Schrödinger equation in moving domains
Abstract
We consider the Schrödinger equation
() |
where is a moving domain depending on the time . 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 and a specific family of unitary maps . We show that the conjugation by provides a new equation of the form
() |
where . The Hamiltonian is a magnetic Laplacian operator of the form
where is an explicit magnetic potential depending on the deformation of the domain . The formulation ( ‣ Schrödinger equation in moving domains) enables to ensure the existence of weak and strong solutions of the initial problem ( ‣ Schrödinger equation in moving domains) on endowed with Dirichlet boundary conditions. In addition, it also indicates that the correct Neumann type boundary conditions for ( ‣ Schrödinger equation in moving domains) are not the homogeneous but the magnetic ones
even though ( ‣ Schrödinger 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 ( ‣ Schrödinger equation in moving domains) and to guarantee the well-posedness for slightly less regular deformations of .
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
(1.1) |
where is an interval of times and is a time-dependent family of bounded domains of with . 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 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 and thus the operator depend on the time. The usual method adopted in these kinds of problems consists of transforming in a bounded reference domain . Such transformation is then used in order to bring back the Schrödinger equation (1.1) in an equivalent equation in the phase space , which does not depend on the time. To this purpose, one can introduce a family of diffeomorphisms such that for each , is a diffeomorphism from onto with (see Figure 1). Assume in addition that the function is of class with respect to the time with .
In order to bring back the Schrödinger equation (1.1) in an equivalent equation in , one can introduce the pullback operator
(1.2) |
and its inverse, the pushforward operator, defined by
(1.3) |
If we compute the equation satisfied by when is solution of (1.1) at least in a formal sense, then we find that (1.1) becomes
(1.4) |
where is the Jacobian matrix of , stands for and corresponds to the scalar product in (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 . 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 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 defined by
(1.5) |
We also denote by its inverse
(1.6) |
Notice that the relation enables to transport the conservative structure through the change of variables. A direct computation, provided in Section 3.1, shows that solves (1.1) if and only if is solution of
(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
(1.8) |
with . Now, the operator appearing in the last equation is the conjugate with respect to and 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 rather than 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 does not appear, see for instance [19, 26]. In [22, 23], the transformation is used on manifolds with time-varying metrics. The authors assume preserving the volumes which yields . 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 of a moving particle of mass , velocity and charge in a magnetic field must be replaced by . 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 be an interval of times and let with be a family of domains. Assume that there exist a bounded reference domain in and a family of diffeomorphisms such that .
Then, Equation (1.8) endowed with Dirichlet boundary conditions generates a unitary flow on and we may define weak solutions of the Schrödinger equation
(1.9) |
by transporting this flow via to a unitary flow .
Assume in addition that the diffeomorphisms are of class with respect to the time and the space variable. Then, for any , the above flow defines a solution in solving (1.9) in the 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 to have any regularity. In particular, it may have corners such as a rectangle for example. Of course, all the domains have to be diffeomorphic to . Hence, we cannot create singular perturbations such as adding or removing corners and holes. Nevertheless, 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 where and is the unit outward normal of . However, these conditions cannot generate a unitary evolution, which is problematic for quantum dynamics. Simply consider the solution , whose norm depends on the volume of .
Even when the volume of is constant, if solves (1.1) with homogeneous Neumann boundary conditions, then the computation (1.12) below shows that the evolution cannot be unitary,
except when is tangent to at all the boundary points, meaning that the shape of 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
(1.10) |
If we denote by the unit outward normal of , then the last identity can be transposed in
(1.11) |
(see Remark 3.3 for further details on the computations). Even though the conditions seems complicated on , they simply write as the classical magnetic Neumann boundary conditions for the original problem in , 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 solves (1.1) at least formally, then
(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 be an interval of times and let with be a family of domains. Assume that there exist a bounded reference domain in of class and a family of diffeomorphisms such that .
Then, Equation (1.8) endowed with the magnetic Neumann boundary conditions (1.10) (or equivalently (1.11)) generates a unitary flow on and we may define weak solutions of the Schrödinger equation
(1.13) |
by transporting this flow via to a unitary flow .
