τ)\Omega(\tau) has a simple isolated eigenvalue λ(τ)\lambda(\tau) with normalized eigenfunction φ(τ)\varphi(\tau), associated with a spectral projector P(τ)P(\tau), all three depending continuously on τ\tau. Following the classical adiabatic principle, we expect that if we start with a quantum state in Ω(0)\Omega(0) close to φ(0)\varphi(0) and we deform very slowly the domain to the shape Ω(1)\Omega(1), then the final quantum state should be close to φ(1)\varphi(1) up to a phase shift (see Figure 3). The slowness of the deformation is represented by a parameter ϵ>0\epsilon>0 and we consider deformations between the times t=0t=0 and t=1/ϵt=1/\epsilon, that is the following Schrödinger equation

{ituϵ(t,x)=Δuϵ(t,x),t[0,1/ϵ],xΩ(ϵt)uϵ(t)0, on Ω(ϵt)uϵ(t=0)=u0L2(Ω(0))\left\{\begin{array}[]{ll}i\partial_{t}u_{\epsilon}(t,x)=-\Delta u_{\epsilon}(t,x),&t\in[0,1/\epsilon]\leavevmode\nobreak\ ,\leavevmode\nobreak\ \leavevmode\nobreak\ x\in\Omega(\epsilon t)\\ u_{\epsilon}(t)\equiv 0,&\text{ on }\partial\Omega(\epsilon t)\\ u_{\epsilon}(t=0)=u_{0}\in L^{2}(\Omega(0))&\end{array}\right. (1.18)

Due to Theorem 1.1, we know how to define a solution of (1.18). Our main equation (1.8) provides the Hamiltonian structure required for applying the adiabatic theory, in contrast to the case of Equation (1.4). It also indicates that we should not consider the Laplacian operators Δ\Delta and its spectrum, but rather the magnetic Laplacian operators of Equation (1.8). With these hints, it is not difficult to adapt the classical adiabatic methods to obtain the following result (see Section 5).

Corollary 1.5.

Consider the above framework. Then, we have

P(1)uϵ(1/ϵ)|uϵ(1/ϵ)ϵ0P(0)u0|u0.\langle P(1)u_{\epsilon}(1/\epsilon)|u_{\epsilon}(1/\epsilon)\rangle\leavevmode\nobreak\ \leavevmode\nobreak\ \xrightarrow[\leavevmode\nobreak\ \leavevmode\nobreak\ \epsilon\longrightarrow 0\leavevmode\nobreak\ \leavevmode\nobreak\ ]{}\leavevmode\nobreak\ \leavevmode\nobreak\ \langle P(0)u_{0}|u_{0}\rangle\leavevmode\nobreak\ .
Refer to captionu0u_{0}very slow deformation of the domainuϵ(1/ϵ)u_{\epsilon}(1/\epsilon)Ω(1)\Omega(1)Ω(0)\Omega(0)
Figure 3: The ground state of a domain Ω(0)\Omega(0) can be transformed into the ground state of a domain Ω(1)\Omega(1) if we slowly and smoothly deform Ω(0)\Omega(0) to Ω(1)\Omega(1) while the quantum state evolves following Schrödinger equation.

Acknowledgements: the present work has been conceived in the vibrant atmosphere of the Workshop-Summer School of Benasque (Spain) “VIII Partial differential equations, optimal design and numerics”. The authors would like to thank Yves Colin de Verdière and Andrea Seppi for the fruitful discussions on the geometric aspects of the work. They are also grateful to Gérard Besson and Emmanuel Russ for the suggestions on the proof of the Moser’s trick and to Alain Joye for the advice on the adiabatic theorem. This work was financially supported by the project ISDEEC ANR-16-CE40-0013.

2 Moving domains and change of variables

The study of the influence of the domain shape on a PDE problem has a long history, in particular in shape optimization. The interested reader may consider [31] or [44] for example. The basic strategy of these kinds of works is to bring back the problem in a fixed reference domain Ω0\Omega_{0} via diffeomorphisms. We thus have to compute the new differential operators in the new coordinates. The formulae presented in this section are well known (see [31] for example). We recall them for sake of completeness and also to fix the notations.

Let Ω0\Omega_{0} be a reference domain and let h(t)h(t) be a 𝒞2\mathcal{C}^{2}-diffeorphism mapping Ω0\Omega_{0} onto Ω(t)\Omega(t) (see Figure 1). In this section, we only work with fixed tt. For this reason, we omit the time dependence when it is not necessary and forgot the question of regularity with respect to the time.

From now on, we denote by xx the points in Ω(t)\Omega(t) and by yy the ones in Ω0\Omega_{0}. For any matrix AA, |A||A| stands for |det(A)||\det(A)| and |\langle\cdot|\cdot\rangle is the scalar product in N\mathbb{C}^{N} with the convention

v|w=k=1Nvk¯wk,v,wN.\langle v|w\rangle=\sum_{k=1}^{N}\overline{v_{k}}\,w_{k}\leavevmode\nobreak\ ,\ \ \ \ \ \ \ \ \ \forall v,w\in\mathbb{C}^{N}.

We use the pullback and pushforward operators defined by (1.2) and (1.3). We denote by J=J(t,y)=Dyh(t,y)J=J(t,y)=D_{y}h(t,y) the Jacobian matrix of hh. The basic rules of differential calculus give that

hDx(h1)=(Dyh)1:=J1,|J1|=|J|1h^{*}D_{x}(h^{-1})=(D_{y}h)^{-1}:=J^{-1}\leavevmode\nobreak\ ,\ \ \ \ \ \ \ \ |J^{-1}|=|J|^{-1} (2.1)

and

Ω(t)f(x)dx=Ω0|J(t,y)|(hf)(y)dy.\int_{\Omega(t)}f(x)\,{\text{\rm d}}x\leavevmode\nobreak\ =\leavevmode\nobreak\ \int_{\Omega_{0}}|J(t,y)|(h^{*}f)(y)\,{\text{\rm d}}y\leavevmode\nobreak\ . (2.2)

The result below is due to the chain rule which implies the following identity

yj(hf)=yjh|hxf.\partial_{y_{j}}(h^{*}f)=\langle\partial_{y_{j}}h|h^{*}\nabla_{x}f\rangle. (2.3)
Proposition 2.1.

For every gH1(Ω0,)g\in H^{1}(\Omega_{0},\mathbb{C}),

(hxh)g(y)=(J(t,y)1)tyg(y)(h^{*}\nabla_{x}h_{*})g(y)=\big{(}J(t,y)^{-1}\big{)}^{t}\cdot\nabla_{y}g(y)

where (J1)t\big{(}J^{-1}\big{)}^{t} is the transposed matrix of (Dyh)1(D_{y}h)^{-1}.

Proof: We apply (2.3) with f=hgf=h_{*}g in order to obtain the identity yg=J(t,y)th(xf)\nabla_{y}g=J(t,y)^{t}h^{*}(\nabla_{x}f). \square

By duality, we obtain the corresponding property for the divergence.

Proposition 2.2.

For every vector field AH1(Ω0,N)A\in H^{1}(\Omega_{0},\mathbb{C}^{N}),

(hdivxh)A(y)=1|J(t,y)|divy(|J(t,y)|J(y,t)1A(y)).(h^{*}\operatorname{div}_{x}h_{*})A(y)=\frac{1}{|J(t,y)|}\operatorname{div}_{y}\big{(}|J(t,y)|J(y,t)^{-1}A(y)\big{)}.

Let ν\nu and σ\sigma respectively be the unit outward normal and the measure on the boundary of Ω(t)\Omega(t). Let ν0\nu_{0} and σ0\sigma_{0} respectively be the unit outward normal and the measure on the boundary of Ω0\Omega_{0}. For every AH1(Ω0,N)A\in H^{1}(\Omega_{0},\mathbb{C}^{N}) and gH1(Ω0,)g\in H^{1}(\Omega_{0},\mathbb{C}),

Ω(t)ν|(hA)(x)(hg)(x)dσ=Ω0|J(t,y)|ν0|J(t,y)1A(y)g(y)dσ0.\int_{\partial\Omega(t)}\langle\nu|(h_{*}A)(x)\rangle(h_{*}g)(x)\,{\text{\rm d}}\sigma=\int_{\partial\Omega_{0}}|J(t,y)|\big{\langle}\nu_{0}\big{|}J(t,y)^{-1}A(y)\big{\rangle}g(y)\,{\text{\rm d}}\sigma_{0}.

Proof: Let φ𝒞0(Ω0,)\varphi\in\mathcal{C}^{\infty}_{0}(\Omega_{0},\mathbb{C}) be a test function and let AH1(Ω0,N)A\in H^{1}(\Omega_{0},\mathbb{C}^{N}). Applying the divergence theorem and the above formulas, we obtain

Ω0|J(t,y)|(hdivxhA)(y)φ(y)dy\displaystyle\int_{\Omega_{0}}|J(t,y)|(h^{*}\operatorname{div}_{x}h_{*}A)(y)\varphi(y)\,{\text{\rm d}}y =Ω(t)(divx(hA))(x)(hφ)(x)dx\displaystyle=\int_{\Omega(t)}(\operatorname{div}_{x}(h_{*}A))(x)(h_{*}\varphi)(x)\,{\text{\rm d}}x
=Ω(t)hA¯(x)|x(hφ)(x)dx\displaystyle=-\int_{\Omega(t)}\langle h_{*}\overline{A}(x)|\nabla_{x}(h_{*}\varphi)(x)\rangle\,{\text{\rm d}}x
=Ω0|J(y,t)|A¯(y)|h(xhφ)(y)dy\displaystyle=-\int_{\Omega_{0}}|J(y,t)|\big{\langle}\overline{A}(y)|h^{*}(\nabla_{x}h_{*}\varphi)(y)\big{\rangle}\,{\text{\rm d}}y
=Ω0|J(y,t)|A¯(y)|(J(t,y)1)tyφ(y)dy\displaystyle=-\int_{\Omega_{0}}|J(y,t)|\big{\langle}\overline{A}(y)|(J(t,y)^{-1})^{t}\nabla_{y}\varphi(y)\big{\rangle}\,{\text{\rm d}}y
=Ω0|J(y,t)|J(t,y)1A¯(y)|yφ(y)dy\displaystyle=-\int_{\Omega_{0}}\big{\langle}|J(y,t)|J(t,y)^{-1}\overline{A}(y)|\nabla_{y}\varphi(y)\big{\rangle}\,{\text{\rm d}}y
=Ω0divy[|J(y,t)|J(t,y)1A(y)]φ(y)dy.\displaystyle=\int_{\Omega_{0}}\operatorname{div}_{y}\big{[}|J(y,t)|J(t,y)^{-1}A(y)\big{]}\varphi(y)\,{\text{\rm d}}y.

The first statement follows from the density of 𝒞0(Ω0,)\mathcal{C}^{\infty}_{0}(\Omega_{0},\mathbb{C}) in L2(Ω0)L^{2}(\Omega_{0}). The second one instead is proved by considering the border terms appearing when we proceed as above with AH1(Ω(t),N)A\in H^{1}(\Omega(t),\mathbb{C}^{N}) and gH1(Ω(t),)g\in H^{1}(\Omega(t),\mathbb{C}). In this context, we use twice the divergence theorem and we obtain that

Ω0|J(t,y)|(hdivxhA)(y)g(y)dy=\displaystyle\int_{\Omega_{0}}|J(t,y)|(h^{*}\operatorname{div}_{x}h_{*}A)(y)g(y)\,{\text{\rm d}}y= Ω(t)ν|(hA)(x)(hg)(x)dσ\displaystyle\int_{\partial\Omega(t)}\langle\nu|(h_{*}A)(x)\rangle(h_{*}g)(x)\,{\text{\rm d}}\sigma
Ω0|J(t,y)|ν0|J(t,y)1A(y)g(y)dσ0\displaystyle-\int_{\partial\Omega_{0}}|J(t,y)|\big{\langle}\nu_{0}\big{|}J(t,y)^{-1}A(y)\big{\rangle}g(y)\,{\text{\rm d}}\sigma_{0}
+Ω0divy(|J(y,t)|J(t,y)1A(y))g(y)dy.\displaystyle+\int_{\Omega_{0}}\operatorname{div}_{y}\big{(}|J(y,t)|J(t,y)^{-1}A(y)\big{)}g(y)\,{\text{\rm d}}y.

The last relation yields the equality of the boundary terms since the first statement is valid in the L2L^{2} sense. \square

By a direct application of the previous propositions, we obtain the operator associated with Δx=divx(x)\Delta_{x}=\operatorname{div}_{x}(\nabla_{x}\cdot).

Proposition 2.3.

For every gH2(Ω0,)g\in H^{2}(\Omega_{0},\mathbb{C}),

(hΔxh)g(y)=1|J(t,y)|divy(|J(t,y)|J(t,y)1(J(t,y)1)tyg(y)).(h^{*}\Delta_{x}h_{*})g(y)=\frac{1}{|J(t,y)|}\operatorname{div}_{y}\Big{(}|J(t,y)|J(t,y)^{-1}(J(t,y)^{-1})^{t}\nabla_{y}g(y)\Big{)}.
Remark 2.4.

For an illustration, let us check, at least formally, that u(t,x)u(t,x) solves the Schrödinger equation (1.1) in Ω(t)\Omega(t) if and only if w=huw=h^{*}u solves (1.4). We have

itw(t,y)\displaystyle i\partial_{t}w(t,y) =it(u(t,h(t,y)))\displaystyle=i\partial_{t}\big{(}u(t,h(t,y))\big{)}
=(itu)(t,h(t,y))+ith(t,y)|(xu)(t,h(t,y))\displaystyle=(i\partial_{t}u)(t,h(t,y))+i\langle\partial_{t}h(t,y)|(\nabla_{x}u)(t,h(t,y))\rangle
=(Δxu)(t,h(t,y))+ith(t,y)|(xu)(t,h(t,y))\displaystyle=(-\Delta_{x}u)(t,h(t,y))+i\langle\partial_{t}h(t,y)|(\nabla_{x}u)(t,h(t,y))\rangle
=(hΔxhw)(t,y)+ith(t,y)|(hxhw)(t,y)\displaystyle=-(h^{*}\Delta_{x}h_{*}w)(t,y)+i\langle\partial_{t}h(t,y)|(h^{*}\nabla_{x}h_{*}w)(t,y)\rangle
=1|J|divy(|J|J1(J1)tyg(t,y))+ith(t,y)|(J1)tyw(t,y)\displaystyle=-\frac{1}{|J|}\operatorname{div}_{y}\Big{(}|J|J^{-1}(J^{-1})^{t}\nabla_{y}g(t,y)\Big{)}+i\langle\partial_{t}h(t,y)|(J^{-1})^{t}\nabla_{y}w(t,y)\rangle
=1|J|divy(|J|J1(J1)tyg(t,y))+iJ1th(t,y)|yw(t,y).\displaystyle=-\frac{1}{|J|}\operatorname{div}_{y}\Big{(}|J|J^{-1}(J^{-1})^{t}\nabla_{y}g(t,y)\Big{)}+i\langle J^{-1}\partial_{t}h(t,y)|\nabla_{y}w(t,y)\rangle.

3 Main results: The Cauchy problem

3.1 Proof of Theorem 1.1: the Dirichlet case

Let II be a time interval and let Ω(t)=h(t,Ω0)\Omega(t)=h(t,\Omega_{0}) be a family of moving domains. In this section, we consider a second order Hamiltonian operator of the type

H(t)\displaystyle H(t) =[D(t,x)x+iA(t,x)]2+V(t,x)\displaystyle=-\Big{[}D(t,x)\nabla_{x}+iA(t,x)\Big{]}^{2}+V(t,x) (3.1)

where

[D(t,x)x+iA(t,x)]2:=(divx(D(t,x)t),+iA(t,x)|)(D(t,x)x+iA(t,x))\displaystyle\Big{[}D(t,x)\nabla_{x}+iA(t,x)\Big{]}^{2}:=\Big{(}\operatorname{div}_{x}\big{(}D(t,x)^{t}\,\cdot\,),+\,i\langle A(t,x)|\ \cdot\ \rangle\Big{)}\circ\Big{(}D(t,x)\nabla_{x}\,+\,iA(t,x)\Big{)} (3.2)

and

  • D𝒞t2(I,𝒞x1(N,N()))D\in\mathcal{C}^{2}_{t}(I,\mathcal{C}^{1}_{x}(\mathbb{R}^{N},\mathcal{M}_{N}(\mathbb{R}))) are symmetric diffusions coefficients such that there exists α>0\alpha>0 satisfying

    tI,xN,ξN,D(x,t)ξ|D(x,t)ξα|ξ|2;\forall t\in I\leavevmode\nobreak\ ,\leavevmode\nobreak\ \leavevmode\nobreak\ \forall x\in\mathbb{R}^{N}\leavevmode\nobreak\ ,\leavevmode\nobreak\ \leavevmode\nobreak\ \forall\xi\in\mathbb{C}^{N}\leavevmode\nobreak\ ,\leavevmode\nobreak\ \leavevmode\nobreak\ \langle D(x,t)\xi|D(x,t)\xi\rangle\geq\alpha|\xi|^{2}\leavevmode\nobreak\ ;
  • A𝒞t2(I,𝒞x1(N,N))A\in\mathcal{C}^{2}_{t}(I,\mathcal{C}^{1}_{x}(\mathbb{R}^{N},\mathbb{R}^{N})) is a magnetic potential;

  • V𝒞t2(I,Lx(N,))V\in\mathcal{C}^{2}_{t}(I,L^{\infty}_{x}(\mathbb{R}^{N},\mathbb{R})) is an electric potential.

