On the Cauchy problem for the Hartree approximation in quantum dynamics
Abstract.
We prove existence and uniqueness results for the time-dependent Hartree approximation arising in quantum dynamics. The Hartree equations of motion form a coupled system of nonlinear Schrödinger equations for the evolution of product state approximations. They are a prominent example for dimension reduction in the context of the the time-dependent Dirac–Frenkel variational principle. We handle the case of Coulomb potentials thanks to Strichartz estimates. Our main result addresses a general setting where the nonlinear coupling cannot be considered as a perturbation. The proof uses a recursive construction that is inspired by the standard approach for the Cauchy problem associated to symmetric quasilinear hyperbolic equations.
1. Introduction
We consider the time-dependent Schrödinger equation
(1.1) |
where the total Hamiltonian is given by
with and , . The potentials and are always real-valued, we will make extra regularity and decay assumptions later on . It is common wisdom that for dealing with quantum systems “a solution of the wave equation in many-dimensional space is far too complicated to be practicable” (Dirac 1930) and one aims at approximative methods that effectively reduce the space dimension. Here, we focus on initial data that decouple the space variables,
Such a form of initial data indeed suggests a dimension reduction approach. Of course, if there is no coupling (), the full solution is itself a product state , with
When the coupling is present, one seeks for approximate solutions of product form in order to reduce the initial system (1.1) in to two systems on spaces of smaller dimensions, and . In situations, where the overall configuration space has a natural decomposition of its dimension , a corresponding product ansatz of factors is sought. Here we only investigate the case , mentioning that repeated application of the binary construction yields the more general case. Applying the time-dependent Dirac–Frenkel variational principle to the manifold
yields the so-called time-dependent Hartree approximation,
where the pair solves the nonlinearly coupled system
(1.2) |
The time-dependent potentials result from the averaging process
under the assumption, made throughout this paper, that
(1.3) |
For any “reasonable” solution (at least with the regularity considered in this paper), the -norms of and , respectively, are independent of time, hence
for all in the time interval where the solution to (1.2) is well-defined; see §6.3 for a proof.
Even though the time-dependent Hartree approximation is one of the most fundamental approximations in quantum dynamics, mathematical existence and uniqueness proofs are rather scarce. Existence and uniqueness have been studied in the case where the interaction potential is of convolution type, i.e. for and with one of the subsystems moving by classical mechanics (see [7, 2, 6] for example). A related investigation has targeted the time-dependent self-consistent field system [12] with coupling potentials of Schwartz class. However, our aim here is to discuss the existence and uniqueness of solutions for system (1.2) when the potentials or need not be bounded, and cannot be considered as a perturbation of or , respectively. This framework requires a different approach. In particular, our result provides the Cauchy theory for the systems discussed in the articles [3, 4] where the accuracy of the Hartree approximation is studied in the broader context of composite quantum dynamics and scale separation.
The steps for our existence and uniqueness proof are strongly inspired by the method which is classical in the study of quasilinear hyperbolic systems, see e.g. [1]. With , we associate the iterative scheme of recursive equations
(1.4) |
with
(1.5) |
The main steps of the proof of our existence and uniqueness result are then:
-
(1)
The iterative scheme is well-defined and enjoys bounds in “large” norm, which control second order derivatives and polynomial growth of order two for some finite time horizon.
-
(2)
The solution of the scheme converges in “small” norm, which is the norm.
-
(3)
It is possible to pass to the limit in the equation, which leads to the construction of a solution that one then proves to be unique and global in time (provided the initial data is regular enough).
1.1. Outline
In the next Section 2, we recall elementary properties of the time-dependent variational principle and formally derive the Hartree equations (1.2). Then we discuss coupling potentials of Coulombic and of polynomial type in Section 3, where we also present our main result Theorem 3.11, which establishes existence and uniqueness of the solutions to the Hartree system for coupling with polynomial growth. The different steps of the proof of Theorem 3.11 are the subject of Section 4 (analysis of the iterative scheme), Section 5 (convergence in small norms) and Section 6 (passing to the limit). A sufficient condition for the growth of the coupling potential is verified in Section 7. The Appendices A and B summarize some technical arguments.