Assume in addition that the diffeomorphisms are of class with respect to the time and the space variable. Then, for any satisfying the magnetic Neumann boundary condition of (1.13), the above flow defines a solution in solving (1.13) in the sense and satisfying the magnetic Neumann boundary condition.
Gauge transformation
As it is well known, the magnetic potential has a gauge invariance. In particular, for any of class
in space, we have
(1.14) |
Thus, it is possible to delete the magnetic term by the change of gauge when there exists of class such that
In such context, the well-posedness of the equations (1.9) and (1.13) can be investigated by considering
and by studying the solution of the following equation endowed with the corresponding boundary conditions
(1.15) |
The gauge transformation, not only simplifies Equation (1.8), but also yields a gain of regularity in the hypotheses on 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 anymore. In details, if we consider , then the existence of weak solutions of (1.9) and (1.13) can be guaranteed when is of class in space and in time. The existence of strong solutions, instead, holds when is of class in space and in time.
Also remark that the gauge transformation is not always possible to use. For example, if is a rotation of a square, is not curl-free and cannot be rectified due to the presence of corners at which 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 such
that the determinant of the Jacobian is independent of . In other words, when satisfies
the following identity
(1.16) |
In this case, the multiplication for the Jacobian of commutes with the spatial derivatives and then, Equations (1.7) and (1.8) can be simplified in the following expression, for ,
(1.17) |
This strategy can be used, not only to simplify the equations, but also to gain some regularity since is now constant and thus smooth. Therefore, maps into as soon as is of class in space.
For fixed, finding a diffeomorphism 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 , and with . Let and let be a connected bounded domain of class . Let an interval of times and assume that there exists a family of domains such that there exists a family of diffeomorphisms
which are of class with respect to and of class with respect to .
Then, there exists a family of diffeomorphisms from onto , with the same regularity as , and such that is constant with respect to , that is that
Even though results similar to Theorem 1.3 have already been stated, the fact that may simply be continuous with respect to the time and that has the same regularity of seems to be new. There are some simple cases where we can define explicit as in dimension 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 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 be an interval of times and let with be a family of domains. Assume that there exist a bounded reference domain in of class with and a family of diffeomorphisms such that .
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 , with 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.
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 appearing in or is replaced by which is constant in space and then smooth. To this end, we simply have to replace the first family of diffeomorphisms by the one given by Theorem 1.3.
Notice that, in Corollary 1.4, we have to assume that the reference domain is smooth. If are simply of class or , this is not a real restriction since we may choose a smooth reference domain and a not so smooth diffeomorphism . In the case where 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 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 of such that there exist a bounded
reference domain and a family of diffeomorphisms such that . Assume
that for each , the Dirichlet Laplacian operator on has a simple isolated
eigenvalue with normalized eigenfunction , associated with a spectral projector
, all three depending continuously on . Following the classical adiabatic principle, we expect that
if we start with a quantum state in close to and we deform very slowly the domain to the
shape , then the final quantum state should be close to up to a phase shift (see Figure
3). The slowness of the deformation is represented by a parameter and we consider
deformations between the times and , that is the following Schrödinger equation
(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 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
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 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 be a reference domain and let be a diffeorphism mapping onto (see Figure 1). In this section, we only work with fixed . 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 the points in and by the ones in . For any matrix , stands for and is the scalar product in with the convention
We use the pullback and pushforward operators defined by (1.2) and (1.3). We denote by the Jacobian matrix of . The basic rules of differential calculus give that
(2.1) |
and
(2.2) |
The result below is due to the chain rule which implies the following identity
(2.3) |
Proposition 2.1.
For every ,
where is the transposed matrix of .
Proof:
We apply (2.3) with in order to obtain the identity .
By duality, we obtain the corresponding property for the divergence.
Proposition 2.2.
For every vector field ,
Let and respectively be the unit outward normal and the measure on the boundary of . Let and respectively be the unit outward normal and the measure on the boundary of . For every and ,
Proof: Let be a test function and let . Applying the divergence theorem and the above formulas, we obtain
The first statement follows from the density of in . The second one instead is proved by considering the border terms appearing when we proceed as above with and . In this context, we use twice the divergence theorem and we obtain that
The last relation yields the equality of the boundary terms since the first statement is valid in the
sense.