In this subsection, we assume that H(t)H(t) is associated with Dirichlet boundary conditions and then its domain is H2(Ω(t),)H01(Ω(t),)H^{2}(\Omega(t),\mathbb{C})\cap H^{1}_{0}(\Omega(t),\mathbb{C}). Of course, the typical example of such Hamiltonian is the Dirichlet Laplacian H(t)=ΔxH(t)=\Delta_{x} defined on H2(Ω(t),)H01(Ω(t),)H^{2}(\Omega(t),\mathbb{C})\cap H^{1}_{0}(\Omega(t),\mathbb{C}). We consider the equation

{itu(t,x)=H(t)u(t,x),tI,xΩ(t)u|Ω(t)0\left\{\begin{array}[]{l}i\partial_{t}u(t,x)=H(t)u(t,x),\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ t\in I\leavevmode\nobreak\ ,\leavevmode\nobreak\ x\in\Omega(t)\\ u_{|\partial\Omega(t)}\equiv 0\end{array}\right. (3.3)

We use the pullback and pushforward operators hh^{\sharp} and hh_{\sharp} defined by (1.5) and (1.6). We also use the notations of Section 2. Notice that hh^{\sharp} is an isometry from L2(Ω(t))L^{2}(\Omega(t)) onto L2(Ω0)L^{2}(\Omega_{0}). Moreover, if hh is a class 𝒞2\mathcal{C}^{2} with respect to the space, then hh^{\sharp} is an isomorphism from H01(Ω(t))H^{1}_{0}(\Omega(t)) onto H01(Ω0)H^{1}_{0}(\Omega_{0}). We set

v(t,y):=(hu)(t,y)=|J(t,y)|u(t,h(t,y)).v(t,y)\leavevmode\nobreak\ :=\leavevmode\nobreak\ (h^{\sharp}u)(t,y)\leavevmode\nobreak\ =\leavevmode\nobreak\ \sqrt{|J(t,y)|}\,u(t,h(t,y))\leavevmode\nobreak\ . (3.4)

At least formally, if uu is solution of (3.3), then a direct computation shows that

itv(t,y)\displaystyle i\partial_{t}v(t,y) =h(itu)(t,y)+i2t(|J(t,y)|)|J(t,y)|u(t,y)+i|J(t,y)|th(t,y)|h(xu)\displaystyle=h^{\sharp}(i\partial_{t}u)(t,y)+\frac{i}{2}\frac{\partial_{t}\big{(}|J(t,y)|\big{)}}{\sqrt{|J(t,y)|}}u(t,y)+i\sqrt{|J(t,y)|}\big{\langle}\partial_{t}h(t,y)|h^{*}(\nabla_{x}u)\big{\rangle}
=(hH(t)h)v(t,y)+i2t(|J(t,y)|)|J(t,y)|v(t,y)+ith(t,y)|(hxh)v(t,y)\displaystyle=\big{(}h^{\sharp}H(t)h_{\sharp}\big{)}v(t,y)+\frac{i}{2}\frac{\partial_{t}\big{(}|J(t,y)|\big{)}}{|J(t,y)|}v(t,y)+i\big{\langle}\partial_{t}h(t,y)|(h^{\sharp}\nabla_{x}h_{\sharp})v(t,y)\big{\rangle}
=(hH(t)h)v(t,y)+i2t(|J(t,y)|)|J(t,y)|v(t,y)+h[i(hth)(t,x)|x]hv(t,y).\displaystyle=\big{(}h^{\sharp}H(t)h_{\sharp}\big{)}v(t,y)+\frac{i}{2}\frac{\partial_{t}\big{(}|J(t,y)|\big{)}}{|J(t,y)|}v(t,y)+h^{\sharp}\Big{[}i\big{\langle}(h_{*}\partial_{t}h)(t,x)|\nabla_{x}\cdot\big{\rangle}\Big{]}h_{\sharp}v(t,y). (3.5)

The first operator (hH(t)h)(h^{\sharp}H(t)h_{\sharp}) is a simple transport of the original operator H(t)H(t) and it is clearly self-adjoint. Both other terms from the last relation come instead from the time derivative of hh^{\sharp}. Since hh^{\sharp} is unitary, we expect that, their sum is also a self-adjoint operator. Notice that the last term may be expressed explicitly by Proposition 2.1 to obtain Equation (1.7). However, we would like to keep the conjugated form to obtain Equation (1.8). Due to Proposition A.2 in the appendix and (2.1), we have

t(|J(t,y)|)|J(t,y)|\displaystyle\frac{\partial_{t}\big{(}|J(t,y)|\big{)}}{|J(t,y)|} =Tr(J(t,y)1tJ(t,y))=Tr(tJ(t,y)J(t,y)1)\displaystyle=\operatorname{Tr}\Big{(}J(t,y)^{-1}\cdot\partial_{t}J(t,y)\Big{)}=\operatorname{Tr}\Big{(}\partial_{t}J(t,y)\cdot J(t,y)^{-1}\Big{)}
=hTr((tDyh)h1(Dyh)1h1)\displaystyle=h^{*}\operatorname{Tr}\Big{(}\big{(}\partial_{t}D_{y}h\big{)}\circ h^{-1}\cdot(D_{y}h)^{-1}\circ h^{-1}\Big{)}
=hTr((Dyth)h1Dx(h1))=hTr(Dx((th)h1))\displaystyle=h^{*}\operatorname{Tr}\Big{(}\big{(}D_{y}\partial_{t}h\big{)}\circ h^{-1}\cdot D_{x}(h^{-1})\Big{)}=h^{*}\operatorname{Tr}\Big{(}D_{x}\big{(}(\partial_{t}h)\circ h^{-1}\big{)}\Big{)}
=hdivx(h(th)).\displaystyle=h^{*}\operatorname{div}_{x}\big{(}h_{*}(\partial_{t}h)\big{)}\leavevmode\nobreak\ .

Since hh^{\sharp} differs from hh^{*} by a multiplication, the conjugacy of a multiplicative operation by either hh^{*} or hh^{\sharp} gives the same result. We obtain that

i2t(|J(t,y)|)|J(t,y)|v(t,y)\displaystyle\frac{i}{2}\frac{\partial_{t}\big{(}|J(t,y)|\big{)}}{|J(t,y)|}v(t,y) =i2[hdivx(h(th))]v=i2h[divx(h(th))hv]\displaystyle=\frac{i}{2}\Big{[}h^{*}\operatorname{div}_{x}\big{(}h_{*}(\partial_{t}h)\big{)}\Big{]}v=\frac{i}{2}h^{*}\Big{[}\operatorname{div}_{x}\big{(}h_{*}(\partial_{t}h)\big{)}h_{*}v\Big{]}
=i2h[divx(h(th))hv].\displaystyle=\frac{i}{2}h^{\sharp}\Big{[}\operatorname{div}_{x}\big{(}h_{*}(\partial_{t}h)\big{)}h_{\sharp}v\Big{]}. (3.6)

We combine the result of (3.6) and half of the last term of (3.5) by using the chain rule

i2divx(h(th))u+i(hth)|xu=i2divx((h(th))u)+i2(hth)|xu.\frac{i}{2}\operatorname{div}_{x}\big{(}h_{*}(\partial_{t}h)\big{)}u+i\big{\langle}(h_{*}\partial_{t}h)|\nabla_{x}u\big{\rangle}=\frac{i}{2}\operatorname{div}_{x}\Big{(}\big{(}h_{*}(\partial_{t}h)\big{)}u\Big{)}+\frac{i}{2}\big{\langle}(h_{*}\partial_{t}h)|\nabla_{x}u\big{\rangle}.

We finally obtain

itv(t,y)=h[H(t)idivx(Ah(t,x))iAh(t,x)|x]hv(t,y)i\partial_{t}v(t,y)\leavevmode\nobreak\ =\leavevmode\nobreak\ h^{\sharp}\Big{[}H(t)\cdot\leavevmode\nobreak\ -\leavevmode\nobreak\ i\operatorname{div}_{x}\big{(}A_{h}(t,x)\cdot\big{)}\leavevmode\nobreak\ -\leavevmode\nobreak\ i\big{\langle}A_{h}(t,x)\big{|}\nabla_{x}\cdot\big{\rangle}\Big{]}h_{\sharp}v(t,y) (3.7)

where Ah(t,x)=12(hth)(t,x)A_{h}(t,x)=-\frac{1}{2}(h_{*}\partial_{t}h)(t,x) is a magnetic field generated by the change of referential. The whole operator of (3.7) can be seen as a modification of the magnetic and electric terms of H(t)H(t). Indeed, using (3.1), the equation (3.7) becomes

{itv(t,y)=(hH~h(t)h)v(t,y),tI,yΩ0v|Ω00\left\{\begin{array}[]{ll}i\partial_{t}v(t,y)\leavevmode\nobreak\ =\leavevmode\nobreak\ \big{(}h^{\sharp}\tilde{H}_{h}(t)h_{\sharp}\big{)}v(t,y),\nobreak\leavevmode\nobreak\leavevmode\nobreak\leavevmode&t\in I\leavevmode\nobreak\ ,\leavevmode\nobreak\ \leavevmode\nobreak\ y\in\Omega_{0}\\ v_{|\partial\Omega_{0}}\equiv 0\end{array}\right. (3.8)

where, using the same standard notation of (3.2),

H~h(t)=[D(t,x)x+iA~h(t,x)]2+V~h(t,x)\tilde{H}_{h}(t)=-\Big{[}D(t,x)\nabla_{x}+i\tilde{A}_{h}(t,x)\Big{]}^{2}+\tilde{V}_{h}(t,x) (3.9)

with

A~h(t,x)=A(t,x)+(D(t,x)1)tAh(t,x),Ah=12h(th),\tilde{A}_{h}(t,x)=A(t,x)+\big{(}D(t,x)^{-1}\big{)}^{t}\cdot A_{h}(t,x)\leavevmode\nobreak\ \leavevmode\nobreak\ ,\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ A_{h}=-\frac{1}{2}h_{*}(\partial_{t}h),
V~h(t,x)=V(t,x)|(D(t,x)1)tAh(t,x)|22(D(t,x)1)tAh(t,x)|A(t,x).\tilde{V}_{h}(t,x)=V(t,x)-\big{|}\big{(}D(t,x)^{-1}\big{)}^{t}\cdot A_{h}(t,x)\big{|}^{2}-2\big{\langle}\big{(}D(t,x)^{-1}\big{)}^{t}\cdot A_{h}(t,x)\big{|}A(t,x)\big{\rangle}\leavevmode\nobreak\ .

Notice that, in the simplest case of H(t)H(t) being the Dirichlet Laplacian operator, we obtain Equation (1.8) discussed in the introduction. Similar computations are provided in [22, 23].

It is now possible to prove the main result of this section which generalizes Theorem 1.1 of the introduction.

Theorem 3.1.

Assume that Ω0N\Omega_{0}\subset\mathbb{R}^{N} is a bounded domain (possibly irregular). Let II\subset\mathbb{R} be an interval of times. Assume that h(t):Ω¯0Ω¯(t):=h(t,Ω¯0)h(t):\overline{\Omega}_{0}\longrightarrow\overline{\Omega}(t):=h(t,\overline{\Omega}_{0}) is a family of diffeomorphisms which is of class 𝒞2\mathcal{C}^{2} with respect to the time and the space variable. Let H(t)H(t) be an operator of the form (3.1) with domain H2(Ω(t),)H01(Ω(t),)H^{2}(\Omega(t),\mathbb{C})\cap H^{1}_{0}(\Omega(t),\mathbb{C}).

Then, Equation (3.8) generates a unitary flow U~(t,s)\tilde{U}(t,s) on L2(Ω0)L^{2}(\Omega_{0}) and we may define weak solutions of the Schrödinger equation (3.3) on the moving domain Ω(t)\Omega(t) by transporting this flow via hh_{\sharp}, that is setting U(t,s)=hU~(t,s)hU(t,s)=h_{\sharp}\tilde{U}(t,s)h^{\sharp}.

Assume in addition that the diffeomorphisms hh are of class 𝒞3\mathcal{C}^{3} with respect to the time and the space variable. Then, for any u0H2(Ω(t0))H01(Ω(t0))u_{0}\in H^{2}(\Omega(t_{0}))\cap H^{1}_{0}(\Omega(t_{0})), the above flow defines a solution u(t)=U(t,t0)u0u(t)=U(t,t_{0})u_{0} in 𝒞0(I,H2(Ω(t))H01(Ω(t)))𝒞1(I,L2(Ω(t)))\mathcal{C}^{0}(I,H^{2}(\Omega(t))\cap H^{1}_{0}(\Omega(t)))\cap\mathcal{C}^{1}(I,L^{2}(\Omega(t))) solving (3.3) in the L2L^{2}-sense.

Proof: Assume that II is compact, otherwise it is sufficient to cover II with compact intervals and to glue the unitary flows defined on each one of them. We simply apply Theorem A.1 in appendix. We notice Ah=12h(th)A_{h}=-\frac{1}{2}h_{*}(\partial_{t}h) is a bounded term in 𝒞t1(I,𝒞x2(Ω¯(t),))\mathcal{C}^{1}_{t}(I,\mathcal{C}^{2}_{x}(\overline{\Omega}(t),\mathbb{R})). First, the assumptions on H(t)H(t) and the regularity of the diffeomorphisms h(t,)h(t,\cdot) imply that H~h(t)\tilde{H}_{h}(t) defined by (3.9) is a well defined self-adjoint operator on L2(Ω(t))L^{2}(\Omega(t)) with domain H2(Ω(t))H01(Ω(t))H^{2}(\Omega(t))\cap H^{1}_{0}(\Omega(t)). Second, it is of class 𝒞1\mathcal{C}^{1} with respect to the time and, for every uH2H1(Ω(t))u\in H^{2}\cap H^{1}(\Omega(t)),

H~h(t)u|uL2(Ω(t))\displaystyle\big{\langle}\tilde{H}_{h}(t)u|u\big{\rangle}_{L^{2}(\Omega(t))} =Ω(t)|(D(t,x)x+iA~h)u(x)|2+V~h(t,x)|u(x)|2dx\displaystyle=\int_{\Omega(t)}\big{|}\big{(}D(t,x)\nabla_{x}+i\tilde{A}_{h}\big{)}u(x)\big{|}^{2}+\tilde{V}_{h}(t,x)|u(x)|^{2}\,{\text{\rm d}}x
γuH1(Ω(t))2κuL2(Ω(t))2\displaystyle\geq\gamma\|u\|_{H^{1}(\Omega(t))}^{2}-\kappa\|u\|_{L^{2}(\Omega(t))}^{2}

for some γ>0\gamma>0 and κ\kappa\in\mathbb{R}. Let (t)=h(t)H~h(t)h(t)\mathcal{H}(t)=h^{\sharp}(t)\circ\tilde{H}_{h}(t)\circ h_{\sharp}(t). Since hh^{\sharp} is an isometry from L2(Ω(t))L^{2}(\Omega(t)) onto L2(Ω0)L^{2}(\Omega_{0}) continuously mapping H1(Ω(t))H^{1}(\Omega(t)) into H1(Ω0)H^{1}(\Omega_{0}), the above properties of H~h(t)\tilde{H}_{h}(t) are also valid for (t)\mathcal{H}(t) because for every vH2(Ω0)H01(Ω0)v\in H^{2}(\Omega_{0})\cap H^{1}_{0}(\Omega_{0})

(t)v|vL2(Ω0)=hH~h(t)hv|vL2(Ω0)=H~h(t)hv|hvL2(Ω(t)).\big{\langle}\mathcal{H}(t)v|v\big{\rangle}_{L^{2}(\Omega_{0})}\leavevmode\nobreak\ =\leavevmode\nobreak\ \big{\langle}h^{\sharp}\tilde{H}_{h}(t)h_{\sharp}v|v\big{\rangle}_{L^{2}(\Omega_{0})}\leavevmode\nobreak\ =\leavevmode\nobreak\ \big{\langle}\tilde{H}_{h}(t)h_{\sharp}v|h_{\sharp}v\big{\rangle}_{L^{2}(\Omega(t))}\leavevmode\nobreak\ .

The only problem concerns the regularities. Indeed, the domain of (t)\mathcal{H}(t) is

D((t))={vL2(Ω0),h(t)vH2(Ω(t))H01(Ω(t))}D(\mathcal{H}(t))\leavevmode\nobreak\ =\leavevmode\nobreak\ \{v\in L^{2}(\Omega_{0})\leavevmode\nobreak\ ,\leavevmode\nobreak\ h_{\sharp}(t)v\in H^{2}(\Omega(t))\cap H^{1}_{0}(\Omega(t))\}

and if hh is only 𝒞2\mathcal{C}^{2} with respect to the space, then D((t))D(\mathcal{H}(t)) is not necessarily H2(Ω0)H01(Ω0)H^{2}(\Omega_{0})\cap H^{1}_{0}(\Omega_{0}) due to the presence of the Jacobian of hh in the definition (1.6) of hh_{\sharp}. However, if hh is of class 𝒞2\mathcal{C}^{2}, then hh_{\sharp} transports the 𝒞1\mathcal{C}^{1}-regularity in space as well as the boundary condition and

D((t)1/2)={vL2(Ω0),h(t)vH01(Ω(t))}=H01(Ω0)D(\mathcal{H}(t)^{1/2})\leavevmode\nobreak\ =\leavevmode\nobreak\ \{v\in L^{2}(\Omega_{0})\leavevmode\nobreak\ ,\leavevmode\nobreak\ h_{\sharp}(t)v\in H^{1}_{0}(\Omega(t))\}\leavevmode\nobreak\ =\leavevmode\nobreak\ H^{1}_{0}(\Omega_{0})

does not depend on the time. We can apply Theorem A.1 in the appendix and obtain the unitary flow U~(t,s)\tilde{U}(t,s). The flow U(t,s)=hU~(t,s)hU(t,s)=h_{\sharp}\tilde{U}(t,s)h^{\sharp} on L2(Ω(t))L^{2}(\Omega(t)) is then well defined but corresponds to solutions of (3.3) only in a formal way.