1.2. Notations
We write for . The notations , , stand for , , , respectively. We denote by , , the corresponding inner products. For , we write whenever there exists a “universal” constant (in the sense that it does not depend on time, space, or , typically) such that .
2. Variational principle
The time-dependent Hartree approximation results from the Dirac–Frenkel variational principle applied to the manifold
(2.1) |
see also [15, §II.3.1] for the analogous discussion with Hartree products of orbitals in . The reader can also refer to [14]. The principle determines an approximate solution for the time-dependent Schrödinger equation
with initial data by requiring that for all times
(2.2) |
where denotes the tangent space of at . For deriving the Hartree equations, we first have to understand the manifold and its tangent space. Note that the representation of a Hartree function is non-unique, since for any . However, we can have unique representations in the tangent space once appropriate gauge conditions are set.
Lemma 2.1 (Tangent space).
For any , ,
Any has a unique representation of the form , if we impose the gauge condition . The tangent spaces are complex linear subspaces of such that for all .
The lemma is proved in Appendix A. The following formal arguments show that, in case that the variational solution is well-defined and sufficiently regular, the norm and the energy expectation value are conserved automatically. Indeed, we differentiate with respect to time and use the variational condition (2.2) for ,
due to self-adjointness of the Hamiltonian. Similarly, using self-adjointness and the variational condition (2.2) for ,
Let us now formally derive the Hartree system (1.2). We write
with . We have
Choosing elements and evaluating (2.2), we obtain the following necessary and sufficient conditions:
-
(i)
If =0, we obtain that for all such that ,
-
(ii)
If =0, we obtain that for all such that ,
The choice of and satisfying the Hartree system (1.2) guarantees (i) and (ii).
3. Main result
We present existence and uniqueness results for the solution of the time-dependent Hartree system (1.2). In §3.1, we discuss how Strichartz estimates may be applied to Coulombic coupling. In §3.2, we give detailed assumptions on polynomial growth conditions. Then, in §3.3 we state our main result Theorem 3.11.
3.1. Coupling potentials of Coulombic form
The case of Coulomb singularities (in combination with classical nuclear dynamics) has already been addressed in [7, 2] by Schauder and Picard fixed point arguments, respectively. We briefly revisit the main result from [7], and show how it may be adapted thanks to Strichartz estimates. We suppose , and have in mind the case
(3.1) |
We consider more generally the case , for a possibly singular . We assume that the potentials and are perturbations of smooth and at most quadratic potentials:
where
and
Typically, we may consider Coulomb potentials, as
The first term on the right hand side belongs to for any , and the second term is obviously bounded. We then make the same assumption on . Denote
Then and enjoy Strichartz estimates, and (1.2) can be solved at the level, by a straightforward adaptation of [10, Corollary 4.6.5]:
Theorem 3.1.
The proof is presented shortly in Appendix B.
Remark 3.2.
The sign of in (3.1) plays no role here. Indeed, the proof relies on local in time Strichartz estimates associated to and , respectively, and the potentials and are treated as perturbations, whose sign is irrelevant in order to guarantee the above global existence result. On the other hand, Theorem 3.1 brings no information regarding the quality of the dynamics or the existence of a ground state.
Remark 3.3.
Under extra assumptions on the potentials and (no extra assumption is needed for , as it is associated to a convolution), it is possible to consider higher regularity properties. In particular, working at the level of -regularity makes it possible to show the conservation of the energy
provided that and also belong to for some . We refer to Remark B.5 for more details.
Remark 3.4.
The role of the set is to guarantee that (local in time) Strichartz estimates are available for and . The same would still be true for a larger class of potentials, including for instance Kato potentials ([20]) or potentials decaying like an inverse square ([5]). The choice of this set is made in order to simplify the presentation, and because it is delicate to keep track of all the classes of potentials for which Strichartz estimates have been proved.
3.2. Coupling potentials with polynomial growth
The core of this paper addresses the case where the coupling potential may grow polynomially. To be more concrete, we recall the example addressed in [4].
Example 3.5.
Assume and that the potentials are given by
Here, corresponds to a double well and to a harmonic bath. The coupling could be locally cubic when choosing for in a neighborhood of zero.