By a direct application of the previous propositions, we obtain the operator associated with .
Proposition 2.3.
For every ,
3 Main results: The Cauchy problem
3.1 Proof of Theorem 1.1: the Dirichlet case
Let be a time interval and let be a family of moving domains. In this section, we consider a second order Hamiltonian operator of the type
(3.1) |
where
(3.2) |
and
-
•
are symmetric diffusions coefficients such that there exists satisfying
-
•
is a magnetic potential;
-
•
is an electric potential.
In this subsection, we assume that is associated with Dirichlet boundary conditions and then its domain is . Of course, the typical example of such Hamiltonian is the Dirichlet Laplacian defined on . We consider the equation
(3.3) |
We use the pullback and pushforward operators and defined by (1.5) and (1.6). We also use the notations of Section 2. Notice that is an isometry from onto . Moreover, if is a class with respect to the space, then is an isomorphism from onto . We set
(3.4) |
At least formally, if is solution of (3.3), then a direct computation shows that
(3.5) |
The first operator is a simple transport of the original operator and it is clearly self-adjoint. Both other terms from the last relation come instead from the time derivative of . Since 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
Since differs from by a multiplication, the conjugacy of a multiplicative operation by either or gives the same result. We obtain that
(3.6) |
We combine the result of (3.6) and half of the last term of (3.5) by using the chain rule
We finally obtain
(3.7) |
where 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 . Indeed, using (3.1), the equation (3.7) becomes
(3.8) |
where, using the same standard notation of (3.2),
(3.9) |
with
Notice that, in the simplest case of 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 is a bounded domain (possibly irregular). Let be an interval of times. Assume that is a family of diffeomorphisms which is of class with respect to the time and the space variable. Let be an operator of the form (3.1) with domain .
Then, Equation (3.8) generates a unitary flow on and we may define weak solutions of the Schrödinger equation (3.3) on the moving domain by transporting this flow via , that is setting .
Assume in addition that the diffeomorphisms are of class with respect to the time and the space variable. Then, for any , the above flow defines a solution in solving (3.3) in the sense.
Proof: Assume that is compact, otherwise it is sufficient to cover 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 is a bounded term in . First, the assumptions on and the regularity of the diffeomorphisms imply that defined by (3.9) is a well defined self-adjoint operator on with domain . Second, it is of class with respect to the time and, for every ,
for some and . Let . Since is an isometry from onto continuously mapping into , the above properties of are also valid for because for every
The only problem concerns the regularities. Indeed, the domain of is
and if is only with respect to the space, then is not necessarily due to the presence of the Jacobian of in the definition (1.6) of . However, if is of class , then transports the regularity in space as well as the boundary condition and
does not depend on the time. We can apply Theorem A.1 in the appendix and obtain the unitary flow . The flow on is then well defined but corresponds to solutions of (3.3) only in a formal way.
Assume finally that the diffeomorphisms are of class with respect to the time and the space variable. Then, we have no more problems with the domains and
which does not depend on the time. Since is now of class with respect to the time, we may apply the
second part of Theorem A.1 and obtain strong solutions of the equation on , which are
transported to strong solutions of the equation on .
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 defined in (3.9) and with domain . This operator corresponds to the following quadratic form in
which is the Friedrichs extension of the quadratic form defined on . Now, it is natural to consider the Friedrichs extension of defined in . This corresponds to the Neumann realization of the magnetic Laplacian defined by (3.9) and with domain
(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 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 of the equation
(3.11) |
Going back to the original moving domain , (3.11) becomes
(3.12) |
It is noteworthy that in this case, the modified magnetic potential appears in the original equation. In particular, notice that the boundary condition of (3.12) is not the natural one associated with for 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 is a bounded domain (possibly irregular). Let be an interval of times. Assume that is a family of diffeomorphisms which is of class with respect to the time and the space variable. Let be a of the form (3.1) and with domain (3.10).