Assume finally that the diffeomorphisms hh are of class 𝒞3\mathcal{C}^{3} with respect to the time and the space variable. Then, we have no more problems with the domains and

D((t))={vL2(Ω0),h(t)vH2(Ω(t))H01(Ω(t))}=H2(Ω0)H01(Ω0)D(\mathcal{H}(t))\leavevmode\nobreak\ =\leavevmode\nobreak\ \{v\in L^{2}(\Omega_{0})\leavevmode\nobreak\ ,\leavevmode\nobreak\ h_{\sharp}(t)v\in H^{2}(\Omega(t))\cap H^{1}_{0}(\Omega(t))\}\leavevmode\nobreak\ =\leavevmode\nobreak\ H^{2}(\Omega_{0})\cap H^{1}_{0}(\Omega_{0})

which does not depend on the time. Since H(t)H(t) is now of class 𝒞2\mathcal{C}^{2} with respect to the time, we may apply the second part of Theorem A.1 and obtain strong solutions of the equation on Ω0\Omega_{0}, which are transported to strong solutions of the equation on Ω(t)\Omega(t). \square

3.2 Proof of Theorem 1.2: the magnetic Neumann case

In the previous section, the well-posedness of the dynamics of (3.8) is ensured by studying the self-adjoint operator H~h(t)\tilde{H}_{h}(t) defined in (3.9) and with domain H2(Ω(t))H01(Ω(t))H^{2}(\Omega(t))\cap H^{1}_{0}(\Omega(t)). This operator corresponds to the following quadratic form in H01(Ω(t))H^{1}_{0}(\Omega(t))

q(ψ):=Ω(t)|D(t,x)ψ(x)+iA~h(t,x)ψ(x)|2dx+Ω(t)V~h(t,x)|ψ(x)|2dx,ψH01(Ω(t))q(\psi):=\int_{\Omega(t)}\Big{|}D(t,x)\nabla\psi(x)+i\tilde{A}_{h}(t,x)\psi(x)\Big{|}^{2}\,{\text{\rm d}}x+\int_{\Omega(t)}\tilde{V}_{h}(t,x)|\psi(x)|^{2}\,{\text{\rm d}}x,\ \ \ \ \ \forall\psi\in H^{1}_{0}(\Omega(t))

which is the Friedrichs extension of the quadratic form qq defined on 𝒞0(Ω(t))\mathcal{C}^{\infty}_{0}(\Omega(t)). Now, it is natural to consider the Friedrichs extension of qq defined in H1(Ω(t))H^{1}(\Omega(t)). This corresponds to the Neumann realization of the magnetic Laplacian H~h(t)\tilde{H}_{h}(t) defined by (3.9) and with domain

D(H~h(t))={uH2(Ω(t),):ν|Dxu+iA~hu(t,x)=0,xΩ(t)}D\big{(}\tilde{H}_{h}(t)\big{)}=\Big{\{}u\in H^{2}(\Omega(t),\mathbb{C})\ :\ \big{\langle}\,\nu\,\big{|}\,D\nabla_{x}u+i\tilde{A}_{h}u\big{\rangle}(t,x)=0,\ \ \forall x\in\partial\Omega(t)\Big{\}} (3.10)

(see [20] for example). Such as the well-posedness of the Schrödinger equation on moving domains can be achieve when it is endowed with Dirichlet boundary conditions, the same result can be addressed in this new framework. Indeed, the operators H~h(t)\tilde{H}_{h}(t) endowed with the domain (3.10) are still self-adjoint and bounded from below. The arguments developed in the previous section lead to the well-posedness in Ω0\Omega_{0} of the equation

{itv(t,y)=(hH~h(t)h)v(t,y),tI,yΩ0(hν|Dx(hv)+iA~h(hv))(t,y)=0,tI,xΩ0\left\{\begin{array}[]{ll}i\partial_{t}v(t,y)\leavevmode\nobreak\ =\leavevmode\nobreak\ \big{(}h^{\sharp}\tilde{H}_{h}(t)h_{\sharp}\big{)}v(t,y),\nobreak\leavevmode\nobreak\leavevmode\nobreak\leavevmode&t\in I\leavevmode\nobreak\ ,\leavevmode\nobreak\ \leavevmode\nobreak\ y\in\Omega_{0}\\ \Big{(}h^{\sharp}\big{\langle}\,\nu\,\big{|}D\nabla_{x}(h_{\sharp}v)+i\tilde{A}_{h}(h_{\sharp}v)\big{\rangle}\Big{)}(t,y)=0,\nobreak\leavevmode\nobreak\leavevmode\nobreak\leavevmode\nobreak\leavevmode\nobreak\leavevmode&t\in I\leavevmode\nobreak\ ,\leavevmode\nobreak\ x\in\partial\Omega_{0}\end{array}\right. (3.11)

Going back to the original moving domain Ω(t)\Omega(t), (3.11) becomes

{itu(t,x)=H(t)u(t,x),tI,xΩ(t)ν|Dxu+iA~hu(t,x)=0,tI,xΩ(t).\begin{split}\begin{cases}i\partial_{t}u(t,x)=H(t)u(t,x),\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ &t\in I\leavevmode\nobreak\ ,\leavevmode\nobreak\ x\in\Omega(t)\\ \big{\langle}\,\nu\,\big{|}D\nabla_{x}u+i\tilde{A}_{h}u\big{\rangle}(t,x)=0,\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ &t\in I\leavevmode\nobreak\ ,\leavevmode\nobreak\ x\in\partial\Omega(t).\end{cases}\end{split} (3.12)

It is noteworthy that in this case, the modified magnetic potential A~h\tilde{A}_{h} appears in the original equation. In particular, notice that the boundary condition of (3.12) is not the natural one associated with H(t)H(t) for tt fixed.

We resume in the following theorem the well-posedness of the dynamics when the Neumann magnetic boundary conditions are considered. The result is a generalization of Theorem 1.2 in the introduction.

Theorem 3.2.

Assume that Ω0N\Omega_{0}\subset\mathbb{R}^{N} is a bounded domain (possibly irregular). Let II\subset\mathbb{R} be an interval of times. Assume that h(t):Ω¯0Ω¯(t):=h(t,Ω¯0)h(t):\overline{\Omega}_{0}\longrightarrow\overline{\Omega}(t):=h(t,\overline{\Omega}_{0}) is a family of diffeomorphisms which is of class 𝒞2\mathcal{C}^{2} with respect to the time and the space variable. Let H(t)H(t) be a of the form (3.1) and with domain (3.10).

Then, Equation (3.11) generates a unitary flow U~(t,s)\tilde{U}(t,s) on L2(Ω0)L^{2}(\Omega_{0}) and we may define weak solutions of the Schrödinger equation (3.12) on the moving domain Ω(t)\Omega(t) by transporting this flow via hh_{\sharp}, that is setting U(t,s)=hU~(t,s)hU(t,s)=h_{\sharp}\tilde{U}(t,s)h^{\sharp}.

Assume in addition that the diffeomorphisms hh are of class 𝒞3\mathcal{C}^{3} with respect to the time and the space variable. Then, for any u0H2(Ω(t0))H01(Ω(t0))u_{0}\in H^{2}(\Omega(t_{0}))\cap H^{1}_{0}(\Omega(t_{0})), the above flow defines a solution u(t)=U(t,t0)u0u(t)=U(t,t_{0})u_{0} in 𝒞0(I,H2(Ω(t))H01(Ω(t)))𝒞1(I,L2(Ω(t)))\mathcal{C}^{0}(I,H^{2}(\Omega(t))\cap H^{1}_{0}(\Omega(t)))\cap\mathcal{C}^{1}(I,L^{2}(\Omega(t))) solving (3.12) in the L2L^{2}-sense.

Proof: Mutatis mutandis, the proof is the same as the one of Theorem 3.1. \square

Remark 3.3.

In (3.11), the boundary conditions satisfied by vv are not explicitly stated in terms of vv but rather in terms of u=hvu=h_{\sharp}v. Even if it not necessary for proving Theorem 3.2, it could be interesting to compute them completely. The second statement of Proposition 2.2 implies

|J|ν0|J1h(Dx(hv))+i(J1h(A~h(hv))=0 on Ω0.|J|\big{\langle}\,\nu_{0}\,\big{|}\,J^{-1}\cdot h^{*}(D\cdot\nabla_{x}(h_{\sharp}v))+i\big{(}J^{-1}\cdot h^{*}(\tilde{A}_{h}(h_{\sharp}v))\big{\rangle}=0\leavevmode\nobreak\ \leavevmode\nobreak\ \text{ on }\partial\Omega_{0}\leavevmode\nobreak\ .

Now, hv=h1|J|vh_{\sharp}v=h_{*}\frac{1}{\sqrt{|J|}}v. Thanks to Proposition 2.1, the fact that |J||J| never vanishes yields

(J1)tν0|(hD)(J1)t|J|y(vJ)+i(hA~h)v=0 on Ω0.\Big{\langle}(J^{-1})^{t}\nu_{0}\Big{|}(h^{*}D)\cdot(J^{-1})^{t}\cdot\sqrt{|J|}\nabla_{y}\big{(}\frac{v}{\sqrt{J}}\big{)}+i(h^{*}\tilde{A}_{h})v\Big{\rangle}=0\leavevmode\nobreak\ \leavevmode\nobreak\ \text{ on }\partial\Omega_{0}\leavevmode\nobreak\ .

When A~h\tilde{A}_{h} corresponds to Ah=12h(th)A_{h}=-\frac{1}{2}h_{*}(\partial_{t}h) and D(t,x)=IdD(t,x)=Id, we obtain the boundary condition presented in (1.11).

4 Moser’s trick

The aim of this section is to prove the following result.

Theorem 4.1.

Let k1k\geq 1 and r0r\geq 0 with krk\geq r and let α(0,1)\alpha\in(0,1). Let Ω0\Omega_{0} be a bounded connected open 𝒞k+2,α\mathcal{C}^{k+2,\alpha} domain of N\mathbb{R}^{N}. Let II\subset\mathbb{R} be an interval and let f𝒞r(I,𝒞k1,α(Ω¯0,+))f\in\mathcal{C}^{r}(I,\mathcal{C}^{k-1,\alpha}(\overline{\Omega}_{0},\mathbb{R}^{*}_{+})) be such that Ω0f(t,y)dy=meas(Ω0)\int_{\Omega_{0}}f(t,y)\,{\text{\rm d}}y={\text{\rm meas}}(\Omega_{0}) for all tIt\in I. Then, there exists a family (φ(t))tI(\varphi(t))_{t\in I} of diffeomorphisms of Ω¯0\overline{\Omega}_{0} of class 𝒞r(I,𝒞k,α(Ω¯0,Ω¯0))\mathcal{C}^{r}(I,\mathcal{C}^{k,\alpha}(\overline{\Omega}_{0},\overline{\Omega}_{0})) satisfying

{det(Dyφ(t,y))=f(t,y),yΩ0φ(t,y)=y,yΩ0.\left\{\begin{array}[]{ll}\det(D_{y}\varphi(t,y))=f(t,y),\nobreak\leavevmode&y\in\Omega_{0}\\ \varphi(t,y)=y,&y\in\partial\Omega_{0}.\end{array}\right. (4.1)

As explained in Section 1, an interesting change of variables for a PDE with moving domains would be a family of diffeomorphisms having Jacobian with constant determinant. This is the goal of Theorem 1.3, which is a direct consequence of Theorem 4.1.

Proof of Theorem 1.3: Let Theorem 4.1 be valid. We set h~(t)=h(t)φ1(t)\tilde{h}(t)=h(t)\circ\varphi^{-1}(t) and we compute

det(Dyh~(t,y))\displaystyle\det\big{(}D_{y}\tilde{h}(t,y)\big{)} =det(Dyh(t,φ1(y)))det(Dy(φ1)(t,y))\displaystyle=\det\big{(}D_{y}h(t,\varphi^{-1}(y))\big{)}\det\big{(}D_{y}(\varphi^{-1})(t,y)\big{)}
=det(Dyh(t,φ1(y)))det((Dyφ)1(t,φ1(y))\displaystyle=\det\big{(}D_{y}h(t,\varphi^{-1}(y))\big{)}\det\big{(}(D_{y}\varphi)^{-1}(t,\varphi^{-1}(y)\big{)}
=(det(Dyh)det(Dyφ))(t,φ1(y)).\displaystyle=\left(\frac{\det(D_{y}h)}{\det(D_{y}\varphi)}\right)(t,\varphi^{-1}(y)).

Since we want to obtain a spatially constant right-hand side, the dependence in φ1(y)\varphi^{-1}(y) is not important. Thus,

yΩ0,det(Dyh~(t,y))=meas(Ω(t))meas(Ω0)\forall y\in\Omega_{0}\leavevmode\nobreak\ ,\leavevmode\nobreak\ \leavevmode\nobreak\ \det(D_{y}\tilde{h}(t,y))=\frac{{\text{\rm meas}}(\Omega(t))}{{\text{\rm meas}}(\Omega_{0})}

if and only if

yΩ0,det(Dyφ(t,y))=meas(Ω0)meas(Ω(t))det(Dyh(t,y)).\forall y\in\Omega_{0}\leavevmode\nobreak\ ,\leavevmode\nobreak\ \leavevmode\nobreak\ \det(D_{y}\varphi(t,y))=\frac{{\text{\rm meas}}(\Omega_{0})}{{\text{\rm meas}}(\Omega(t))}\det(D_{y}h(t,y))\leavevmode\nobreak\ .

To obtain such a diffeomorphism φ\varphi, it remains to apply Theorem 4.1 with f(t,y)f(t,y) being the above right-hand side. \square

4.1 The classical results

Theorem 4.1 is an example of a family of results aiming to find diffeomorphisms having prescribed determinant of the Jacobian. Such outcomes are often referred as “Moser’s trick” because they were originated by the famous work of Moser [42]. A lot of variants can be found in the literature, depending on the needs of the reader. We refer to [15] for a review on the subject. The following result comes from [14, 16], see also [15].

Theorem 4.2.

Dacorogna-Moser (1990)
Let k1k\geq 1 and α(0,1)\alpha\in(0,1). Let Ω0\Omega_{0} be a bounded connected open 𝒞k+2,α\mathcal{C}^{k+2,\alpha} domain. Then, the following statements are equivalent.

  1. (i)

    The function f𝒞k1,α(Ω¯0,+)f\in\mathcal{C}^{k-1,\alpha}(\overline{\Omega}_{0},\mathbb{R}^{*}_{+}) satisfies Ω0f=meas(Ω0)\int_{\Omega_{0}}f={\text{\rm meas}}(\Omega_{0}).

  2. (ii)

    There exists φDiffk,α(Ω¯0,Ω¯0)\varphi\in{\text{\rm Diff}}^{k,\alpha}(\overline{\Omega}_{0},\overline{\Omega}_{0}) satisfying

    {det(Dyφ(y))=f(y),yΩ0φ(y)=y,yΩ0.\left\{\begin{array}[]{ll}\det(D_{y}\varphi(y))=f(y),\nobreak\leavevmode&y\in\Omega_{0}\\ \varphi(y)=y,&y\in\partial\Omega_{0}.\end{array}\right.

Furthermore, if c>0c>0 is such that max(f,1/f)c\max(\|f\|_{\infty},\|1/f\|_{\infty})\leq c then there exists a constant C=C(c,α,k,Ω0)C=C(c,\alpha,k,\Omega_{0}) such that

φid𝒞k,αCf1𝒞k1,α.\|\varphi-{\text{\rm id}}\|_{\mathcal{C}^{k,\alpha}}\leavevmode\nobreak\ \leq\leavevmode\nobreak\ C\|f-1\|_{\mathcal{C}^{k-1,\alpha}}\leavevmode\nobreak\ .

A complete proof of theorem 4.2 can be found in [15]. There, a discussion concerning the optimality of the regularity of the diffeomorphisms is also provided. In particular, notice that the natural gain of regularity from f𝒞k1,αf\in\mathcal{C}^{k-1,\alpha} to φ𝒞k,α\varphi\in\mathcal{C}^{k,\alpha} was not present in the first work of Moser [42] and cannot be obtained through the original method. Cases with other space regularity are studied in [48, 49].

For tt fixed, Theorem 4.1 corresponds to Theorem 4.2. Our main goal is to extend the result to a time-dependent measure f(t,y)f(t,y). In the original proof of [42], Moser uses a flow method and constructs φ\varphi by a smooth deformation starting at f(t=0)=idf(t=0)={\text{\rm id}} and reaching f(t=1)=ff(t=1)=f. In this proof, the smooth deformation is a linear interpolation f(t)=(1t)id+tff(t)=(1-t){\text{\rm id}}+tf and one of the main steps consists in solving an ODE with a non-linearity as smooth as tf(t)\partial_{t}f(t). The linear interpolation is of course harmless in such a context. But when we consider another type of time-dependence, we need to have enough smoothness to solve the ODE. Typically, at least 𝒞1\mathcal{C}^{1}-smoothness in time is required to use the arguments from the original paper [42] of Moser. Thus, the original proof of [42] does not provide an optimal regularity, neither in space or time. In particular, in Theorem 1.3, this type of proof provides a diffeomorphism h~\tilde{h} with one space regularity less than hh. We refer to [1, 4, 25] for other time-dependent versions of Theorem 4.2.