We emphasize that in Example 3.5, the average grows quadratically in : in terms of growth, is comparable to and cannot be considered as a perturbation as far as the Cauchy problem is concerned. This setting turns out to be very different from the one in [7, 2] (see also Theorem 3.1), and requires a different approach to be developed below.
3.2.1. Restriction to non-negative potentials
In the general case, we assume . First, the potentials and are smooth, real-valued, , , and bounded from below:
for some constants . The operators and then are self-adjoint operators. Up to changing to (which amounts to replacing by in (1.2)), and to , we may actually assume
(H1) |
as we are only interested in existence results for the Cauchy problem (1.2). Thus and are sums of a nonnegative operator (Laplacian in and , respectively) and of a nonnegative potential. We use them to measure the regularity of the solutions of the system (1.2).
3.2.2. Functional setting
For , we define the Hilbert spaces
which are the natural analogues of Sobolev spaces in the presence of (nonnegative) potentials (in view of (H1)), equipped with the norms given by
For , , we set
All along the paper, we use that in view of (H1), and .
As (1.2) is reversible, from now on we consider positive time only. We shall work with the time-dependent functional spaces
If , we set
We choose to consider integer exponents and for the sake of simplicity. We emphasize however that our approach requires ; see Section 4 for a more precise discussion on this aspect. We note that Theorem 3.11 allows .
3.2.3. Main assumptions
We assume that the coupling potential satisfies
(H2) There exist with such that for all ,
We emphasize that no condition is required concerning the above constant : for instance if and are bounded, then we may always pick so that (H2) is satisfied. For unbounded potentials, the requirement can be understood as some smallness property, in the sense that is a perturbation of . This actually corresponds to the physical framework where the system (1.2) is introduced in order to approximate the exact solution of (1.1) through the formula ; see [3] for a derivation of error estimates. We also note that the assumption implies that is -bounded with relative bound , hence by Kato–Rellich Theorem (see e.g. [18, Theorem X.12]), is self-adjoint.
Remark 3.6.
If is at most quadratic, in the sense that
then the assumption (H2) is not needed to guarantee that is self-adjoint (see the Faris–Lavine Theorem, [18, Theorem X.38]). Such a framework corresponds to the assumptions made for the error analysis in [3, 4]. For such potentials , the assumption in (H2) is needed only in order to ensure that the Hartree solutions are global in time.
We also assume some conditions on the regularity of commutators of the coupling potential with the operators and . For integers , we consider the condition:
(H3)α,β. There exist such that for all , , for all , in the Schwartz class (),
Assumption (H3)α,β is made in order to generalize the framework of Example 3.5. The subsequent proofs do not use the special form of the Hamiltonians , , so that our result extends as soon as they are self-adjoint operators and assumptions (H1), (H2), (H3)α,β are satisfied. This applies in particular for magnetic Schrödinger operators.
Remark 3.7.
It is not necessary to assume that the potential is smooth, . We only need enough regularity in order to write assumption for the and that we consider (recalling that and are self-adjoint). For example, if , then Theorem 3.11 holds with . We make this regularity assumption for simplicity, as the most important properties are those discussed in this subsection.
Remark 3.8.
We next present sufficient conditions on the potentials guaranteeing that assumptions (H2) and (H3)2,2 hold.
Lemma 3.9 (Sufficient conditions).
Let and such that . The above assumptions (H2) and (H3)2,2 are satisfied provided that the following estimates hold:
-
•
There exists such that
(3.2) -
•
There exist and independent of and such that
(3.3)
The proof of this lemma is given in Section 7.
3.3. Main result and comments
Before stating our main result we informally summarize the previous assumptions on the potentials for the case of polynomial coupling:
We have the following result on existence and uniqueness as well norm and energy conservation of the time-dependent Hartree approximation.
Theorem 3.11.
We will see that the assumption in (H2) arises in two steps of the proof of Theorem 3.11. First, to make sure that the approximating scheme (1.4) introduced below is well-defined, we invoke Kato–Rellich Theorem, to show essentially that (or more precisely, ) is -bounded with relative bound smaller than one, and that the same holds when the roles of and are swapped. Second, the assumption guarantees that the conserved energy , defined in Theorem 3.11, is a coercive functional, so the conservation of provides uniform in times a priori estimates, which in turn allow to show that the local in time solutions are actually global in time solutions.