Then, Equation (3.11) generates a unitary flow on and we may define weak solutions of the Schrödinger equation (3.12) on the moving domain by transporting this flow via , that is setting .
Assume in addition that the diffeomorphisms are of class with respect to the time and the space variable. Then, for any , the above flow defines a solution in solving (3.12) in the sense.
Proof: Mutatis mutandis, the proof is the same as the one of Theorem 3.1.
Remark 3.3.
In (3.11), the boundary conditions satisfied by are not explicitly stated in terms of but rather in terms of . 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
Now, . Thanks to Proposition 2.1, the fact that never vanishes yields
When corresponds to and , 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 and with and let . Let be a bounded connected open domain of . Let be an interval and let be such that for all . Then, there exists a family of diffeomorphisms of of class satisfying
(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 and we compute
Since we want to obtain a spatially constant right-hand side, the dependence in is not important. Thus,
if and only if
To obtain such a diffeomorphism , it remains to apply Theorem 4.1 with being the
above right-hand side.
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 and . Let be a bounded connected open domain. Then,
the
following statements are equivalent.
-
(i)
The function satisfies .
-
(ii)
There exists satisfying
Furthermore, if is such that then there exists a constant such that
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 to 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 fixed, Theorem 4.1 corresponds to Theorem 4.2. Our main goal is to extend the result to a time-dependent measure . In the original proof of [42], Moser uses a flow method and constructs by a smooth deformation starting at and reaching . In this proof, the smooth deformation is a linear interpolation and one of the main steps consists in solving an ODE with a non-linearity as smooth as . 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 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 with one space regularity less than . 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 and with . Let be a bounded connected open domain of for some . Let be an interval of times and let be such that for all . Then, there exists a family of diffeomorphisms of of class satisfying (4.1).
Moreover, there exist continuous functions and such that, if
then
Proof: Let . Due to theorem 4.2, there is with the required space regularity satisfying (4.1) at . Setting , we replace (4.1) by the condition
Thus, we may assume without loss of generality that .
Let be the right-inverse of the divergence introduced in Appendix A.3. Notice that is of class and hence of class . Since we get that for all and we can define
(4.2) |
We get that is well defined and it is of class in time and in space. For , we define as the flow corresponding to the ODE
(4.3) |
Notice that is locally well defined because is at least lipschitzian in space and continuous in time. Moreover, the trajectories are globally defined because on the boundary of , providing a barrier of equilibrium points. The classical regularity results for ODEs show that is of class with respect to time and 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 is a diffeomorphism for all . We set .
Using Proposition A.2, we compute for
Since we have for any function , we use the trick
Since due to (4.3), we obtain
(4.4) | ||||
(4.5) |
It remains to notice that by (4.2), that and that . Now, from (4.4), we have and then
(4.6) |
We conclude that, as required,
Finally, it remains to notice that the bounds on yield bounds on due to (4.2). The
classical bounds on the flow recalled in Proposition A.4 in the appendix provide the claimed bounds
on
and . Indeed, they satisfy (4.3) and the dual ODE where the
time is reversed.
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 . It is proved following the fixed point method of [15, 16].
Proposition 4.4.
Let and with and let . Let be a bounded connected open domain of . There exists such that, for any interval and any being such that
Then, there exists a family of diffeomorphisms of of class satisfying (4.1). Moreover, there exists a constant independent of such that
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
Notice that is the sum of monomials of degree between and with respect to the coefficients because and are the first order terms of the development of . Using the fact that is a Banach algebra (see for example [14]), there exists a constant such that, for any functions
(4.7) |
We seek for a function solving (4.1) for times close to by setting
Since we would like that and as , the identity (4.1) is equivalent to
(4.8) |
Let and the spaces introduced in Appendix A.3. First notice that, by assumption,
and thus belongs to . Since is close to , we seek for small in . By Corollary A.8, there exists small such that if for all , then is a diffeomorphism from onto itself. If this is the case, we have that
This means that we can look for small as a solution in of
(4.9) |
since we expect to belong to . We construct by a fixed point argument by applying Theorem A.3 to the function
where is the right-inverse of the divergence (see Appendix A.3) and with as above. As a consequence, is a diffeomorphism and the above computations are valid. Due to (4.7), we have
where . Using that , we also have
We choose small enough such that is a diffeomorphism and
which also implies that . Then, we choose and assume that
(4.10) |
By construction, is lipschitzian from into . The classical fixed point theorem for
contraction maps shows the existence of solving (4.9) for each . The regularity of
with respect to the time is directly given by Theorem A.3. By construction
, and since , this implies that and thus . Also remember that, by
construction, is close enough to the identity in and is the identity on
so that is a diffeomorphism (see Appendix A.6 and the
topological arguments of [41]).