Our aim is to obtain a better regularity in space and time as well as to provide a complete proof of a time-dependent version of Moser’s trick. To this end, we follow a method coming from [15, 16], which is already known for obtaining the optimal space regularity. This method uses a fixed point argument, which can easily be parametrized with respect to time. Nevertheless, the fixed point argument only provides a local construction which is difficult to extend by gluing several similar construction (equations as (4.1) have an infinite number of solutions and the lack of uniqueness makes difficult to glue smoothly the different curves). That is the reason why we follow a similar strategy to the ones adopted in [15, 16]. First, at the cost of losing some regularity, we prove the global result with the flow method. Second, we exploit a fixed point argument in order to ensure the result with respect to the optimal regularity.

4.2 The flow method

By using the flow method of the original work of Moser [42], we obtain the following version of Theorem 4.1, where the statements on the regularities are weakened.

Proposition 4.3.

Let k2k\geq 2 and r1r\geq 1 with krk\geq r. Let Ω0\Omega_{0} be a bounded connected open 𝒞k,α\mathcal{C}^{k,\alpha} domain of N\mathbb{R}^{N} for some α(0,1)\alpha\in(0,1). Let II\subset\mathbb{R} be an interval of times and let f𝒞r(I,𝒞k1(Ω¯0,+))f\in\mathcal{C}^{r}(I,\mathcal{C}^{k-1}(\overline{\Omega}_{0},\mathbb{R}^{*}_{+})) be such that Ω0f(t,y)dy=meas(Ω0)\int_{\Omega_{0}}f(t,y)\,{\text{\rm d}}y={\text{\rm meas}}(\Omega_{0}) for all tIt\in I. Then, there exists a family (φ(t))tI(\varphi(t))_{t\in I} of diffeomorphisms of Ω¯0\overline{\Omega}_{0} of class 𝒞r(I,𝒞k1(Ω¯0,Ω¯0))\mathcal{C}^{r}(I,\mathcal{C}^{k-1}(\overline{\Omega}_{0},\overline{\Omega}_{0})) satisfying (4.1).

Moreover, there exist continuous functions MC(M)M\mapsto C(M) and Mλ(M)M\mapsto\lambda(M) such that, if

tI,f(t,)𝒞k1M and minyΩ0f(t,y)1M,\forall t\in I\leavevmode\nobreak\ ,\leavevmode\nobreak\ \leavevmode\nobreak\ \|f(t,\cdot)\|_{\mathcal{C}^{k-1}}\leavevmode\nobreak\ \leq\leavevmode\nobreak\ M\text{ and }\min_{y\in\Omega_{0}}f(t,y)\geq\frac{1}{M},

then

tI,φ(t,)𝒞k1+φ1(t,)𝒞k1C(M)eλ(M)|t|.\forall t\in I\leavevmode\nobreak\ ,\leavevmode\nobreak\ \leavevmode\nobreak\ \|\varphi(t,\cdot)\|_{\mathcal{C}^{k-1}}\,+\,\|\varphi^{-1}(t,\cdot)\|_{\mathcal{C}^{k-1}}\,\leq\,C(M)e^{\lambda(M)|t|}.

Proof: Let t0It_{0}\in I. Due to theorem 4.2, there is φ(t0)\varphi(t_{0}) with the required space regularity satisfying (4.1) at t=t0t=t_{0}. Setting φ(t)=φ~(t)φ(t0)\varphi(t)=\tilde{\varphi}(t)\circ\varphi(t_{0}), we replace (4.1) by the condition

det(Dyφ~(t,y))=f(t,φ1(t0,y))f(t0,φ1(t0,y)).\det(D_{y}\tilde{\varphi}(t,y))=\frac{f(t,\varphi^{-1}(t_{0},y))}{f(t_{0},\varphi^{-1}(t_{0},y))}\leavevmode\nobreak\ .

Thus, we may assume without loss of generality that f(t0,y)1f(t_{0},y)\equiv 1.

Let L1L^{-1} be the right-inverse of the divergence introduced in Appendix A.3. Notice that tf(t,)\partial_{t}f(t,\cdot) is of class 𝒞k1\mathcal{C}^{k-1} and hence of class 𝒞k2,α\mathcal{C}^{k-2,\alpha}. Since Ω0tf=tΩ0f=0\int_{\Omega_{0}}\partial_{t}f=\partial_{t}\int_{\Omega_{0}}f=0 we get that f(t)Ymk2,αf(t)\in Y^{k-2,\alpha}_{m} for all tIt\in I and we can define

U(t,)=1f(t,)L1(tf(t,)).U(t,\cdot)\leavevmode\nobreak\ =\leavevmode\nobreak\ -\frac{1}{f(t,\cdot)}L^{-1}(\partial_{t}f(t,\cdot))\leavevmode\nobreak\ . (4.2)

We get that UU is well defined and it is of class 𝒞r1\mathcal{C}^{r-1} in time and 𝒞k1\mathcal{C}^{k-1} in space. For xΩ¯0x\in\overline{\Omega}_{0}, we define tψ(t,x)t\mapsto\psi(t,x) as the flow corresponding to the ODE

ψ(t0,x)=x and tψ(t,x)=U(t,ψ(t,x))tI.\psi(t_{0},x)=x\leavevmode\nobreak\ \leavevmode\nobreak\ \text{ and }\leavevmode\nobreak\ \leavevmode\nobreak\ \partial_{t}\psi(t,x)=U(t,\psi(t,x))\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ t\in I\leavevmode\nobreak\ . (4.3)

Notice that ψ\psi is locally well defined because (t,y)U(t,y)(t,y)\mapsto U(t,y) is at least lipschitzian in space and continuous in time. Moreover, the trajectories tIψ(t,x)Ω¯0t\in I\mapsto\psi(t,x)\in\overline{\Omega}_{0} are globally defined because U(t,y)=0U(t,y)=0 on the boundary of Ω0\Omega_{0}, providing a barrier of equilibrium points. The classical regularity results for ODEs show that ψ\psi is of class 𝒞r\mathcal{C}^{r} with respect to time and 𝒞k1\mathcal{C}^{k-1} in space (see Proposition A.4 in the appendix). Moreover, by reversing the time, the flow of a classical ODE as (4.3) is invertible and ψ(t,)\psi(t,\cdot) is a diffeomorphism for all tIt\in I. We set φ(t,)=ψ1(t,)\varphi(t,\cdot)=\psi^{-1}(t,\cdot).

Using Proposition A.2, we compute for tIt\in I

t[det(Dyψ(t,y))f(t,ψ(t,y))]\displaystyle\partial_{t}\big{[}\det(D_{y}\psi(t,y))\,f(t,\psi(t,y))\big{]} =(tdet(Dyψ(t,y)))fψ(t,y)+det(Dyψ(t,y))(tf)ψ(t,y)\displaystyle=(\partial_{t}\det(D_{y}\psi(t,y)))\,f\circ\psi(t,y)+\det(D_{y}\psi(t,y))\,(\partial_{t}f)\circ\psi(t,y)
+det(Dyψ(t,y))((Dyf)ψ)(tψ)(t,y)\displaystyle\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ +\det(D_{y}\psi(t,y))((D_{y}f)\circ\psi)\cdot(\partial_{t}\psi)(t,y)
=det(Dyψ)[Tr((Dyψ)1t(Dyψ))fψ\displaystyle=\det(D_{y}\psi)\Big{[}\operatorname{Tr}\big{(}(D_{y}\psi)^{-1}\partial_{t}(D_{y}\psi)\big{)}f\circ\psi
+(tf)ψ+((Dyf)ψ)(tψ)](t,y).\displaystyle\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ +(\partial_{t}f)\circ\psi+\big{(}(D_{y}f)\circ\psi\big{)}\cdot(\partial_{t}\psi)\Big{]}(t,y)\leavevmode\nobreak\ .

Since we have Dy(ρφ)ψ=(Dyρ)(Dyψ)1D_{y}(\rho\circ\varphi)\circ\psi=(D_{y}\rho)(D_{y}\psi)^{-1} for any function ρ\rho, we use the trick

Tr((Dyψ)1tDy(ψ))=Tr(Dy(tψ)(Dyψ)1)=Tr[Dy((tψ)φ)]ψ=divy((tψ)φ)ψ.\operatorname{Tr}\big{(}(D_{y}\psi)^{-1}\partial_{t}D_{y}(\psi)\big{)}=\operatorname{Tr}\big{(}D_{y}(\partial_{t}\psi)(D_{y}\psi)^{-1}\big{)}=\operatorname{Tr}\Big{[}D_{y}\big{(}(\partial_{t}\psi)\circ\varphi\big{)}\Big{]}\circ\psi=\operatorname{div}_{y}\big{(}(\partial_{t}\psi)\circ\varphi\big{)}\circ\psi.

Since (tψ)(t,φ(t,y))=U(t,ψφ(t,y))=U(t,y)(\partial_{t}\psi)(t,\varphi(t,y))=U(t,\psi\circ\varphi(t,y))=U(t,y) due to (4.3), we obtain

t[det(Dyψ(t,y))f(t,ψ(t,y))]\displaystyle\partial_{t}\Big{[}\det\big{(}D_{y}\psi(t,y)\big{)}\,f\big{(}t,\psi(t,y)\big{)}\Big{]} =det(Dyψ(t,y))[divy(U)f+(tf)+(Dyf)U]ψ\displaystyle=\det\big{(}D_{y}\psi(t,y)\big{)}\big{[}\operatorname{div}_{y}(U)f+(\partial_{t}f)+(D_{y}f)U\big{]}\circ\psi (4.4)
=det(Dyψ(t,y))[divy(Uf)+(tf)]ψ.\displaystyle=\det\big{(}D_{y}\psi(t,y)\big{)}\big{[}\operatorname{div}_{y}(Uf)+(\partial_{t}f)\big{]}\circ\psi\leavevmode\nobreak\ . (4.5)

It remains to notice that divy(Uf)=tf\operatorname{div}_{y}(Uf)=-\partial_{t}f by (4.2), that f(t,y0)=1f(t,y_{0})=1 and that det(Dyψ(t0,x))=det(id)=1\det(D_{y}\psi(t_{0},x))=\det({\text{\rm id}})=1. Now, from (4.4), we have det(Dyψ(t,y))f(t,ψ(t,y))=det(Dyψ(t0,y))f(t,ψ(t0,y))=1\det\big{(}D_{y}\psi(t,y)\big{)}\,f\big{(}t,\psi(t,y)\big{)}=\det\big{(}D_{y}\psi(t_{0},y)\big{)}\,f\big{(}t,\psi(t_{0},y)\big{)}=1 and then

tI,xΩ0,det(Dyψ(t,y))=1f(t,ψ(t,y)).\forall t\in I,x\in\Omega_{0}\leavevmode\nobreak\ ,\leavevmode\nobreak\ \leavevmode\nobreak\ \det(D_{y}\psi(t,y))=\frac{1}{f(t,\psi(t,y))}\leavevmode\nobreak\ . (4.6)

We conclude that, as required,

tI,xΩ0,det(Dφ(t,y))=det((Dψ)1(t,φ(t,y)))=f(t,ψφ(t,y))=f(t,y).\forall t\in I,x\in\Omega_{0}\leavevmode\nobreak\ ,\leavevmode\nobreak\ \leavevmode\nobreak\ \det(D\varphi(t,y))=\det((D\psi)^{-1}(t,\varphi(t,y)))=f(t,\psi\circ\varphi(t,y))=f(t,y)\leavevmode\nobreak\ .

Finally, it remains to notice that the bounds on ff yield bounds on UU due to (4.2). The classical bounds on the flow recalled in Proposition A.4 in the appendix provide the claimed bounds on φ\varphi and φ1=ψ\varphi^{-1}=\psi. Indeed, they satisfy (4.3) and the dual ODE where the time is reversed. \square

The trick of the above proof is the use of the flow of the ODE (4.3), which has of course a geometrical background. Actually, this idea can be extended to other differential forms than the volume form considered here, see [15] and [42]. Also notice the explicit bounds stated in our version, which is not usual. We will need them for the proof of Theorem 4.1 because we slightly adapt the strategy of [16].

4.3 The fixed point method

Proposition 4.4 below is an improvement of Proposition 4.3 regarding regularity in both time and space, but it only deals with local perturbations of f(t,y)1f(t,y)\equiv 1. It is proved following the fixed point method of [15, 16].

Proposition 4.4.

Let k1k\geq 1 and r0r\geq 0 with krk\geq r and let α(0,1)\alpha\in(0,1). Let Ω0\Omega_{0} be a bounded connected open 𝒞k+1,α\mathcal{C}^{k+1,\alpha} domain of N\mathbb{R}^{N}. There exists ε>0\varepsilon>0 such that, for any interval II\subset\mathbb{R} and any f𝒞r(I,𝒞k1,α(Ω¯0,+))f\in\mathcal{C}^{r}(I,\mathcal{C}^{k-1,\alpha}(\overline{\Omega}_{0},\mathbb{R}^{*}_{+})) being such that

tI,Ω0f(t,y)dy=meas(Ω0) and f(t,)1𝒞k1,αε.\forall t\in I\leavevmode\nobreak\ ,\leavevmode\nobreak\ \leavevmode\nobreak\ \int_{\Omega_{0}}f(t,y)\,{\text{\rm d}}y={\text{\rm meas}}(\Omega_{0})\leavevmode\nobreak\ \leavevmode\nobreak\ \text{ and }\leavevmode\nobreak\ \leavevmode\nobreak\ \|f(t,\cdot)-1\|_{\mathcal{C}^{k-1,\alpha}}\leq\varepsilon\leavevmode\nobreak\ .

Then, there exists a family (φ(t))tI(\varphi(t))_{t\in I} of diffeomorphisms of Ω¯0\overline{\Omega}_{0} of class 𝒞r(I,𝒞k,α(Ω¯0,Ω¯0))\mathcal{C}^{r}(I,\mathcal{C}^{k,\alpha}(\overline{\Omega}_{0},\overline{\Omega}_{0})) satisfying (4.1). Moreover, there exists a constant KK independent of ff such that

suptIφ(t)𝒞k,α+suptIφ1(t)𝒞k,αKsuptIf(t,)1𝒞k1,α.\sup_{t\in I}\|\varphi(t)\|_{\mathcal{C}^{k,\alpha}}+\sup_{t\in I}\|\varphi^{-1}(t)\|_{\mathcal{C}^{k,\alpha}}\leavevmode\nobreak\ \leq\leavevmode\nobreak\ K\sup_{t\in I}\|f(t,\cdot)-1\|_{\mathcal{C}^{k-1,\alpha}}\leavevmode\nobreak\ .

Proof: We follow Section 5.6.2 of [15]. We reproduce the proof for sake of completeness and to explain why we can add for free the time-dependence. We set

Q:Md()det(id+M)1Tr(M).Q\leavevmode\nobreak\ :\leavevmode\nobreak\ M\in\mathcal{M}_{d}(\mathbb{R})\leavevmode\nobreak\ \longmapsto\leavevmode\nobreak\ \det({\text{\rm id}}+M)-1-\operatorname{Tr}(M)\in\mathbb{R}\leavevmode\nobreak\ .

Notice that QQ is the sum of monomials of degree between 22 and dd with respect to the coefficients because 11 and Tr(M)\operatorname{Tr}(M) are the first order terms of the development of det(id+M)\det({\text{\rm id}}+M). Using the fact that 𝒞k,α\mathcal{C}^{k,\alpha} is a Banach algebra (see for example [14]), there exists a constant K2K_{2} such that, for any functions u,v𝒞k,α(Ω¯,N)u,v\in\mathcal{C}^{k,\alpha}(\overline{\Omega},\mathbb{R}^{N})

Q(Dyu)Q(Dyv)𝒞k1,αK1(1+\displaystyle\|Q(D_{y}u)-Q(D_{y}v)\|_{\mathcal{C}^{k-1,\alpha}}\leq K_{1}(1+ u𝒞k,αd2+v𝒞k,αd2)\displaystyle\|u\|_{\mathcal{C}^{k,\alpha}}^{d-2}+\|v\|_{\mathcal{C}^{k,\alpha}}^{d-2})
max(u𝒞k,α,v𝒞k,α)uv𝒞k,α.\displaystyle\max(\|u\|_{\mathcal{C}^{k,\alpha}},\|v\|_{\mathcal{C}^{k,\alpha}})\|u-v\|_{\mathcal{C}^{k,\alpha}}\leavevmode\nobreak\ . (4.7)

We seek for a function φ(t,y)\varphi(t,y) solving (4.1) for times tt close to t0t_{0} by setting

φ(t,y)=id+η(t,y).\varphi(t,y)={\text{\rm id}}+\eta(t,y)\leavevmode\nobreak\ .