The property means that is locally bounded on . The map is bounded on in view of the conservation of the coercive energy , but higher order norms may not be bounded as goes to infinity (recall that Theorem 3.11 requires ).
4. Analysis of the iterative scheme: existence and uniform bounds
This section is devoted to the analysis of the system (1.4). For , we denote by , the solution to the scheme (1.4) and we prove local in time uniform estimates. At this stage, we only need that , for integers .
Lemma 4.1.
The proof of this lemma relies on the fact that the control of involves terms which are linear in and quadratic in , and require the -norm of . For this reason, the Lemma holds as soon as . However in Theorem 3.11 , we require at least an regularity. The reason will appear in Section 5, as we do need uniform (in ) estimates in to show that the sequence converges in .
In Section 4.1, we address the construction of the family , which relies on a commutation lemma that we prove in Section 4.3. Section 4.2 is devoted to the proof of the uniform bound stated in Lemma 4.1.
4.1. Well-posedness of the scheme
Before entering into the proof of Lemma 4.1, let us discuss why the scheme is indeed well-defined: as solves a decoupled system of linear Schrödinger equations, it suffices to study the properties of the time-dependent potentials and . We fix arbitrary and take , and , with . For , is obviously well-defined with , and
(4.2) |
holds for . We argue by induction. If satisfies (4.2), then in view of (H2), and are well-defined. In addition, for , (H2) yields
(4.3) | ||||
for some constant whose value is irrelevant here, unlike the fact that we assume . Indeed, together with (4.2), this implies that is -bounded with relative bound at most . By Kato–Rellich Theorem (see e.g. [18, Theorem X.12]), is well-defined (see e.g. [19, Section VIII.4]), and (4.2) holds at level . Next, we prove that . Applying the operator to the first equation in (1.4), we find
(4.4) |
Since is self-adjoint, we deduce
Minkowski inequality yields, in view of (H3)1,1,
(4.5) |
We infer the existence of a universal constant such that
We deduce
(4.6) |
We have a similar estimate for :
(4.7) |
and so Gronwall lemma and the inductive assumption yields and completes the construction of the sequence .
4.2. Uniform bounds
We conclude the proof of Lemma 4.1 in analyzing the regularity of the solutions.
Proof of Lemma 4.1.
In view of the definition of the scheme and of the conservations
we need now consider and for . Let
and introduce
We distinguish two cases for the ease of presentation.
First case: . In that case, if , then estimates (4.6) and (4.7) imply
We infer that choosing sufficiently small in terms of , but independently of , implies .
Higher regularity: The control of higher order regularity is obtained by a similar recursive argument which uses an iterated commutator estimate. Let . We have
(4.8) |
The next lemma allows to control the commutators.
Lemma 4.2.
Let for some , and be integers. Suppose that (H3)k,ℓ is satisfied. For all ,
Taking the lemma for granted, (4.8) implies, since is self-adjoint,
Remark 4.3.
Lemma 4.1 holds as soon as , but this is not enough in order to conclude that the sequence converges to some solution of (1.2). Indeed, the mere boundedness in only implies the convergence of a subsequence in the weak-* topology: this is not enough to pass to the limit in (1.4), both because the subsequence need not retain consecutive indices, and because the topology considered is too large to pass to the limit in nonlinear terms. These issues are overcome by requiring in Sections 5 and 6.
4.3. Proof of Lemma 4.2
Of course, (4.5) implies the result when . Take and assume that the result holds for all . We write
We deduce from (4.5) and the recursive assumption that for , we have
By the recursive assumption
Finally, in view of (H3)k,ℓ and Minkowski inequality, we have
which concludes the proof, after arguing similarly with .
5. Convergence in small norms
The second step of the proof of Theorem 3.11 consists in passing to the limit and prove the existence of a limit to the sequence of solutions to (1.4). The main result in this section is:
Lemma 5.1.
Assume that there exist and such that
Then there exist and such that
(5.1) |
If in addition is bounded in for some integers , then .
Proof.