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 -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 with for all . Let . We consider a regularization of , which is of class with and , satisfying
(4.11) |
Assume that we first apply the flow method of Proposition 4.3 to obtain a smooth family of diffeomorphisms such that
Now, we seek for of the form satisfying the statement of Theorem 4.1. Thus, we need to find such that
i.e.
(4.12) |
We would like to apply the fixed point method of Proposition 4.4 by choosing small enough such that
(4.13) |
with as in Proposition 4.4. However, we notice that in (4.13) appears the composition with . Fix as in Proposition 4.4 and smaller than . When we consider Proposition 4.3 with , we have the upper bound for for the derivatives of any constructed as above when in (4.11). In particular, these bounds indicate how much the composition by increases the norm. Having this in mind, we may choose small (and exponentially decreasing) such that, for any satisfying (4.11), the associated diffeomorphisms are such that (4.13) holds.
We construct the diffeomorphism as follows. We choose a regularization of the function , which is of class with and , satisfying (4.11) for the suitable . We will also need that , which is provided by first choosing satisfying (4.11) with a smaller and then set . By Proposition 4.3, there exists a family of diffeomorphisms satisfying (4.11) and of class .
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 the Dirichlet Laplacian in , i.e. and , a continuous curve such that for every is in the discrete spectrum of . We also assume that is a simple isolated eigenvalue for every , associated with the spectral projectors .
We consider on the time interval the equation (3.7) in when is a Dirichlet Laplacian. Fixed we substitute by and set to obtain
(5.1) |
where
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 , (see [11, Chapter 4] or [52]), we may wonder if it is the same for the equation (5.1) because
depends on and no spectral assumptions have been made on this family. First, we notice that, by conjugation, also belongs to the discrete spectrum of the operator in and is associated with the spectral projection . Then, for each , is a small relatively compact self-adjoint perturbation of . Thus, for all small enough, there exists a curve of simple isolated eigenvalues of , associated with spectral projectors , such that and converge uniformly when to and respectively (see [33] for further details). In this framework, even if depends on , 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)
Finally, we notice that, since converges to when goes to zero,
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 and any smooth family of vectors .
The family of translated domains is where . By explicit computations, we
obtain that and .
Since does not depend on , we do not need Moser’s trick to get (1.17). We can apply the
gauge transformation since satisfies . Then,
satisfies Equation (1.15), which becomes in our framework
(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 . For example, we consider
where satisfies Equation (5.2). Then, is solution of the equation
(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 . Clearly, the same results can be extended by considering
rotations in with .
Let and let with
and . Using the classical notation , it is straightforward to check that , ,
We obtain by direct computation or by the first line of (1.17) that satisfies the following Schrödinger equation in the rotating frame
(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
which is given by the second line of (1.17) or obtained directly from (5.4) by using the fact that is divergence free. We recover a repulsive potential corresponding to the centrifugal force.
Moving domains with diagonal diffeomorphisms
Let and let with
and . As above, we obtain
(5.5) |
We apply again the gauge transformation and then,
(5.6) |
Finally, satisfies the equation
(5.7) |
In the homothetical case where (see for instance [3, 5, 7, 38, 40, 43, 47, 50]). Equation (5.7) becomes
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
and introducing the implicitly defined function
We obtain
(5.8) |
In this simple case, we see that the general framework of this paper coincides with the previous computations introduced in dimension by Beauchard in [5]. Indeed, the transformations described in [5, Section 1.3] corresponds to application 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 . 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 -norm.