Since we would like that det(id+Dyη)=f\det({\text{\rm id}}+D_{y}\eta)=f and as divη=Tr(Dyη)\operatorname{div}\eta=\operatorname{Tr}(D_{y}\eta), the identity (4.1) is equivalent to

{divη(y,t)=f(t,y)1Q(Dyη(t,y))yΩ0η(t,y)=0yΩ0\left\{\begin{array}[]{ll}\operatorname{div}\eta(y,t)=f(t,y)-1-Q(D_{y}\eta(t,y))\nobreak\leavevmode&y\in\Omega_{0}\\ \eta(t,y)=0&y\in\partial\Omega_{0}\end{array}\right. (4.8)

Let X0k,αX^{k,\alpha}_{0} and Ymk1,αY^{k-1,\alpha}_{m} the spaces introduced in Appendix A.3. First notice that, by assumption,

Ω0f(t,y)dy=meas(Ω0)=Ω01dy\int_{\Omega_{0}}f(t,y)\,{\text{\rm d}}y={\text{\rm meas}}(\Omega_{0})=\int_{\Omega_{0}}1\,{\text{\rm d}}y

and thus (f(t)1)(f(t)-1) belongs to Ymk1,αY^{k-1,\alpha}_{m}. Since f(t,)f(t,\cdot) is close to 11, we seek for η\eta small in 𝒞r(I,𝒞k,α(Ω¯0,N))\mathcal{C}^{r}(I,\mathcal{C}^{k,\alpha}(\overline{\Omega}_{0},\mathbb{R}^{N})). By Corollary A.8, there exists R>0R>0 small such that if η(t,)𝒞k,αR\|\eta(t,\cdot)\|_{\mathcal{C}^{k,\alpha}}\leq R for all tIt\in I, then φ(t,y)=id+η(t,y)\varphi(t,y)={\text{\rm id}}+\eta(t,y) is a diffeomorphism from Ω0\Omega_{0} onto itself. If this is the case, we have that

Ω0Q(Dyη(t,y))dy\displaystyle\int_{\Omega_{0}}Q(D_{y}\eta(t,y))\,{\text{\rm d}}y =Ω0(det(id+Dyη)(t,y)1+Tr(Dyη(t,y)))dy\displaystyle=\int_{\Omega_{0}}\big{(}\det({\text{\rm id}}+D_{y}\eta)(t,y)-1+\operatorname{Tr}(D_{y}\eta(t,y))\big{)}\,{\text{\rm d}}y
=Ω0det(Dyφ(t,y))dymeas(Ω0)+Ω0div(η(t,y))dy\displaystyle=\int_{\Omega_{0}}\det(D_{y}\varphi(t,y))\,{\text{\rm d}}y-{\text{\rm meas}}(\Omega_{0})+\int_{\Omega_{0}}\operatorname{div}(\eta(t,y))\,{\text{\rm d}}y
=Ω0𝟙(φ(y))dymeas(Ω0)+Ω0η(y)dy\displaystyle=\int_{\Omega_{0}}\mathbbm{1}(\varphi(y))\,{\text{\rm d}}y-{\text{\rm meas}}(\Omega_{0})+\int_{\partial\Omega_{0}}\eta(y)\,{\text{\rm d}}y
=meas(Ω0)meas(Ω0)+0=0.\displaystyle={\text{\rm meas}}(\Omega_{0})-{\text{\rm meas}}(\Omega_{0})+0=0\leavevmode\nobreak\ .

This means that we can look for η\eta small as a solution in 𝒞r(I,𝒞k,α(Ω¯0,N))\mathcal{C}^{r}(I,\mathcal{C}^{k,\alpha}(\overline{\Omega}_{0},\mathbb{R}^{N})) of

η(t)=L1(f(t)1Q(Dyη(t)))\eta(t)\,=\,L^{-1}\big{(}f(t)-1-Q(D_{y}\eta(t))\big{)} (4.9)

since we expect Q(Dyη(t))Q(D_{y}\eta(t)) to belong to Ymk1,αY^{k-1,\alpha}_{m}. We construct η(t)\eta(t) by a fixed point argument by applying Theorem A.3 to the function

Φ:(η,t)R×IL1(f(t)1Q(Dyη))X0k,α\Phi\leavevmode\nobreak\ :\leavevmode\nobreak\ (\eta,t)\in\mathcal{B}_{R}\times I\leavevmode\nobreak\ \longmapsto\leavevmode\nobreak\ L^{-1}\big{(}f(t)-1-Q(D_{y}\eta)\big{)}\in X^{k,\alpha}_{0}

where L1L^{-1} is the right-inverse of the divergence (see Appendix A.3) and R={ηX0k,α,ηXk,αR}\mathcal{B}_{R}=\{\eta\in X^{k,\alpha}_{0},\|\eta\|_{X^{k,\alpha}}\leq R\} with RR as above. As a consequence, id+η{\text{\rm id}}+\eta is a diffeomorphism and the above computations are valid. Due to (4.7), we have

Φ(η1,t)Φ(η2,t)𝒞k,αRK1K2(1+2Rd2)η1η2𝒞k,α\|\Phi(\eta_{1},t)-\Phi(\eta_{2},t)\|_{\mathcal{C}^{k,\alpha}}\leavevmode\nobreak\ \leq\leavevmode\nobreak\ RK_{1}K_{2}(1+2R^{d-2})\|\eta_{1}-\eta_{2}\|_{\mathcal{C}^{k,\alpha}}

where K2=|L1|K_{2}=|||L^{-1}|||. Using that Q(0)=0Q(0)=0, we also have

Φ(η,t)𝒞k,αK2(f(t)1𝒞k1,α+K1(R2+Rd)).\|\Phi(\eta,t)\|_{\mathcal{C}^{k,\alpha}}\leavevmode\nobreak\ \leq\leavevmode\nobreak\ K_{2}\big{(}\|f(t)-1\|_{\mathcal{C}^{k-1,\alpha}}+K_{1}(R^{2}+R^{d})\big{)}\leavevmode\nobreak\ .

We choose R(0,1/2]R\in(0,1/2] small enough such that id+η{\text{\rm id}}+\eta is a diffeomorphism and

RK1K2(1+2Rd2)12RK_{1}K_{2}(1+2R^{d-2})\leavevmode\nobreak\ \leq\leavevmode\nobreak\ \frac{1}{2}

which also implies that K1K2(R2+Rd)R/2K_{1}K_{2}(R^{2}+R^{d})\leq R/2. Then, we choose ε=R/2K2\varepsilon=R/2K_{2} and assume that

tI,f(t)1𝒞k1,αε=R2K1.\forall t\in I\leavevmode\nobreak\ ,\leavevmode\nobreak\ \leavevmode\nobreak\ \|f(t)-1\|_{\mathcal{C}^{k-1,\alpha}}\leavevmode\nobreak\ \leq\leavevmode\nobreak\ \varepsilon=\frac{R}{2K_{1}}\leavevmode\nobreak\ . (4.10)

By construction, Φ\Phi is 1/21/2-lipschitzian from R\mathcal{B}_{R} into R\mathcal{B}_{R}. The classical fixed point theorem for contraction maps shows the existence of η(t)\eta(t) solving (4.9) for each tt. The regularity of η\eta with respect to the time tt is directly given by Theorem A.3. By construction η(t)=L1(f(t)1Q(Dyη(t))\eta(t)=L^{-1}(f(t)-1-Q(D_{y}\eta(t)), and since LL1=idLL^{-1}={\text{\rm id}}, this implies that divη(t)=f(t)1Q(Dyη(t))\operatorname{div}\eta(t)=f(t)-1-Q(D_{y}\eta(t)) and thus f(t)=det(id+η(t))=det(φ(t))f(t)=\det({\text{\rm id}}+\eta(t))=\det(\varphi(t)). Also remember that, by construction, φ\varphi is 𝒞k\mathcal{C}^{k} close enough to the identity in Ω0\Omega_{0} and is the identity on Ω0\partial\Omega_{0} so that φ(t)\varphi(t) is a 𝒞k\mathcal{C}^{k}-diffeomorphism (see Appendix A.6 and the topological arguments of [41]). \square

4.4 Proof of Theorem 4.1

As in the work of Dacorogna and Moser (see [15] and [16]), we obtain Theorem 4.1 by combining the propositions 4.3 and 4.4. In the original method, the authors used first the fixed point method and then the flow method. However, this approach does not provide the optimal time regularity. Therefore, we couple the two methods in the other sense. In this case, we need the 𝒞k1\mathcal{C}^{k-1}-bounds on the diffeomorphism provided by Proposition 4.3. This is the reason why we made them explicit, which is not common for this type of results.

Let f𝒞r(I,𝒞k1,α(Ω¯0,+))f\in\mathcal{C}^{r}(I,\mathcal{C}^{k-1,\alpha}(\overline{\Omega}_{0},\mathbb{R}^{*}_{+})) with Ωf(t,y)dy=meas(Ω0)\int_{\Omega}f(t,y)\,{\text{\rm d}}y={\text{\rm meas}}(\Omega_{0}) for all tIt\in I. Let ε>0\varepsilon>0. We consider a regularization f1f_{1} of ff, which is of class 𝒞r(I,𝒞k1(Ω¯0,+))\mathcal{C}^{r^{\prime}}(I,\mathcal{C}^{k^{\prime}-1}(\overline{\Omega}_{0},\mathbb{R}^{*}_{+})) with r=max(1,r)r^{\prime}=\max(1,r) and k=k+2k^{\prime}=k+2, satisfying

tI,f(t,)f1(t,)1𝒞k1,αε(t).\forall t\in I\leavevmode\nobreak\ ,\leavevmode\nobreak\ \leavevmode\nobreak\ \left\|\frac{f(t,\cdot)}{f_{1}(t,\cdot)}-1\right\|_{\mathcal{C}^{k-1,\alpha}}\leq\varepsilon(t)\leavevmode\nobreak\ . (4.11)

Assume that we first apply the flow method of Proposition 4.3 to obtain a smooth family of diffeomorphisms φ1𝒞r(I,𝒞k1(Ω¯0,Ω¯0))𝒞r(I,𝒞k+1Ω¯0,Ω¯0))\varphi_{1}\in\mathcal{C}^{r^{\prime}}(I,\mathcal{C}^{k^{\prime}-1}(\overline{\Omega}_{0},\overline{\Omega}_{0}))\subset\mathcal{C}^{r}(I,\mathcal{C}^{k+1}\overline{\Omega}_{0},\overline{\Omega}_{0})) such that

det(Dyφ1(t,y))=f1(t,y) and φ1|Ω0(t,)=id|Ω0.\det(D_{y}\varphi_{1}(t,y))=f_{1}(t,y)\leavevmode\nobreak\ \leavevmode\nobreak\ \text{ and }\leavevmode\nobreak\ \leavevmode\nobreak\ \varphi_{1|\partial\Omega_{0}}(t,\cdot)={\text{\rm id}}_{|\partial\Omega_{0}}\leavevmode\nobreak\ .

Now, we seek for φ\varphi of the form φ=φ2φ1\varphi=\varphi_{2}\circ\varphi_{1} satisfying the statement of Theorem 4.1. Thus, we need to find φ2𝒞r(I,𝒞k,α(Ω¯0,+))\varphi_{2}\in\mathcal{C}^{r}(I,\mathcal{C}^{k,\alpha}(\overline{\Omega}_{0},\mathbb{R}^{*}_{+})) such that

det(Dyφ2)(t,φ1(t,y))det(Dyφ1(t,y))=f(t,y)\det(D_{y}\varphi_{2})(t,\varphi_{1}(t,y)){\det(D_{y}\varphi_{1}(t,y))}={f(t,y)}

i.e.

det(Dyφ2)(t,y)=f(t,φ11(t,y))f1(t,φ11(t,y)).\det(D_{y}\varphi_{2})(t,y)=\frac{f(t,\varphi_{1}^{-1}(t,y))}{f_{1}(t,\varphi_{1}^{-1}(t,y))}\leavevmode\nobreak\ . (4.12)

We would like to apply the fixed point method of Proposition 4.4 by choosing ε(t)\varepsilon(t) small enough such that

tI,f(t,φ11(t,y))f1(t,φ11(t,y))1𝒞k1,αε\forall t\in I\leavevmode\nobreak\ ,\leavevmode\nobreak\ \leavevmode\nobreak\ \left\|\frac{f(t,\varphi_{1}^{-1}(t,y))}{f_{1}(t,\varphi_{1}^{-1}(t,y))}-1\right\|_{\mathcal{C}^{k-1,\alpha}}\leq\varepsilon (4.13)

with ε\varepsilon as in Proposition 4.4. However, we notice that in (4.13) appears the composition with φ11\varphi_{1}^{-1}. Fix ε\varepsilon as in Proposition 4.4 and smaller than 1/21/2. When we consider Proposition 4.3 with M=2M=2, we have the upper bound for C(2)eλ(2)tC(2)e^{\lambda(2)t} for the derivatives of any φ1\varphi_{1} constructed as above when ε(t)ε\varepsilon(t)\leq\varepsilon in (4.11). In particular, these bounds indicate how much the composition by φ1\varphi^{-1} increases the 𝒞k1,α\mathcal{C}^{k-1,\alpha}-norm. Having this in mind, we may choose ε(t)\varepsilon(t) small (and exponentially decreasing) such that, for any f1f_{1} satisfying (4.11), the associated diffeomorphisms φ1\varphi_{1} are such that (4.13) holds.

We construct the diffeomorphism as follows. We choose a regularization f1f_{1} of the function ff, which is of class 𝒞r(I,𝒞k1(Ω¯0,+))\mathcal{C}^{r^{\prime}}(I,\mathcal{C}^{k^{\prime}-1}(\overline{\Omega}_{0},\mathbb{R}^{*}_{+})) with r=max(1,r)r^{\prime}=\max(1,r) and k=k+2k^{\prime}=k+2, satisfying (4.11) for the suitable ε(t)\varepsilon(t). We will also need that Ω0(f/f1)(t,y)dy=meas(Ω0)\int_{\Omega_{0}}(f/f_{1})(t,y)\,{\text{\rm d}}y={\text{\rm meas}}(\Omega_{0}), which is provided by first choosing f~\tilde{f} satisfying (4.11) with a smaller ε(t)\varepsilon(t) and then set f1=f~(f/f~)/meas(Ω0)f_{1}=\tilde{f}\cdot\int(f/\tilde{f})/{\text{\rm meas}}(\Omega_{0}). By Proposition 4.3, there exists a family of diffeomorphisms tφ1(t,)t\mapsto\varphi_{1}(t,\cdot) satisfying (4.11) and of class 𝒞r(I,𝒞k1(Ω¯0,Ω¯0))𝒞r(I,𝒞k+1(Ω¯0,Ω¯0))\mathcal{C}^{r^{\prime}}(I,\mathcal{C}^{k^{\prime}-1}(\overline{\Omega}_{0},\overline{\Omega}_{0}))\subset\mathcal{C}^{r}(I,\mathcal{C}^{k+1}(\overline{\Omega}_{0},\overline{\Omega}_{0})).

Now we consider f2=f/f1f_{2}=f/f_{1} which satisfies by construction (4.13) with ε\varepsilon as in Proposition 4.4. Notice that f2f_{2} is of class 𝒞r(I,𝒞k1,α(Ω¯0,+))\mathcal{C}^{r}(I,\mathcal{C}^{k-1,\alpha}(\overline{\Omega}_{0},\mathbb{R}^{*}_{+})) and

Ω0f2(t,y)dy\displaystyle\int_{\Omega_{0}}f_{2}(t,y)\,{\text{\rm d}}y =Ω0f(t,φ11(t,y))det(Dyφ1(t,φ11(t,y)))dy\displaystyle=\int_{\Omega_{0}}\frac{f(t,\varphi_{1}^{-1}(t,y))}{\det(D_{y}\varphi_{1}(t,\varphi_{1}^{-1}(t,y)))}\,{\text{\rm d}}y
=Ω0f(t,φ11(t,y))det(Dy(φ11(t,y)))dy\displaystyle=\int_{\Omega_{0}}{f(t,\varphi_{1}^{-1}(t,y))}{\det(D_{y}(\varphi_{1}^{-1}(t,y)))}\,{\text{\rm d}}y
=Ω0f(t,y)dy=meas(Ω0).\displaystyle=\int_{\Omega_{0}}f(t,y)\,{\text{\rm d}}y={\text{\rm meas}}(\Omega_{0})\leavevmode\nobreak\ .

Thus, we can apply Proposition 4.4 to obtain a family of diffeomorphisms φ2\varphi_{2} of class 𝒞r(I,𝒞k,α(Ω¯0,Ω¯0))\mathcal{C}^{r}(I,\mathcal{C}^{k,\alpha}(\overline{\Omega}_{0},\overline{\Omega}_{0})) such that

det(Dyφ2(t,y))=f2(t,y) and φ2|Ω0(t,)=id|Ω0.\det(D_{y}\varphi_{2}(t,y))=f_{2}(t,y)\leavevmode\nobreak\ \leavevmode\nobreak\ \text{ and }\leavevmode\nobreak\ \leavevmode\nobreak\ \varphi_{2|\partial\Omega_{0}}(t,\cdot)={\text{\rm id}}_{|\partial\Omega_{0}}\leavevmode\nobreak\ .

By construction, φ=φ2φ1\varphi=\varphi_{2}\circ\varphi_{1} satisfies the conclusion of Theorem 4.1.

5 Some applications of our results

5.1 Adiabatic dynamics for quantum states on moving domains

In this subsection, we show how to ensure an adiabatic result for the Schrödinger equation on a moving domain as (1.9). We consider the framework of Corollary 1.5. We denote by H(τ)=ΔH(\tau)=-\Delta the Dirichlet Laplacian in L2(Ω(τ),)L^{2}(\Omega(\tau),\mathbb{C}), i.e. D(H(τ))=H2(Ω(τ))H01(Ω(τ))D(H(\tau))=H^{2}(\Omega(\tau))\cap H^{1}_{0}(\Omega(\tau)) and τ[0,1]λ(τ)\tau\in[0,1]\mapsto\lambda(\tau), a continuous curve such that λ(τ)\lambda(\tau) for every τ[0,1]\tau\in[0,1] is in the discrete spectrum of H(τ)H(\tau). We also assume that λ(τ)\lambda(\tau) is a simple isolated eigenvalue for every τ[0,1]\tau\in[0,1], associated with the spectral projectors P(τ)P(\tau).