Consider (1.4) at steps and , respectively, and subtract the corresponding equations. We find, for ,
and energy estimates yield, for , since ,
(5.2) |
In view of (H2), the key term is estimated by
Writing , and using Cauchy-Schwarz inequality,
(5.3) |
Plugging this estimate into (5.2), we infer, thanks to Minkowski inequality,
We obtain a similar estimate by exchanging the roles of and , and so
(5.4) |
Fixing sufficiently small, the series
converges geometrically, and converges in , to some .
On the other hand, the boundedness of in implies that a subsequence is converging in the weak-* topology of . By uniqueness of limits in the sense of distributions, we infer . The same holds when is replaced by for . ∎
6. Passing to the limit in the equation
We now have all the elements in hands for proving Theorem 3.11 by showing that the limit function constructed in Lemma 5.1 is a solution to equation (1.2) with the properties stated in Theorem 3.11.
6.1. Existence of a local solution
Combining Lemmas 4.1 and 5.1, we infer that under the assumptions of Theorem 3.11, there exists such that in . By uniqueness of the limit, we also have in (and no extraction of a subsequence is needed). Resuming the estimates from the proof of Lemma 5.1, we observe that for , and ,
Passing to the limit , we obtain that for , and ,
Therefore, keeping the same notation as from Lemma 5.1,
where we have used (4.3) and the normalization (1.3). The first term on the right hand side goes to zero as , uniformly in . So does the last one in the case , since by interpolation then converges to strongly in . In the case where or is equal to , we can only claim a weak convergence,
Similarly,
and solves (1.2) for , in the sense of distributions. In view of the regularity , Duhamel’s formula,
then shows the continuity in time .
6.2. Uniqueness
At this stage, it is rather clear that uniqueness holds in , no matter how large and are. Suppose that is another solution to (1.2) for : the system satisfied by is similar to the one satisfied by , and considered in the proof of Lemma 5.1. Since , there exists such that
and repeating the computations presented in the proof of Lemma 5.1, we obtain, for any ,
Picking such that shows that for , and we infer that on by covering by finitely many intervals of length at most .
6.3. Conservations
We now address the second point in Theorem 3.11: we assume that (1.2) has a unique solution for some . This implies in particular, in view of (1.2), that and , and the multiplier techniques evoked below are justified without using regularizing argument as in e.g. [10].
For the conservation of the -norms, multiply the first equation in (1.2) by , integrate in space on , and consider the imaginary part: we readily obtain
We proceed similarly for , and the conservation of the -norms follows.
For the energy, consider the multiplier in the equation for : as evoked above, all the products are well-defined, in the worst possible case as products of two functions. Integrate in space and consider the real part: we obtain
6.4. Globalization
In view of Lemmas 4.1 and 5.1, it suffices to prove a priori estimates on , showing that this quantity is locally bounded in , to infer that for all , and then globalize the solution by the standard ODE alternative.
We use the conservation of the total energy, whose expression we develop:
Since in (H2), we infer
The conservations established above yield
This is the coercivity property announced in the introduction, showing that there exists depending only on such that
for any interval on which the solution is well-defined. Proceeding like in the proof of Lemma 4.1, we have
(6.1) |
In view of Lemma 4.2 with , and , we infer
The conservation of the -norm of implies
hence an exponential a priori control of the -norm of by Gronwall lemma. The same holds for , hence the conclusion of Theorem 3.11.
7. Proof of Lemma 3.9
We briefly explain why (3.3) implies (H3)2,2, thanks to an integration by parts, in view of (3.2). Typically, for ,
Therefore, for almost all , Cauchy-Schwarz inequality yields
Using (3.3),
We deduce the expected relation for :
For , write
where we have used the estimate . For the last term, (3.3) yields
For the first term on the last right hand side, we use an integration by parts:
By Cauchy-Schwarz inequality, the first term on the right hand side is estimated by
Invoking (3.2), and using Cauchy-Schwarz inequality again,
where we have used Young inequality for the last estimate.
To estimate , we use the self-adjointness of and write
We use the above estimate, where the roles of and have been swapped, to conclude that the first inequality in (H3)2,2 holds. The proof of the second one is similar.