For a second application of this simple case, we consider the case of a family of cylinders
for a varying length. We would like to consider the Schrödinger equation in 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 is not constant. Theorem 1.2 shows us the correct ones. We can choose
with the cylinder of length . Then, we get by (5.5) the term and we can compute that the suitable boundary conditions are
(5.9) |
and 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 . 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 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 be a Hilbert space and let be a family of self-adjoint positive operators
on such that is independent of time . Also set
and assume that is of class with respect to . In other words, we assume that the sesquilinear form
associated with has a domain independent of the time and is of
class with respect to . Also assume that
there exist and such that,
(A.1) |
Then, for any , there is a unique solution belonging to of the equation
(A.2) |
Moreover, for all and we may extend by density the flow of (A.2) on as a unitary flow such that for all solutions of (A.2).
If in addition is of class and , then belongs to for all and is of class .
A.2 The derivative of the determinant
We recall the following standard result.
Proposition A.2.
Let be an interval of times and let . Let be a family of complex matrices which is differentiable with respect to the parameter . If is invertible for every , then
More generally, we have where is the comatrix of .
Proof: Without loss of generality, let us consider the derivative at time and assume (since is a trivial case). First assume that , where is the identity matrix. Then,
In the case where is invertible at , we write
and apply the above computation to .
For any invertible , we have . Thus, we obtain the last statement by
extending the formula by a density argument.
A.3 Right-inverse of the divergence
Let , let and let be a domain of . We define
and
Notice that is well defined since if . It is shown in [16, Theorem 30] that the operator admits a bounded linear right-inverse
that is and there exists such that .
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 be an open subset of a Banach space and let be an open subset of a Banach space . Let be a closed subset of and let
Assume that
-
(i)
For all , maps into .
-
(ii)
The maps are uniformly contracting in the sense that there exists such that
Then, for all , there exists a unique solution of in . Moreover, if is of class with , then is also of class (the derivatives being understood in the Fréchet sense).
Proof: The existence and uniqueness of correspond of course to the classical Banach fixed point theorem. Assume that is continuous. Then, we write
Since and is continuous, we obtain the continuity of
. If is of class with , then we apply the implicit function
theorem to the equation with . Notice that, due to the
contraction property, and thus is invertible everywhere.
A.5 The flow of a vector field on a compact domain
Let , and . Let be a compact smooth domain of and be a compact interval of times. Let a vector field which is of class in time and in space, meaning that all derivatives exist and they are continuous for all and . We also assume that on .
We define as the flow corresponding to the ODE
(A.3) |
Notice that is locally well defined because is at least lipschitzian in space and continuous in time. Moreover, the trajectories are globally defined because on the boundary of , 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 and . If is of class in time and in space, then the flow defined by (A.3) is of class in time and in space. Moreover, for all , there exist and such that, if satisfies
then
Proof: The fact that the -bound on yields a bound on simply comes from (A.3). If is of class with respect to , it is well know that is a of class with respect to and the derivatives solve the ODE
(A.4) |
see for example [27]. We have , thus (A.4) and Grönwall’s inequality ensures the bound on .
If is of class in , the above arguments show that is of class and we apply again the procedure to (A.4). We obtain that is a of class with respect to and the derivatives solve the ODE
where is the unknown. Since we already have bounds on and its first derivatives, again, Grönwall’s inequality yields the bound on .
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 .
A.6 Globalization of local diffeomorphisms
In this appendix, we consider a function for a domain into itself such that is a local diffeomorphism and . We would like to obtain that is in fact a global diffeomorphism from 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 writes as follows.
Theorem A.5.
(Meisters-Olech, 1963)
Let be the open ball of center and radius of and let the closed ball. Let
be a continuous mapping of into itself which is locally one-to-one on , where is discrete and does not cover the whole boundary . If is one-to-one
from into itself, then is an homeomorphism of 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 is the identity at the boundary.
Theorem A.6.