We consider on the time interval [0,1/ϵ][0,1/\epsilon] the equation (3.7) in L2(Ω0,)L^{2}(\Omega_{0},\mathbb{C}) when H(t)H(t) is a Dirichlet Laplacian. Fixed ϵ>0,\epsilon>0, we substitute tt by τ=ϵt\tau=\epsilon t and set v~ϵ(τ)=v(τ/ϵ)=huϵ(τ/ϵ)\tilde{v}_{\epsilon}(\tau)=v(\tau/\epsilon)=h^{\sharp}u_{\epsilon}(\tau/\epsilon) to obtain

{iϵτv~ϵ(τ)=(hH(τ)h+ϵ(τ))v~ϵ(τ),τ[0,1]v~ϵ(τ)|Ω00,v~ϵ(τ=0)=hu0\begin{split}\begin{cases}i\epsilon\partial_{\tau}\tilde{v}_{\epsilon}(\tau)=\Big{(}h^{\sharp}H(\tau)h_{\sharp}+\epsilon\mathcal{H}(\tau)\Big{)}\tilde{v}_{\epsilon}(\tau),&\ \ \ \ \ \ \ \ \tau\in[0,1]\\ \tilde{v}_{\epsilon}(\tau)_{|\partial\Omega_{0}}\equiv 0,&\\ \tilde{v}_{\epsilon}(\tau=0)=h^{\sharp}u_{0}&\end{cases}\end{split} (5.1)

where

(τ)=h[idivx(Ah(τ,x))+iAh(τ,x)|x]h,Ah(τ,x):=12(hτh)(τ,x).\mathcal{H}(\tau)=-h^{\sharp}\Big{[}i\operatorname{div}_{x}\big{(}A_{h}(\tau,x)\cdot\big{)}\leavevmode\nobreak\ +\leavevmode\nobreak\ i\big{\langle}A_{h}(\tau,x)\big{|}\nabla_{x}\cdot\big{\rangle}\Big{]}h_{\sharp}\leavevmode\nobreak\ ,\ \ \ \ \ \ A_{h}(\tau,x):=-\frac{1}{2}(h_{*}\partial_{\tau}h)(\tau,x).

The problem (5.1) generates a unitary flow thanks to Section 3. Even though it is well known that the classical adiabatic theorem is valid for the dynamics iϵτuϵ(τ)=hH(τ)huϵ(t)i\epsilon\partial_{\tau}u_{\epsilon}(\tau)=h^{\sharp}H(\tau)h_{\sharp}u_{\epsilon}(t), (see [11, Chapter 4] or [52]), we may wonder if it is the same for the equation (5.1) because

H~ϵ(τ)=hH(τ)h+ϵ(τ)\tilde{H}_{\epsilon}(\tau)=h^{\sharp}H(\tau)h_{\sharp}+\epsilon\mathcal{H}(\tau)

depends on ϵ\epsilon and no spectral assumptions have been made on this family. First, we notice that, by conjugation, λ(τ)\lambda(\tau) also belongs to the discrete spectrum of the operator hH(τ)hh^{\sharp}H(\tau)h_{\sharp} in L2(Ω0,)L^{2}(\Omega_{0},\mathbb{C}) and is associated with the spectral projection (hPh)(τ)(h^{\sharp}Ph_{\sharp})(\tau). Then, for each τ\tau, H~ϵ(τ)\tilde{H}_{\epsilon}(\tau) is a small relatively compact self-adjoint perturbation of hH(τ)hh^{\sharp}H(\tau)h_{\sharp}. Thus, for all ϵ>0\epsilon>0 small enough, there exists a curve λ~ϵ(τ)\tilde{\lambda}_{\epsilon}(\tau) of simple isolated eigenvalues of H~ϵ(τ)\tilde{H}_{\epsilon}(\tau), associated with spectral projectors P~ϵ(τ)\tilde{P}_{\epsilon}(\tau), such that λ~ϵ\tilde{\lambda}_{\epsilon} and P~ϵ\tilde{P}_{\epsilon} converge uniformly when ϵ0\epsilon\rightarrow 0 to λ\lambda and hPhh^{\sharp}Ph_{\sharp} respectively (see [33] for further details). In this framework, even if H~ϵ(τ)\tilde{H}_{\epsilon}(\tau) depends on ϵ\epsilon, it is known that the classical adiabatic arguments can still be applied (see for example Nenciu [45, Remarks; p. 16; (4)], Teufel [52, Theorem 4.15] or the works [2, 24, 32]). Thus, we obtain the following convergence for the solution of (5.1)

P~ϵ(1)v~ϵ(1)|v~ϵ(1)ϵ0P~0(0)v(0)|v(0)=(hPh)(0).hu0|hu0=P(0)u0|u0.\langle\tilde{P}_{\epsilon}(1)\tilde{v}_{\epsilon}(1)|\tilde{v}_{\epsilon}(1)\rangle\leavevmode\nobreak\ \leavevmode\nobreak\ \xrightarrow[\leavevmode\nobreak\ \leavevmode\nobreak\ \epsilon\longrightarrow 0\leavevmode\nobreak\ \leavevmode\nobreak\ ]{}\leavevmode\nobreak\ \leavevmode\nobreak\ \langle\tilde{P}_{0}(0)v(0)|v(0)\rangle\leavevmode\nobreak\ =\leavevmode\nobreak\ \langle(h^{\sharp}Ph_{\sharp})(0).h^{\sharp}u_{0}|h^{\sharp}u_{0}\rangle\leavevmode\nobreak\ =\leavevmode\nobreak\ \langle P(0)u_{0}|u_{0}\rangle\leavevmode\nobreak\ .

Finally, we notice that, since P~ϵ(1)\tilde{P}_{\epsilon}(1) converges to hP(1)hh^{\sharp}P(1)h_{\sharp} when ϵ\epsilon goes to zero,

P~ϵ(1)v~ϵ(1)|v~ϵ(1)=(hP(1)h)v~ϵ(1)|v~ϵ(1)+o(1)=P(1)uϵ(1/ϵ)|uϵ(1/ϵ)+o(1).\langle\tilde{P}_{\epsilon}(1)\tilde{v}_{\epsilon}(1)|\tilde{v}_{\epsilon}(1)\rangle\leavevmode\nobreak\ =\leavevmode\nobreak\ \langle(h^{\sharp}P(1)h_{\sharp})\tilde{v}_{\epsilon}(1)|\tilde{v}_{\epsilon}(1)\rangle\leavevmode\nobreak\ +\leavevmode\nobreak\ o(1)\leavevmode\nobreak\ =\leavevmode\nobreak\ \langle P(1)u_{\epsilon}(1/\epsilon)|u_{\epsilon}(1/\epsilon)\rangle\leavevmode\nobreak\ +\leavevmode\nobreak\ o(1)\leavevmode\nobreak\ .

This concludes the proof of Corollary 1.5.

5.2 Explicit examples of time-varying domains


Translation of a potential well
Let us consider any domain Ω0N\Omega_{0}\subset\mathbb{R}^{N} and any smooth family of vectors D(t)𝒞2([0,T],N)D(t)\in\mathcal{C}^{2}([0,T],\mathbb{R}^{N}). The family of translated domains is Ω(t)=h(t,Ω0)\Omega(t)=h(t,\Omega_{0}) where h(t,y)=y+D(t)h(t,y)=y+D(t). By explicit computations, we obtain that hΔxh=Δyh^{\sharp}\Delta_{x}h_{\sharp}=\Delta_{y} and (hth)(t,x)=D(t)(h_{*}\partial_{t}h)(t,x)=D^{\prime}(t). Since |J||J| does not depend on yy, we do not need Moser’s trick to get (1.17). We can apply the gauge transformation since ϕ(t,x)=12D(t)|x\phi(t,x)=\frac{1}{2}\langle D(t)|x\rangle satisfies hth=2xϕh_{*}\partial_{t}h=2\nabla_{x}\phi. Then, w=heiϕuw=h^{\sharp}e^{-i\phi}u satisfies Equation (1.15), which becomes in our framework

itw=Δyw+14(2D′′(t)|(y+D(t))+|D(t)|2)w.i\partial_{t}w\leavevmode\nobreak\ =\leavevmode\nobreak\ -\Delta_{y}w\leavevmode\nobreak\ +\leavevmode\nobreak\ \frac{1}{4}\Big{(}2\langle D^{\prime\prime}(t)|(y+D(t))\rangle+|D(t)|^{2}\Big{)}w. (5.2)

In this very particular case, we can further simplify the expression thanks to an interesting fact. Two terms of the electric potential in (5.2) do not depend on the space variable. We can thus apply another transformation to the system by adding a phase which is an antiderivative of 2D′′(t)|D(t)+|D(t)|22\langle D^{\prime\prime}(t)|D(t)\rangle+|D(t)|^{2}. For example, we consider

w~=ei4(2D(t)|D(t)0t|D(s)|2ds)w\tilde{w}\leavevmode\nobreak\ =\leavevmode\nobreak\ e^{\frac{i}{4}(2\langle D^{\prime}(t)|D(t)\rangle-\int_{0}^{t}|D(s)|^{2}\,{\text{\rm d}}s)}w

where ww satisfies Equation (5.2). Then, w~\tilde{w} is solution of the equation

itw~=Δyw~+12D′′(t)|yw~.i\partial_{t}\tilde{w}\leavevmode\nobreak\ =\leavevmode\nobreak\ -\Delta_{y}\tilde{w}\leavevmode\nobreak\ +\leavevmode\nobreak\ \frac{1}{2}\langle D^{\prime\prime}(t)|y\rangle\tilde{w}. (5.3)

These explicit computations are not new and appear for example in [6] for the one dimensional case of a moving interval.


Rotating domains
Let us consider a family of rotating domains in 2\mathbb{R}^{2}. Clearly, the same results can be extended by considering rotations in N\mathbb{R}^{N} with N3N\geq 3. Let Ω02\Omega_{0}\subset\mathbb{R}^{2} and let Ω(t)=h(t,Ω0)\Omega(t)=h(t,\Omega_{0}) with

h(t):y=(y1,y2)Ω0(cos(ωt)y1sin(ωt)y2,cos(ωt)y1+sin(ωt)y2)h(t):\leavevmode\nobreak\ y=(y_{1},y_{2})\in\Omega_{0}\leavevmode\nobreak\ \longmapsto\leavevmode\nobreak\ \big{(}\cos(\omega t)y_{1}-\sin(\omega t)y_{2},\cos(\omega t)y_{1}+\sin(\omega t)y_{2}\big{)}

and ω\omega\in\mathbb{R}. Using the classical notation (y1,y2)=(y2,y1)(y_{1},y_{2})^{\perp}=(-y_{2},y_{1}), it is straightforward to check that |J|=1|J|=1, J1(J1)t=IdJ^{-1}(J^{-1})^{t}=Id,

(hth)(t,x)=ωx and J1th(t,y)=ωy.(h_{*}\partial_{t}h)(t,x)=\omega x^{\perp}\leavevmode\nobreak\ \leavevmode\nobreak\ \text{ and }\leavevmode\nobreak\ \leavevmode\nobreak\ J^{-1}\partial_{t}h(t,y)=\omega y^{\perp}\leavevmode\nobreak\ .

We obtain by direct computation or by the first line of (1.17) that v=huv=h^{*}u satisfies the following Schrödinger equation in the rotating frame

itv=Δyv+iωy,yv.i\partial_{t}v\leavevmode\nobreak\ =\leavevmode\nobreak\ -\Delta_{y}v\leavevmode\nobreak\ +\leavevmode\nobreak\ i\omega\langle y^{\perp},\nabla_{y}v\rangle\leavevmode\nobreak\ . (5.4)

This is an obvious and well-known computation (used for studying quantum systems in rotating potentials frames). The general Hamiltonian structure highlighted in this paper simply writes here as

itv=(yiω2y)2vω24|y|2v.i\partial_{t}v\leavevmode\nobreak\ =\leavevmode\nobreak\ -\Big{(}\nabla_{y}-i\frac{\omega}{2}y^{\perp}\Big{)}^{2}v\leavevmode\nobreak\ -\leavevmode\nobreak\ \frac{\omega^{2}}{4}|y|^{2}v\leavevmode\nobreak\ .

which is given by the second line of (1.17) or obtained directly from (5.4) by using the fact that yy^{\perp} is divergence free. We recover a repulsive potential ω2|y|2/4\omega^{2}|y|^{2}/4 corresponding to the centrifugal force.


Moving domains with diagonal diffeomorphisms
Let Ω0N\Omega_{0}\subset\mathbb{R}^{N} and let Ω(t)=h(t,Ω0)\Omega(t)=h(t,\Omega_{0}) with

h(t):y=(yi)i=1NΩ0(fi(t)yi)i=1Nh(t):\leavevmode\nobreak\ y=(y_{i})_{i=1\ldots N}\in\Omega_{0}\leavevmode\nobreak\ \longmapsto\leavevmode\nobreak\ \big{(}f_{i}(t)\,y_{i}\big{)}_{i=1\ldots N}

and fi𝒞2([0,T],)f_{i}\in\mathcal{C}^{2}([0,T],\mathbb{R}). As above, we obtain

hΔxh=i=1N1fi(t)2yiyi2,(hth)(t,x)=(fi(t)fi(t)xi)i=1N.h^{\sharp}\Delta_{x}h_{\sharp}=\sum_{i=1}^{N}\frac{1}{f_{i}(t)^{2}}\partial_{y_{i}y_{i}}^{2},\ \ \ \ \ (h_{*}\partial_{t}h)(t,x)=\left(\frac{f_{i}^{\prime}(t)}{f_{i}(t)}\,x_{i}\right)_{i=1\ldots N}. (5.5)

We apply again the gauge transformation and then,

ϕ(t,x)=14j=1N(fi(t)fi(t)xi2)satisfieshth=2xϕ.\phi(t,x)=\frac{1}{4}\sum_{j=1}^{N}\left(\frac{f_{i}^{\prime}(t)}{f_{i}(t)}\,x_{i}^{2}\right)\ \ \ \ \ \text{satisfies}\ \ \ \ \ h_{*}\partial_{t}h=2\nabla_{x}\phi. (5.6)

Finally, w=heiϕuw=h^{\sharp}e^{-i\phi}u satisfies the equation

itw=i=1N1fi(t)2yiyi2w+14(i=1Nfi′′(t)fi(t)yi2)w.i\partial_{t}w\leavevmode\nobreak\ =\leavevmode\nobreak\ -\sum_{i=1}^{N}\frac{1}{f_{i}(t)^{2}}\partial_{y_{i}y_{i}}^{2}w\leavevmode\nobreak\ +\leavevmode\nobreak\ \frac{1}{4}\left(\sum_{i=1}^{N}f^{\prime\prime}_{i}(t)f_{i}(t)y_{i}^{2}\right)w. (5.7)

In the homothetical case where f1=f2==:f(t)f_{1}=f_{2}=\ldots=:f(t) (see for instance [3, 5, 7, 38, 40, 43, 47, 50]). Equation (5.7) becomes

itw(t)=1f(t)2Δw(t)+14f′′(t)f(t)|y|2w(t),tI.i\partial_{t}w(t)=-\frac{1}{f(t)^{2}}\Delta w(t)+\frac{1}{4}f^{\prime\prime}(t)f(t)|y|^{2}w(t),\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ t\in I.

In this case, it is usual to make a further simplification to eliminate the time-dependence of the main operator by changing the time variable for

τ=0t1f(s)2𝑑s\tau=\int_{0}^{t}\frac{1}{f(s)^{2}}ds

and introducing the implicitly defined function

U(τ)=f(t)f(t)4.U(\tau)=\frac{f^{\prime}(t)f(t)}{4}.

We obtain

iτw(τ)=Δw(τ)+(U(τ)4U(τ))|y|2w(τ),τ[ 0,0T1/f(s)2𝑑s].i\partial_{\tau}w(\tau)=-\Delta w(\tau)+\Big{(}U^{\prime}(\tau)-4U(\tau)\Big{)}|y|^{2}w(\tau),\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \tau\in\Big{[}\,0\,,\,\int_{0}^{T}1/{f(s)^{2}}ds\Big{]}. (5.8)

In this simple case, we see that the general framework of this paper coincides with the previous computations introduced in dimension d=1d=1 by Beauchard in [5]. Indeed, the transformations described in [5, Section 1.3] corresponds to application uw=heiϕuu\longmapsto w=h^{\sharp}e^{-i\phi}u as the multiplication for the square root of the exponential [5, relation (1.4)] is nothing else than the multiplication for the square root of the Jacobian appearing in the definition of hh^{\sharp}. Our paper put the change of variable of [5] in a more general geometric framework.

A similar expression to (5.8) is also obtained by Moyano in [43] for the case of the two-dimensional disk, nevertheless the transformations adopted in [43] are different from the ones considered in our work. In particular, they are not unitary with respect to the classical L2L^{2}-norm.