Appendix A Tangent space
For completeness, we give the elementary considerations for determining the tangent spaces of the Hartree manifold, Lemma 2.1.
Proof.
We consider a curve with . Then,
which verifies the claimed representation of any tangent function as
Let us consider with . We set and . Then, satisfies
Choosing and , we obtain a representation of satisfying the claimed gauge condition. We verify that this condition implies uniqueness. We assume that with . Then, for any ,
which implies . Then, for any ,
which implies . Choosing and , we have so that . ∎
Appendix B Coulombic type coupling
We recall standard definition and results.
Definition B.1 (Admissible pairs in ).
A pair is admissible if , and
As the range allowed for is compact, we set, for a time interval,
Proposition B.2.
Let and . Denote
.
There exists such that for all interval such that ,
Denote
There exists such that for all interval such that ,
The existence of (local in time) Strichartz estimates of Proposition B.2 is the main ingredient of the proof of Theorem 3.1. Actually, as soon as such estimates are available for the operators and , then Theorem 3.1 remains valid. As mentioned in the introduction, such cases can be found in e.g. [20] or [5]. On the other hand, we emphasize that for superquadratic potentials, like in Example 3.5, Strichartz estimates suffer a loss of regularity; see [16, 21].
Remark B.3.
The case of a harmonic potential, , shows that may have eigenvalues, and explains why the above time intervals are required to have finite length.
Remark B.4.
The potential may also be time dependent, in view the original framework of [11]: is real-valued, and smooth with respect to the space variable: for (almost) all , is a map. Moreover, it is at most quadratic in space:
Under these assumptions, suitable modifications of Proposition B.2 are needed, but they do not alter the conclusion of Theorem 3.1 (see [8]). See also [20] for another class of time dependent potentials.
Proof of Theorem 3.1.
We give the main technical steps of the proof, and refer to [10] for details. By Duhamel’s formula, we write (1.2) as
Theorem 3.1 follows from a standard fixed point argument based on Strichartz estimates. For , we introduce
and the distance
where stems from Proposition B.2. Then is a complete metric space.
By using Strichartz estimates and Hölder inequality, we have:
for any . By assumption (see Theorem 3.1), we may write
and the value can obviously be the same for the three potentials, by taking the minimum between the three ’s if needed. Regarding , we write
Let be such that
Note that this exponent is the one introduced in the statement of Theorem 3.1. The assumption implies . Let be such that is admissible: implies . Hölder inequality yields
where is such that
Note that is finite, as
For the convolution term, first write
Introduce such that
(B.1) |
The assumption implies . Let be such that is admissible: . Hölder inequality yields
where is such that
Note that since , we have . In view of (B.1), Young inequality yields
where is given by .
The same inequalities obviously holds by switching and , and so for sufficiently small, leaves invariant.
Using similar estimates, again relying on Strichartz and Hölder inequalities involving the same Lebesgue exponents ( is the sum of a linear and a trilinear term in ), we infer that up to decreasing , is a contraction on , and so there exists a unique solving (1.2). The global existence of the solution for (1.2) follows from the conservation of the -norms of and , respectively.
Uniqueness of such solutions follows once again from Strichartz and Hölder inequalities involving the same Lebesgue exponents as above, like for the contraction part of the argument. The main remark consists in noticing that the above Lebesgue indices satisfy , hence , and so . ∎
Remark B.5 (-regularity).
If in Theorem 3.1, we assume in addition that
then for and (this last assumption may be removed when – the minimal assumption to work at the -level with is , see [9]), the global solution constructed in Theorem 3.1 satisfies
To see this, it suffices to resume the above proof, and check that and satisfy essentially the same estimates as in . One first has to commute the gradient with or . Typically,
where the last factor accounts for the possible lack of commutation between and , . Since is at most quadratic, is at most linear, and we obtain a closed system of estimates by considering
where we have used . We omit the details, and refer to [10] (see also [8]). As pointed in Remark 3.3, the energy
which is well defined with the above regularity, is independent of time. Formally, this can be seen by multiplying the first equation in (1.2) by , the second by , integrating in space, considering the real part, and summing the two identities. To make the argument rigorous (we may not have enough regularity to be allowed to proceed as described), one may use a regularization procedure as in [10], or rely on a clever use of the regularity provided by Strichartz estimates, as in [17].