Let be a bounded open domain. Let be a continuous mapping of into itself which is locally one-to-one on , where is a finite set. Assume moreover that is the identity on . Then, is an homeomorphism of onto itself.
Proof:
For large enough, is included inside the ball . We extend continuously to a function
by setting on . Notice that maps 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 is locally one-to-one at all these points since the extension maps the outside of
into itself. We apply Theorem A.5 to and obtain that is
an homeomorphism of . Since it is the identity outside , must be an
homeomorphism of .
If we consider of class and its jacobian matrix, then we may check the local one-to-one property by assuming that only vanishes at a finite number of points. More importantly, if never vanishes, then is a diffeomorphism.
Corollary A.7.
Let , let be a bounded open domain of class and let . Assume that does not vanish on and that is the identity on . Then, is a diffeomorphism of onto itself.
We could also be interested in the following other consequence.
Corollary A.8.
Let , let be a bounded open domain of class and let be a diffeomorphism from onto itself. Assume that is the identity on . Then, for all , there exists such that, for all functions with and , is also a diffeomorphism of onto itself and .
Proof:
We simply notice that is compact and so for some uniform positive
. Thus, for which is close to , is still invertible everywhere and is a
diffeomorphism due to Corollary A.7.
References
- [1] H. Abdallah, A Varadhan type estimate on manifolds with time-dependent metrics and constant volume, Journal de Mathématiques Pures et Appliquées n99 (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 n110 (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 n65 (2002), p. 033607.
- [4] G. Bande and D. Kotschick, Moser stability for locally conformally symplectic structures, Proceedings of the American Mathematical Society n137 (2009), pp. 2419–2424.
- [5] K. Beauchard, Controllablity of a quantum particle in a 1D variable domain, ESAIM, Control, Optimization and Calculus of Variations n14 (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 n232 (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 n53 (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 n211 (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 n125 (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 n41 (2005), pp. 215–258.
- [11] F. Bornemann, Homogenization in time of singularly perturbed mechanical systems. Lecture Notes in Mathematics n1687. Springer-Verlag, Berlin, 1998.
- [12] J. Bourgain and H. Brezis, On the equation and application to control of phases, Journal of the American Mathematical Society n16 (2003), pp. 393–426.
- [13] P. Bousquet and G. Csató, The equation , 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 n83. 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 n2179, 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 n7 (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 n12 (2015).
- [18] S.W. Doescher and M.H. Rice, Infinite Square-Well Potential with a Moving Wall, American Journal of Physics n37 (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 n77. 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 n56 (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 n65 (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 n5 (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 n16 (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. n23. 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 n202 (1999), pp. 629–649.
- [30] B. Helffer and Y.A. Kordyukov, Eigenvalue estimates for a three-dimensional magnetic Schrödinger operator, Asymptotic Analysis n82 (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 n318. 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 n40 (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 n10 (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 n134 (1998), pp. 489–545.
- [37] J. Kisyński, Sur les opérateurs de Green des problèmes de Cauchy abstraits, Studia Mathematica n23 (1963), pp. 285–328.
- [38] E. Knobloch and R. Krechetnikov, Problems on time-varying domains: formulation, dynamics, and challenges, Acta Applicandae Mathematicae n137 (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 n163 (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 n30 (1963), pp. 63–80.
- [42] J. Moser, On the volume elements on a manifold, Transaction of the American Mathematical Society n120 (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 n455 (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. n41, 1976.
- [45] G. Nenciu. On the adiabatic theorem of quantum mechanics, Journal of Physics A: Mathematical and General n13 (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 n354 (2002), pp. 4201–4227.
- [47] D.N. Pinder, The contracting square quantum well, American Journal of Physics n58 (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, n319 (1994), pp. 25–28.
- [49] T. Rivière and D. Ye, Resolutions of the prescribed volume form equation, NoDEA Nonlinear Differential Equation and Applications n3 (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 n6. Pitman (Advanced Publishing Program), Boston, Mass.-London, 1979.
- [52] S. Teufel, Adiabatic perturbation theory in quantum dynamics. Lecture Notes in Mathematics n1821. Springer-Verlag, Berlin, 2003.