For a second application of this simple case, we consider the case of a family of cylinders

Ω(t)={(x1,x2,x3)+3x1(0,(t)),x22+x32<r2,}\Omega(t)=\Big{\{}(x_{1},x_{2},x_{3})\in\mathbb{R}_{+}^{3}\ \>\ x_{1}\in(0,\ell(t)),\ x_{2}^{2}+x_{3}^{2}<r^{2},\ \Big{\}}

for \ell a 𝒞2\mathcal{C}^{2}-varying length. We would like to consider the Schrödinger equation itu=Δxui\partial_{t}u=-\Delta_{x}u in Ω(t)\Omega(t) with boundary conditions of the Neumann type. This example corresponds to the situation of Figure 2 in Section 1. As shown in this paper, the conditions at the boundaries cannot be pure homogeneous Neumann ones everywhere if (t)\ell(t) is not constant. Theorem 1.2 shows us the correct ones. We can choose

h(t,):(y1,y2,y3)Ω0((t)y1,y2,y3)Ω(t)h(t,\cdot):\big{(}y_{1},y_{2},y_{3}\big{)}\in\Omega_{0}\longmapsto\big{(}\ell(t)y_{1},y_{2},y_{3}\big{)}\in\Omega(t)

with Ω0\Omega_{0} the cylinder of length 11. Then, we get by (5.5) the term hthh_{*}\partial_{t}h and we can compute that the suitable boundary conditions are

νui2(t)u=0 at xΩ(t) with x1=(t)\partial_{\nu}u-\frac{i}{2}\ell^{\prime}(t)u=0\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \text{ at }x\in\partial\Omega(t)\text{ with }x_{1}=\ell(t) (5.9)

and νu=0\partial_{\nu}u=0 on the other parts of the boundary (see Figure 2). It is important to notice that, contrary to the above changes of variables, this computation is independent of the choice of h(t)h(t). Equations as (5.8) can be seen as auxiliary equations and they depend on several choices, while (5.9) is stated for the original variable uu and have physical meaning.

Appendix A Appendix

A.1 Unitary semigroups

Defining solutions of an evolution equation with a time-dependent family of operators is nowadays a classical result (see [51]). In the present article, we are interested in the Hamiltonian structure and we use the following result of [37] (see also [52]).

Theorem A.1.

(Kisyński, 1963)
Let 𝒳\mathcal{X} be a Hilbert space and let ((t))t[0,T](\mathcal{H}(t))_{t\in[0,T]} be a family of self-adjoint positive operators on 𝒳\mathcal{X} such that 𝒳1/2=D((t)1/2)\mathcal{X}^{1/2}=D(\mathcal{H}(t)^{1/2}) is independent of time tt. Also set 𝒳1/2=D((t)1/2)=(𝒳1/2)\mathcal{X}^{-1/2}=D(\mathcal{H}(t)^{-1/2})=(\mathcal{X}^{1/2})^{*} and assume that (t):𝒳1/2𝒳1/2\mathcal{H}(t):\mathcal{X}^{1/2}\rightarrow\mathcal{X}^{-1/2} is of class 𝒞1\mathcal{C}^{1} with respect to t[0,T]t\in[0,T]. In other words, we assume that the sesquilinear form ϕt(u,v)=(t)u|v\phi_{t}(u,v)=\langle\mathcal{H}(t)u|v\rangle associated with (t)\mathcal{H}(t) has a domain 𝒳1/2\mathcal{X}^{1/2} independent of the time and is of class 𝒞1\mathcal{C}^{1} with respect to tt. Also assume that there exist γ>0\gamma>0 and κ\kappa\in\mathbb{R} such that,

t[0,T],u𝒳1/2,ϕt(u,u)=(t)u|u𝒳γu𝒳1/22κu𝒳2.\forall t\in[0,T]\leavevmode\nobreak\ ,\leavevmode\nobreak\ \forall u\in\mathcal{X}^{1/2}\leavevmode\nobreak\ ,\leavevmode\nobreak\ \leavevmode\nobreak\ \phi_{t}(u,u)=\langle\mathcal{H}(t)u|u\rangle_{\mathcal{X}}\geq\gamma\|u\|^{2}_{\mathcal{X}^{1/2}}-\kappa\|u\|^{2}_{\mathcal{X}}\leavevmode\nobreak\ . (A.1)

Then, for any u0𝒳1/2u_{0}\in\mathcal{X}^{1/2}, there is a unique solution uu belonging to 𝒞0([0,T],𝒳1/2)𝒞1([0,T],𝒳1/2)\mathcal{C}^{0}([0,T],\mathcal{X}^{1/2})\cap\mathcal{C}^{1}([0,T],\mathcal{X}^{-1/2}) of the equation

itu(t)=(t)u(t)u(0)=u0.i\partial_{t}u(t)=\mathcal{H}(t)u(t)\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ u(0)=u_{0}\leavevmode\nobreak\ . (A.2)

Moreover, u0𝒳=u(t)𝒳\|u_{0}\|_{\mathcal{X}}=\|u(t)\|_{\mathcal{X}} for all t[0,t]t\in[0,t] and we may extend by density the flow of (A.2) on 𝒳\mathcal{X} as a unitary flow U(t,s)U(t,s) such that U(t,s)u(s)=u(t)U(t,s)u(s)=u(t) for all solutions uu of (A.2).

If in addition (t):𝒳1/2𝒳1/2\mathcal{H}(t):\mathcal{X}^{1/2}\rightarrow\mathcal{X}^{-1/2} is of class 𝒞2\mathcal{C}^{2} and u0D((0))u_{0}\in D(\mathcal{H}(0)), then u(t)u(t) belongs to D((t))D(\mathcal{H}(t)) for all t[0,T]t\in[0,T] and uu is of class 𝒞1([0,T],𝒳)\mathcal{C}^{1}([0,T],\mathcal{X}).

A.2 The derivative of the determinant

We recall the following standard result.

Proposition A.2.

Let II\subset\mathbb{R} be an interval of times and let N1N\geq 1. Let A(t)A(t) be a family of N×NN\times N complex matrices which is differentiable with respect to the parameter tIt\in I. If A(t)A(t) is invertible for every tIt\in I, then

tdet(A(t))=det(A(t))Tr(A(t)1tA(t)).\partial_{t}\det(A(t))=\det(A(t))\operatorname{Tr}\big{(}A(t)^{-1}\partial_{t}A(t)\big{)}.

More generally, we have tdet(A(t))=Tr(com(A(t))ttA(t))\partial_{t}\det(A(t))=\operatorname{Tr}(\text{\rm com}(A(t))^{t}\partial_{t}A(t)) where com(A)\text{\rm com}(A) is the comatrix of AA.

Proof: Without loss of generality, let us consider the derivative at time t=0t=0 and assume N2N\geq 2 (since N=1N=1 is a trivial case). First assume that A(0)=IA(0)=I, where II is the identity matrix. Then,

t=0det(A(t))\displaystyle\partial_{t=0}\det(A(t)) =limt0det[I+tA(0)+o(t)]det(I)t\displaystyle=\lim_{t\rightarrow 0}\frac{\det\big{[}I+tA^{\prime}(0)+o(t)\big{]}-\det(I)}{t}
=limt0Tr[I+tA(0)]+o(t)Tr(I)t=Tr(A(0)).\displaystyle=\lim_{t\rightarrow 0}\frac{\operatorname{Tr}\big{[}I+tA^{\prime}(0)\big{]}+o(t)-\operatorname{Tr}(I)}{t}=\operatorname{Tr}(A^{\prime}(0)).

In the case where A(t)A(t) is invertible at t=0t=0, we write

det(A(t))=det(A(0))det(A(0)1A(t))\det(A(t))=\det(A(0))\det(A(0)^{-1}A(t))

and apply the above computation to A~h(t)=A(0)1A(t)\tilde{A}_{h}(t)=A(0)^{-1}A(t).

For any invertible AA, we have com(A)t=det(A)A1\text{\rm com}(A)^{t}=\det(A)A^{-1}. Thus, we obtain the last statement by extending the formula tdet(A(t))=Tr(com(A(t))ttA(t))\partial_{t}\det(A(t))=\operatorname{Tr}(\text{\rm com}(A(t))^{t}\partial_{t}A(t)) by a density argument. \square

A.3 Right-inverse of the divergence

Let k1k\geq 1, let α(0,1)\alpha\in(0,1) and let Ω0\Omega_{0} be a 𝒞k+1,α\mathcal{C}^{k+1,\alpha}-domain of N\mathbb{R}^{N}. We define

X0k,α={u𝒞k,α(Ω¯0,N),u=0 on Ω0},X^{k,\alpha}_{0}=\{u\in\mathcal{C}^{k,\alpha}(\overline{\Omega}_{0},\mathbb{R}^{N})\,,\,u=0\text{ on }\partial\Omega_{0}\},
Ymk1,α={v𝒞k1,α(Ω¯0,N),Ω0v=0}Y^{k-1,\alpha}_{m}=\Big{\{}v\in\mathcal{C}^{k-1,\alpha}(\overline{\Omega}_{0},\mathbb{R}^{N})\,,\,\int_{\Omega_{0}}v=0\,\Big{\}}

and

L:uX0k,αdivuYmk1,α.L\leavevmode\nobreak\ :\leavevmode\nobreak\ u\in X^{k,\alpha}_{0}\leavevmode\nobreak\ \longmapsto\leavevmode\nobreak\ \operatorname{div}u\in Y^{k-1,\alpha}_{m}\leavevmode\nobreak\ .

Notice that LL is well defined since Ω0divu=0\int_{\Omega_{0}}\operatorname{div}u=0 if uΩ00u_{\partial_{\Omega_{0}}}\equiv 0. It is shown in [16, Theorem 30] that the operator LL admits a bounded linear right-inverse

L1:Ymk1,αX0k,αL^{-1}:Y^{k-1,\alpha}_{m}\rightarrow X^{k,\alpha}_{0}

that is LL1=idLL^{-1}={\text{\rm id}} and there exists K>0K>0 such that L1v𝒞k,αKv𝒞k1,α\|L^{-1}v\|_{\mathcal{C}^{k,\alpha}}\leq K\|v\|_{\mathcal{C}^{k-1,\alpha}}.

This kind of result is classical, in particular in the framework of Sobolev spaces. In addition to [16] see also [12, 13, 15].

A.4 Fixed point theorem with parameter

Even though the Banach fixed point theorem is long-established, in this paper we need its extension to the case where the contraction depends on a parameter. Of course, this extension is also very classical. We briefly recall it for sake of completeness in order to detail the problem of the regularity.

Theorem A.3.

Let UU be an open subset of a Banach space XX and let VV be an open subset of a Banach space Λ\Lambda. Let FUF\subset U be a closed subset of XX and let

Φ:(x,λ)U×VΦ(x,λ)X.\Phi\leavevmode\nobreak\ :\leavevmode\nobreak\ (x,\lambda)\in U\times V\leavevmode\nobreak\ \longmapsto\leavevmode\nobreak\ \Phi(x,\lambda)\in X\leavevmode\nobreak\ .

Assume that

  1. (i)

    For all λ\lambda, Φ(,λ)\Phi(\cdot,\lambda) maps FF into FF.

  2. (ii)

    The maps Φ(,λ)\Phi(\cdot,\lambda) are uniformly contracting in the sense that there exists k[0,1)k\in[0,1) such that

    (x,y,λ)U×U×V,Φ(x,λ)Φ(y,λ)XkxyX.\forall(x,y,\lambda)\in U\times U\times V\leavevmode\nobreak\ ,\leavevmode\nobreak\ \leavevmode\nobreak\ \|\Phi(x,\lambda)-\Phi(y,\lambda)\|_{X}\,\leq\,k\|x-y\|_{X}\leavevmode\nobreak\ .

Then, for all λV\lambda\in V, there exists a unique solution x(λ)x(\lambda) of x=Φ(x,λ)x=\Phi(x,\lambda) in FF. Moreover, if Φ\Phi is of class 𝒞k(U×V,X)\mathcal{C}^{k}(U\times V,X) with kk\in\mathbb{N}, then x(λ)x(\lambda) is also of class 𝒞k(V,F)\mathcal{C}^{k}(V,F) (the derivatives being understood in the Fréchet sense).

Proof: The existence and uniqueness of x(λ)x(\lambda) correspond of course to the classical Banach fixed point theorem. Assume that Φ\Phi is continuous. Then, we write

x(λ)x(λ)X\displaystyle\|x(\lambda)-x(\lambda^{\prime})\|_{X} =Φ(x(λ),λ)Φ(x(λ),λ)X\displaystyle=\|\Phi(x(\lambda),\lambda)-\Phi(x(\lambda^{\prime}),\lambda^{\prime})\|_{X}
Φ(x(λ),λ)Φ(x(λ),λ)X+Φ(x(λ),λ)Φ(x(λ),λ)X\displaystyle\leq\|\Phi(x(\lambda),\lambda)-\Phi(x(\lambda),\lambda^{\prime})\|_{X}+\|\Phi(x(\lambda),\lambda^{\prime})-\Phi(x(\lambda^{\prime}),\lambda^{\prime})\|_{X}
Φ(x(λ),λ)Φ(x(λ),λ)X+kx(λ)x(λ)X.\displaystyle\leq\|\Phi(x(\lambda),\lambda)-\Phi(x(\lambda),\lambda^{\prime})\|_{X}+k\|x(\lambda)-x(\lambda^{\prime})\|_{X}.

Since k<1k<1 and λΦ(x(λ),λ)\lambda^{\prime}\mapsto\Phi(x(\lambda),\lambda^{\prime}) is continuous, we obtain the continuity of λx(λ)\lambda\mapsto x(\lambda). If Φ\Phi is of class 𝒞k\mathcal{C}^{k} with k1k\geq 1, then we apply the implicit function theorem to the equation Ψ(x,λ)=0\Psi(x,\lambda)=0 with Ψ(x,λ)=xΦ(x,λ)\Psi(x,\lambda)=x-\Phi(x,\lambda). Notice that, due to the contraction property, DxΦ(x,λ)(X)k\|D_{x}\Phi(x,\lambda)\|_{\mathcal{L}(X)}\leq k and thus DxΨD_{x}\Psi is invertible everywhere. \square

A.5 The flow of a vector field on a compact domain

Let d1d\geq 1, r0r\geq 0 and p1p\geq 1. Let Ω¯0\overline{\Omega}_{0} be a compact smooth domain of N\mathbb{R}^{N} and II be a compact interval of times. Let (t,y)I×Ω¯0U(t,y)N(t,y)\in I\times\overline{\Omega}_{0}\longmapsto U(t,y)\in\mathbb{R}^{N} a vector field which is of class 𝒞r\mathcal{C}^{r} in time and 𝒞p\mathcal{C}^{p} in space, meaning that all derivatives trypU\partial_{t}^{r^{\prime}}\partial_{y}^{p^{\prime}}U exist and they are continuous for all rrr^{\prime}\leq r and ppp^{\prime}\leq p. We also assume that U(t,y)=0U(t,y)=0 on Ω0\partial\Omega_{0}.

We define tψ(t,x)t\mapsto\psi(t,x) as the flow corresponding to the ODE

ψ(t0,x)=x and tψ(t,x)=U(t,ψ(t,x))tI.\psi(t_{0},x)=x\leavevmode\nobreak\ \leavevmode\nobreak\ \text{ and }\leavevmode\nobreak\ \leavevmode\nobreak\ \partial_{t}\psi(t,x)=U(t,\psi(t,x))\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ t\in I\leavevmode\nobreak\ . (A.3)

Notice that ψ\psi is locally well defined because (t,y)U(t,y)(t,y)\mapsto U(t,y) is at least lipschitzian in space and continuous in time. Moreover, the trajectories tIψ(t,x)Ω¯0t\in I\mapsto\psi(t,x)\in\overline{\Omega}_{0} are globally defined because U(t,y)=0U(t,y)=0 on the boundary of Ω0\Omega_{0}, providing a barrier of equilibrium points.

The purpose of this appendix is to show the following regularity result. It is of course a well known property. However, the uniform bounds are often not stated explicitly and that is why we quickly recall here the arguments to obtain them.

Proposition A.4.

Let r0r\geq 0 and p1p\geq 1. If UU is of class 𝒞r\mathcal{C}^{r} in time and 𝒞p\mathcal{C}^{p} in space, then the flow ψ\psi defined by (A.3) is of class 𝒞r+1\mathcal{C}^{r+1} in time and 𝒞p\mathcal{C}^{p} in space. Moreover, for all M>0M>0, there exist C(M)>0C(M)>0 and λ(M)>0\lambda(M)>0 such that, if UU satisfies

tI,U(t,)𝒞pM,\forall t\in I\leavevmode\nobreak\ ,\leavevmode\nobreak\ \leavevmode\nobreak\ \|U(t,\cdot)\|_{\mathcal{C}^{p}}\leavevmode\nobreak\ \leq\leavevmode\nobreak\ M,

then

tI,ψ(t,)𝒞pC(M)eλ(M)|tt0|.\forall t\in I\leavevmode\nobreak\ ,\leavevmode\nobreak\ \leavevmode\nobreak\ \|\psi(t,\cdot)\|_{\mathcal{C}^{p}}\leavevmode\nobreak\ \leq\leavevmode\nobreak\ C(M)e^{\lambda(M)|t-t_{0}|}\leavevmode\nobreak\ .

Proof: The fact that the 𝒞0\mathcal{C}^{0}-bound on UU yields a 𝒞0\mathcal{C}^{0}-bound on ψ\psi simply comes from (A.3). If U(t,y)U(t,y) is of class 𝒞1\mathcal{C}^{1} with respect to yy, it is well know that ψ(t,y)\psi(t,y) is a of class 𝒞1\mathcal{C}^{1} with respect to yy and the derivatives solve the ODE

t(yiψ(t,y))=DyU(t,ψ(t,y)).yiψ(t,y),\partial_{t}(\partial_{y_{i}}\psi(t,y))=D_{y}U(t,\psi(t,y)).\partial_{y_{i}}\psi(t,y)\leavevmode\nobreak\ , (A.4)

see for example [27]. We have yiψ(t=0)0\partial_{y_{i}}\psi(t=0)\equiv 0, thus (A.4) and Grönwall’s inequality ensures the bound on yiψ\partial_{y_{i}}\psi.

