Paper The following article is Open access

Separation of scales: dynamical approximations for composite quantum systems*

, , , and

Published 16 September 2021 © 2021 The Author(s). Published by IOP Publishing Ltd
, , Koopman Methods in Classical and Quantum-Classical Mechanics Citation Irene Burghardt et al 2021 J. Phys. A: Math. Theor. 54 414002 DOI 10.1088/1751-8121/ac219d

1751-8121/54/41/414002

Abstract

We consider composite quantum-dynamical systems that can be partitioned into weakly interacting subsystems, similar to system–bath type situations. Using a factorized wave function ansatz, we mathematically characterize dynamical scale separation. Specifically, we investigate a coupling régime that is partially flat, i.e. slowly varying with respect to one set of variables, for example, those of the bath. Further, we study the situation where one of the sets of variables is semiclassically scaled and derive a quantum–classical formulation. In both situations, we propose two schemes of dimension reduction: one based on Taylor expansion (collocation) and the other one based on partial averaging (mean-field). We analyze the error for the wave function and for the action of observables, obtaining comparable estimates for both approaches. The present study is the first step towards a general analysis of scale separation in the context of tensorized wavefunction representations.

Export citation and abstract BibTeX RIS

Original content from this work may be used under the terms of the Creative Commons Attribution 4.0 licence. Any further distribution of this work must maintain attribution to the author(s) and the title of the work, journal citation and DOI.

1. Introduction

We consider composite quantum-dynamical systems that can be partitioned into weakly interacting subsystems, with the goal of developing effective dynamical descriptions that simplify the original, fully quantum-mechanical formulation. Typical examples are small reactive molecular fragments embedded in a large molecular bath, namely, a protein, or a solvent, all being governed dynamically by quite distinct energy and time scales. To this end, various régimes of intersystem couplings are considered, and a quantum–classical approximation is explored. A key aspect is dimension reduction at the wave function level, without referring to the conventional 'reduced dynamics' approaches that are employed in system–bath theories.

1.1. The mathematical setting

The quantum system is described by a time-dependent Schrödinger equation

Equation (1)

governed by a Hamiltonian of the form

where the coupling potential W(x, y) is a smooth function, that satisfies growth estimates guaranteeing existence and uniqueness of the solution to the Schrödinger equation (1) for a rather general set of initial data, as we shall see later in section 2. The overall set of space variables is denoted as $(x,y)\in {\mathbb{R}}^{n}\times {\mathbb{R}}^{d}$ such that the total dimension of the configuration space is n + d. The wave function depends on time t ⩾ 0 and both space variables, that is, ψ = ψ(t, x, y). We suppose that initially scales are separable, that is, we work with initial data of product form

Equation (2)

In the simple case without coupling, that is, W ≡ 0, the solution stays separated, ψ(t, x, y) = φx (t, x)φy (t, y) for all time, where

and this is an exact formula. Here, we aim at investigating the case of an actual coupling with ∂x y W ≢ 0 and look for approximate solutions of the form

where the individual components satisfy evolution equations that account for the coupling between the variables. The main motivation for such approximations is dimension reduction, since φx (t, x) and φy (t, y) depend on variables of lower dimension than the initial (x, y). Of crucial importance is the choice of the approximate Hamiltonian Happ = Hx + Hy + Wapp(t, x, y), that governs the approximate dynamics. We consider two different approximate coupling potentials: one is the time-dependent Hartree mean-field potential, the other one, computationally less demanding, is based on a brute force single-point collocation. Time-dependent Hartree methods have been known for a long time and have earned the reputation of oversimplifying the dynamics of real molecular systems. We emphasize, that our present study does not aim at rejuvenating but at deriving rigorous mathematical error estimates, which seem to be missing in the literature. Surprisingly, our error analysis provides similar estimates for both methods, the collocation and the mean-field approach. We investigate the size of the difference between the true and the approximate solution in the L2-norm,

and in Sobolev norms. We present error estimates that explicitly depend on derivatives of the coupling potential W(x, y) and on moments of the approximate solution. As an additional error measure we also consider the deviation of true and approximate expectation values $\left\langle \psi (t),A\psi (t)\right\rangle -\left\langle {\psi }_{\mathrm{app}}(t),A{\psi }_{\mathrm{app}}(t)\right\rangle $, for self-adjoint linear operators A. Roughly speaking, the estimates we obtain for observables depend on one more derivative of the coupling potential W(x, y) than the norm estimates. This means that in many situations expectation values are more accurately described than the wave function itself. Even though rigorous error estimates that quantify the decoupling of quantum subsystems in terms of flatness properties of the coupling potential W(x, y) are naturally important, our results here seem to be the first ones of their kind.