References
- [1] S. Alinhac and P. Gérard. Pseudo-differential operators and the Nash-Moser theorem, volume 82 of Graduate Studies in Mathematics. American Mathematical Society, Providence, RI, 2007. Translated from the 1991 French original by Stephen S. Wilson.
- [2] L. Baudouin. Existence and regularity of the solution of a time dependent Hartree-Fock equation coupled with a classical nuclear dynamics. Rev. Mat. Complut., 18(2):285–314, 2005.
- [3] I. Burghardt, R. Carles, C. Fermanian Kammerer, B. Lasorne, and C. Lasser. Separation of scales: Dynamical approximations for composite quantum systems. J. Phys. A, 54(41):414002, 2021.
- [4] I. Burghardt, R. Carles, C. Fermanian Kammerer, B. Lasorne, and C. Lasser. Dynamical approximations for composite quantum systems: assessment of error estimates for a separable ansatz. J. Phys. A, 55(22):224010, 2022.
- [5] N. Burq, F. Planchon, J. G. Stalker, and A. S. Tahvildar-Zadeh. Strichartz estimates for the wave and Schrödinger equations with potentials of critical decay. Indiana Univ. Math. J., 53(6):1665–1680, 2004.
- [6] F. Cacciafesta, A.-S. de Suzzoni, and D. Noja. A Dirac field interacting with point nuclear dynamics. Math. Ann., 376(3-4):1261–1301, 2020.
- [7] E. Cancès and C. Le Bris. On the time-dependent Hartree-Fock equations coupled with a classical nuclear dynamics. Math. Models Methods Appl. Sci., 9(7):963–990, 1999.
- [8] R. Carles. Nonlinear Schrödinger equation with time dependent potential. Commun. Math. Sci., 9(4):937–964, 2011.
- [9] R. Carles. Sharp weights in the Cauchy problem for nonlinear Schrödinger equations with potential. Z. Angew. Math. Phys., 66(4):2087–2094, 2015.
- [10] T. Cazenave. Semilinear Schrödinger equations, volume 10 of Courant Lecture Notes in Mathematics. New York University Courant Institute of Mathematical Sciences, New York, 2003.
- [11] D. Fujiwara. Remarks on the convergence of the Feynman path integrals. Duke Math. J., 47(3):559–600, 1980.
- [12] S. Jin, C. Sparber, and Z. Zhou. On the classical limit of a time-dependent self-consistent field system: analysis and computation. Kinet. Relat. Models, 10(1):263–298, 2017.
- [13] M. Keel and T. Tao. Endpoint Strichartz estimates. Amer. J. Math., 120(5):955–980, 1998.
- [14] C. Lasser and C. Su. Various variational approximations of quantum dynamics. Journal of Mathematical Physics, 63(7):072107, 2022.
- [15] C. Lubich. From quantum to classical molecular dynamics: reduced models and numerical analysis. European Mathematical Society (EMS), Zürich, 2008.
- [16] H. Mizutani. Strichartz estimates for Schrödinger equations with variable coefficients and unbounded potentials II. Superquadratic potentials. Commun. Pure Appl. Anal., 13(6):2177–2210, 2014.
- [17] T. Ozawa. Remarks on proofs of conservation laws for nonlinear Schrödinger equations. Calc. Var. Partial Differ. Equ., 25(3):403–408, 2006.
- [18] M. Reed and B. Simon. Methods of modern mathematical physics. II. Fourier analysis, self-adjointness. Academic Press [Harcourt Brace Jovanovich Publishers], New York, 1975.
- [19] M. Reed and B. Simon. Methods of modern mathematical physics. I Functional analysis. Academic Press, Inc. [Harcourt Brace Jovanovich, Publishers], New York, second edition, 1980.
- [20] I. Rodnianski and W. Schlag. Time decay for solutions of Schrödinger equations with rough and time-dependent potentials. Invent. Math., 155(3):451–513, 2004.
- [21] K. Yajima and G. Zhang. Local smoothing property and Strichartz inequality for Schrödinger equations with potentials superquadratic at infinity. J. Differential Equations, 202(1):81–110, 2004.