If UU is of class 𝒞2\mathcal{C}^{2} in yy, the above arguments show that yDyU(t,ψ(t,y))y\mapsto D_{y}U(t,\psi(t,y)) is of class 𝒞1\mathcal{C}^{1} and we apply again the procedure to (A.4). We obtain that ψ(t,y)\psi(t,y) is a of class 𝒞2\mathcal{C}^{2} with respect to yy and the derivatives solve the ODE

t(yiyj2ψ(t,y))=DyU(t,ψ(t,y)).yiyj2ψ(t,y)+Dyy2U(t,ψ(t,y)).(yiψ(t,y),yjψ(t,y)),\partial_{t}(\partial^{2}_{y_{i}y_{j}}\psi(t,y))=D_{y}U(t,\psi(t,y)).\partial^{2}_{y_{i}y_{j}}\psi(t,y)+D^{2}_{yy}U(t,\psi(t,y)).(\partial_{y_{i}}\psi(t,y),\partial_{y_{j}}\psi(t,y))\leavevmode\nobreak\ ,

where yiyj2ψ(t,y)\partial^{2}_{y_{i}y_{j}}\psi(t,y) is the unknown. Since we already have bounds on ψ\psi and its first derivatives, again, Grönwall’s inequality yields the bound on yiyj2ψ\partial^{2}_{y_{i}y_{j}}\psi.

By applying the argument as many times as needed, we obtain the uniform bounds for all the wanted derivatives. We also proceed in the same way to obtain the regularity with respect to the time tt. \square

A.6 Globalization of local diffeomorphisms

In this appendix, we consider a 𝒞1\mathcal{C}^{1}-function φ\varphi for a domain Ω0\Omega_{0} into itself such that φ\varphi is a local diffeomorphism and φ|Ω=id\varphi_{|\partial\Omega}={\text{\rm id}}. We would like to obtain that φ\varphi is in fact a global diffeomorphism from Ω0\Omega_{0} into itself. This extension needs topological arguments contained in the article [41] of Meisters and Olech.

Theorem 1 of [41] applied to the ball of N\mathbb{R}^{N} writes as follows.

Theorem A.5.

(Meisters-Olech, 1963)
Let BRB_{R} be the open ball of center 0 and radius R>0R>0 of N\mathbb{R}^{N} and let B¯R\overline{B}_{R} the closed ball. Let ff be a continuous mapping of B¯R\overline{B}_{R} into itself which is locally one-to-one on B¯RZ\overline{B}_{R}\setminus Z, where ZBRZ\cap B_{R} is discrete and ZZ does not cover the whole boundary BR\partial B_{R}. If ff is one-to-one from BR\partial B_{R} into itself, then ff is an homeomorphism of B¯R\overline{B}_{R} onto itself.

In fact, the original result of [41] includes different domains than the balls. Nevertheless, if we consider any smooth domain, then it has to be diffeomorphic to a ball (typically, annulus are excluded). To consider more general domains, we assume that ff is the identity at the boundary.

Theorem A.6.

Let ΩN\Omega\subset\mathbb{R}^{N} be a bounded open domain. Let ff be a continuous mapping of Ω¯\overline{\Omega} into itself which is locally one-to-one on Ω¯Z\overline{\Omega}\setminus Z, where ZZ is a finite set. Assume moreover that ff is the identity on Ω\partial\Omega. Then, ff is an homeomorphism of Ω¯\overline{\Omega} onto itself.

Proof: For RR large enough, Ω¯\overline{\Omega} is included inside the ball BRB_{R}. We extend continuously ff to a function f~\tilde{f} by setting f~=id\tilde{f}={\text{\rm id}} on B¯RΩ\overline{B}_{R}\setminus\Omega. Notice that ff maps Ω¯\overline{\Omega} into itself and is locally one-to-one at all the points of the boundary, except maybe at a finite number of them. This yields that f~\tilde{f} is locally one-to-one at all these points since the extension maps the outside of Ω¯\overline{\Omega} into itself. We apply Theorem A.5 to f~\tilde{f} and obtain that f~\tilde{f} is an homeomorphism of BRB_{R}. Since it is the identity outside Ω\Omega, f=f~|Ω¯f=\tilde{f}_{|\overline{\Omega}} must be an homeomorphism of Ω¯\overline{\Omega}. \square

If we consider ff of class 𝒞k\mathcal{C}^{k} and DfDf its jacobian matrix, then we may check the local one-to-one property by assuming that det(Df)\det(Df) only vanishes at a finite number of points. More importantly, if det(Df)\det(Df) never vanishes, then ff is a diffeomorphism.

Corollary A.7.

Let k1k\geq 1, let ΩN\Omega\subset\mathbb{R}^{N} be a bounded open domain of class 𝒞k\mathcal{C}^{k} and let f𝒞k(Ω¯,Ω¯)f\in\mathcal{C}^{k}(\overline{\Omega},\overline{\Omega}). Assume that det(Df)\det(Df) does not vanish on Ω¯\overline{\Omega} and that ff is the identity on Ω\partial\Omega. Then, ff is a 𝒞k\mathcal{C}^{k}-diffeomorphism of Ω¯\overline{\Omega} onto itself.

We could also be interested in the following other consequence.

Corollary A.8.

Let k1k\geq 1, let ΩN\Omega\subset\mathbb{R}^{N} be a bounded open domain of class 𝒞k\mathcal{C}^{k} and let ff be a 𝒞k\mathcal{C}^{k}-diffeomorphism from Ω¯\overline{\Omega} onto itself. Assume that ff is the identity on Ω\partial\Omega. Then, for all ε>0\varepsilon>0, there exists η>0\eta>0 such that, for all functions g𝒞k(Ω¯,Ω¯)g\in\mathcal{C}^{k}(\overline{\Omega},\overline{\Omega}) with gΩ=idg_{\partial\Omega}={\text{\rm id}} and fg𝒞kη\|f-g\|_{\mathcal{C}^{k}}\leq\eta, gg is also a 𝒞k\mathcal{C}^{k}-diffeomorphism of Ω¯\overline{\Omega} onto itself and f1g1𝒞kε\|f^{-1}-g^{-1}\|_{\mathcal{C}^{k}}\leq\varepsilon.

Proof: We simply notice that Ω¯\overline{\Omega} is compact and so |det(Df)|α>0|\det(Df)|\geq\alpha>0 for some uniform positive α\alpha. Thus, for gg which is 𝒞1\mathcal{C}^{1}-close to ff, DgDg is still invertible everywhere and gg is a 𝒞k\mathcal{C}^{k}-diffeomorphism due to Corollary A.7. \square

References

  • [1] H. Abdallah, A Varadhan type estimate on manifolds with time-dependent metrics and constant volume, Journal de Mathématiques Pures et Appliquées no{}^{\text{o}}99 (2013), pp.409–418.
  • [2] J. E. Avron, R. Seiler and L. G. Yaffe, Adiabatic theorems and applications to the quantum hall effect, Communications in Mathematical Physics no{}^{\text{o}}110 (1987), pp. 33–49.
  • [3] Y. Band, B. Malomed and M. Trippenbach, Adiabaticity in nonlinear quantum dynamics: Bose-Einstein condensate in a time varying box, Physical Review A no{}^{\text{o}}65 (2002), p. 033607.
  • [4] G. Bande and D. Kotschick, Moser stability for locally conformally symplectic structures, Proceedings of the American Mathematical Society no{}^{\text{o}}137 (2009), pp. 2419–2424.
  • [5] K. Beauchard, Controllablity of a quantum particle in a 1D variable domain, ESAIM, Control, Optimization and Calculus of Variations no{}^{\text{o}}14 (2008), pp. 105–147.
  • [6] K. Beauchard and J. M. Coron, Controllability of a quantum particle in a moving potential well, J. of Functional Analysis no{}^{\text{o}}232 (2006), pp. 328–389.
  • [7] K. Beauchard, H. Lange and H. Teismann, Local exact controllability of a 1D Bose-Einstein condensate in a time-varying box, SIAM Journal on Control and Optimization no{}^{\text{o}}53 (2015), pp. 2781–2818.
  • [8] M.L. Bernardi, G. Bonfanti and F. Luterotti, Abstract Schrödinger-type differential equations with variable domain, Journal of Mathematical Analysis and Applications no{}^{\text{o}}211 (1997), pp. 84–105.
  • [9] V. Bisognin, C. Buriol, M.V. Ferreira, M. Sepúlveda and O. Vera, Asymptotic behaviour for a nonlinear Schrödinger equation in domains with moving boundaries, Acta Applicandae Mathematicae no{}^{\text{o}}125 (2013), pp. 159–172.
  • [10] V. Bonnaillie, On the fundamental state energy for a Schrödinger operator with magnetic field in domains with corners, Asymptotic Analysis no{}^{\text{o}}41 (2005), pp. 215–258.
  • [11] F. Bornemann, Homogenization in time of singularly perturbed mechanical systems. Lecture Notes in Mathematics no{}^{\text{o}}1687. Springer-Verlag, Berlin, 1998.
  • [12] J. Bourgain and H. Brezis, On the equation divY=f\operatorname{div}Y=f and application to control of phases, Journal of the American Mathematical Society no{}^{\text{o}}16 (2003), pp. 393–426.
  • [13] P. Bousquet and G. Csató, The equation divu+a,u=f\operatorname{div}u+\langle a,u\rangle=f, Communications in Contemporary Mathematics (2019).
  • [14] G. Csató, B. Dacorogna and O. Kneuss, The pullback equation for differential forms. Progress in Nonlinear Differential Equations and Their Applications no{}^{\text{o}}83. Birkhäuser, New-York, 2012.
  • [15] B. Dacorogna, The pullback equation. Paolo Marcellini (ed.) et al. Vector-valued partial differential equations and applications, Cetraro, Italy, July 8-12, 2013. Lecture Notes in Mathematics no{}^{\text{o}}2179, pp. 1–72. CIME Foundation Subseries, Springer, 2017.
  • [16] B. Dacorogna and J. Moser, On a partial differential equation involving the Jacobian determinant, Annales de l’Institut Henri Poincaré, Analyse Non Linéaire no{}^{\text{o}}7 (1990), pp. 1–26.
  • [17] S. Di Martino and P. Facchi, Quantum systems with time-dependent boundaries, International Journal of Geometric Methods in Modern Physics no{}^{\text{o}}12 (2015).
  • [18] S.W. Doescher and M.H. Rice, Infinite Square-Well Potential with a Moving Wall, American Journal of Physics no{}^{\text{o}}37 (1969), p. 1246.
  • [19] P. Exner and H. Kovařík, Quantum waveguides. Theoretical and Mathematical Physics. Springer International Publishing, Switzerland, 2015.
  • [20] S. Fournais and B. Helffer, Spectral methods in surface superconductivity. Progress in Nonlinear Differential Equations and their Applications no{}^{\text{o}}77. Birkhäuser Boston Inc., Boston, MA, 2010.
  • [21] S. Fournais and B. Helffer, Accurate eigenvalue asymptotics for the magnetic Neumann Laplacian, Université de Grenoble. Annales de l’Institut Fourier no{}^{\text{o}}56 (2006), pp. 1–67.
  • [22] J. Fröhlich and U. M. Studer, Gauge Invariance and Current Algebra in Non-Relativistic Many-Body Theory, Reviews of Modern Physics no{}^{\text{o}}65 (1993), pp. 733–802.
  • [23] J. Fröhlich, U. M. Studer and E. Thiran, Quantum Theory of Large Systems of Non-Relativistic Matter, arXiv:cond-mat/9508062 (1995).
  • [24] L.M. Garrido, Generalized adiabatic invariance, Journal of Mathematical Physics no{}^{\text{o}}5 (1964), pp. 335–362.
  • [25] H. Geiges, Contact geometry. Chapter 5 of Handbook of Differential Geometry, vol. 2. Editors J.E. Franki, J.E. Dillen and L.C.A. Verstraelen. North-Holland, 2006.
  • [26] S. Haag, J. Lampart and S. Teufel, Generalised quantum waveguides, Annales Henri Poincaré. A Journal of Theoretical and Mathematical Physics no{}^{\text{o}}16 (2015), pp. 2535–2568.
  • [27] J.K. Hale, Ordinary differential equations. Pure and Applied Mathematics, Vol. 21. Orig. publ. by Wiley-Interscience. Huntington, New York: Robert E. Krieger Publishing Company, 1980.
  • [28] B. Helffer, On spectral theory for Schrödinger operators with magnetic potentials. Spectral and scattering theory and applications. Adv. Stud. Pure Math. no{}^{\text{o}}23. Math. Soc. Japan, Tokyo, 1994.
  • [29] B. Helffer, M. Hoffmann-Ostenhof, T. Hoffmann-Ostenhof and M.P. Owen, Nodal sets for groundstates of Schrödinger operators with zero magnetic field in non-simply connected domains, Communications in Mathematical Physics no{}^{\text{o}}202 (1999), pp. 629–649.
  • [30] B. Helffer and Y.A. Kordyukov, Eigenvalue estimates for a three-dimensional magnetic Schrödinger operator, Asymptotic Analysis no{}^{\text{o}}82 (2013), pp. 65–89.
  • [31] D. Henry, Perturbation of the boundary in boundary-value problems of partial differential equations. With editorial assistance from J. Hale and A.L. Pereira. London Mathematical Society Lecture Note Series no{}^{\text{o}}318. Cambridge University Press, 2005.
  • [32] A. Joye and C.E. Pfister, Quantum adiabatic evolution, in: On Three Levels. Eds. M. Fannes, C. Maes, A. Verbeure, pp. 139–148. Plenum, New York, 1994.
  • [33] T. Kato, Perturbation Theory for Linear Operators. Principles of Mathematical Sciences, Reprint of the 1980 edition. Springer-Verlag, Berlin, 2003.
  • [34] C.E. Kenig, G. Ponce and L. Vega, Oscillatory Integrals and Regularity of Dispersive Equations, Indiana University Mathematics Journal no{}^{\text{o}}40 (1991), pp. 33–69.
  • [35] C.E. Kenig, G. Ponce and L. Vega, Small solutions to nonlinear Schrödinger equations, Annales de l’Institut Henri Poincaré, Analyse Non Linéaire no{}^{\text{o}}10 (1993), pp. 255–288.
  • [36] C.E. Kenig, G. Ponce and L. Vega, Smoothing effects and local existence theory for the generalized nonlinear Schrödinger equations, Inventiones Mathematicae no{}^{\text{o}}134 (1998), pp. 489–545.
  • [37] J. Kisyński, Sur les opérateurs de Green des problèmes de Cauchy abstraits, Studia Mathematica no{}^{\text{o}}23 (1963), pp. 285–328.
  • [38] E. Knobloch and R. Krechetnikov, Problems on time-varying domains: formulation, dynamics, and challenges, Acta Applicandae Mathematicae no{}^{\text{o}}137 (2015), pp. 123–157.
  • [39] F. Linares and G. Ponce, Introduction to nonlinear dispersive equations. Universitext. New York, NY: Springer, 2015.
  • [40] A.J. Makowski and P. Pepłowski, On the behaviour of quantum systems with time-dependent boundary conditions, Physics Letters A no{}^{\text{o}}163 (1992), pp. 143–151.
  • [41] G.H. Meisters and C. Olech, Locally One to One Mappings and a Classical Theorem on Schlicht Functions, Duke Mathematical Journal no{}^{\text{o}}30 (1963), pp. 63–80.
  • [42] J. Moser, On the volume elements on a manifold, Transaction of the American Mathematical Society no{}^{\text{o}}120 (1965), pp. 286–294.
  • [43] I. Moyano, Controllability of a 2D quantum particle in a time-varying disc with radial data, Journal of Mathematical Analysis and Applications no{}^{\text{o}}455 (2017), pp. 1323–1350.
  • [44] F. Murat and J. Simon, Étude de problèmes d’optimal design. Proceedings of the 7th IFIP Conference, Nice, pp. 54–62. Lect. Notes Comput. Sci. no{}^{\text{o}}41, 1976.
  • [45] G. Nenciu. On the adiabatic theorem of quantum mechanics, Journal of Physics A: Mathematical and General no{}^{\text{o}}13 (1980), pp. 15–18.
  • [46] X.B. Pan and K.H. Kwek, Schrödinger operators with non-degenerately vanishing magnetic fields in bounded domains, Transactions of the American Mathematical Society no{}^{\text{o}}354 (2002), pp. 4201–4227.
  • [47] D.N. Pinder, The contracting square quantum well, American Journal of Physics no{}^{\text{o}}58 (1990), p. 54.
  • [48] T. Rivière and D. Ye, A resolution of the prescribed volume form equation, Comptes Rendus de l’Académie des Sciences de Paris, Série I, no{}^{\text{o}}319 (1994), pp. 25–28.
  • [49] T. Rivière and D. Ye, Resolutions of the prescribed volume form equation, NoDEA Nonlinear Differential Equation and Applications no{}^{\text{o}}3 (1996), pp. 323–369.
  • [50] P. Rouchon, Control of a quantum particle in a moving potential well, 2nd IFAC Workshop on Lagrangian and Hamiltonian Methods for Nonlinear Control (2003), Seville.
  • [51] H. Tanabe, Equations of evolution. Translated from the Japanese by N. Mugibayashi and H. Haneda. Monographs and Studies in Mathematics no{}^{\text{o}}6. Pitman (Advanced Publishing Program), Boston, Mass.-London, 1979.
  • [52] S. Teufel, Adiabatic perturbation theory in quantum dynamics. Lecture Notes in Mathematics no{}^{\text{o}}1821. Springer-Verlag, Berlin, 2003.