1.2. Relation with previous work

Interacting quantum systems have traditionally been formulated from the point of view of reduced dynamics theories, based on quantum master equations in a Markovian or non-Markovian setting [1]. More recently, also tensorized representations of the full quantum system have been considered, as for example by matrix product states [2, 3] or within a multiconfiguration time-dependent Hartree (MCTDH) approach [47]. Both wavefunctions (pure states) and density operators (mixed states) can be described in this framework, and wavefunction-based computations can be used to obtain density matrices [8]. In the chemical physics literature, dimension reduction for quantum systems has been proposed in the context of mean-field methods [9, 10], and the quantum–classical mean-field Ehrenfest approach [11, 12]. Also, quantum–classical formulations have been derived in Wigner phase space [13, 14] and in a quantum hydrodynamic setting [1517]. Our present mathematical formulation circumvents formal difficulties of these approaches [1820], by preserving a quantum wavefunction description for the entire system. Previous mathematical work we are aware of is concerned with rather specific coupling models, as for example the coupling of Hartree–Fock and classical equations in [21], or the time-dependent self-consistent field equations in [22], or with adiabatic approximations which rely on eigenfunctions for one part of the system, see for example [23, 24]. To the best of our knowledge, the rather general mathematical analysis of scale separation in quantum systems we are developing here is new.

1.3. Partially flat coupling

For a first approximate potential, we consider a brute-force approach, where we collocate partially at a single point, for definiteness we choose the origin, and set

In comparison, following the more conventional time-dependent Hartree approach, we set

where we perform partial and full averages of the coupling potential,

For both approximations, the brute-force and the mean-field approximation, we derive various types of estimates for the error in L2-norm. Our key finding is that both methods come with error bounds that are qualitatively the same, since they draw from either evaluations or averages of the function

Depending on whether one chooses to control the auxiliary function δW in terms of ∇x W, ∇y W or ∇x y W, the estimate requires a balancing with corresponding moments of the approximate solution. For example, proposition 1 provides L2-norm estimates of the form

where const. ∈ {1, 2, 4}, depending on whether ψapp(t) results from the brute-force or the mean-field approximation. Example 3 discusses important variants of this estimate using different ways of quantifying the flatness of the coupling potential. Proposition 2 gives analogous estimates in Sobolev norms. In addition, we analyze the deviation of the true and the approximate expectation values in a similar vein. For the expectation values, we again obtain qualitatively similar error estimates for Wbf and Wmf. The upper bounds differ from the norm bounds in so far as they involve one more derivative of the coupling potential W and low order Sobolev norms of the approximate solution, see proposition 3. Hence, from the perspective of approximation accuracy, the brute force and the mean-field approach differ only slightly. Therefore, other assessment criteria are needed for explaining the prevalence of the Hartree method in many applications, as we will discuss in section 3.5.

1.4. Dimension reduction via semiclassical analysis

In the second part of the paper we turn to a specific case of the previous general class of coupled Hamiltonians ${H}^{\varepsilon }={H}_{x}+{H}_{y}^{\varepsilon }+W(x,y)$ and consider for one part of the system a semiclassically scaled Schrödinger operator

We will discuss in section 5.1 system–bath Hamiltonians that can be recast in this semiclassical format. The initial data are still a product of the form 2, but the y-factor is chosen as

that is, ${\varphi }_{0}^{y}$ is a semiclassical wave packet with a smooth and rapidly decaying amplitude function $a\in \mathcal{S}({\mathbb{R}}^{d})$, and an arbitrary phase space centre $({q}_{0},{p}_{0})\in {\mathbb{R}}^{2d}$. We will choose a semiclassical wave packet approximation for φy (t, y) exploring two different choices for the centre (q(t), p(t)). As a first option we consider the classical trajectory

and as a second option the corresponding trajectory resulting from the averaged gradient of the potential V2,

Correspondingly, the approximative factor φx (t, x) is evolved by the partial Hamiltonian Hx + Weff with

We obtain error estimates in L2-norm, see proposition 4 and expectation values, see proposition 5. These estimates are given in terms of the semiclassical parameter ɛ and derivatives of the coupling potential. Again, both choices for the effective potential differ only slightly in approximation accuracy. Measuring the coupling strength in terms of $\eta ={\Vert}{\nabla }_{y}W{{\Vert}}_{{L}^{\infty }}$, we obtain two-parameter estimates of order $\sqrt{\varepsilon }+\eta /\sqrt{\varepsilon }$ in norm, while the ones for the expectation values are of order ɛ + η. Hence, again the accuracy of quadratic observables is higher than the one for wavefunctions.

2. Assumptions

We describe here the mathematical setting that will be ours, and discuss it in the context of system–bath Hamiltonians [1, 25]. Our Hamiltonian is of the form

Equation (3)

where the potentials V1(x), V2(y) and the coupling potential W(x, y) are all smooth functions, that satisfy growth conditions as given in assumption 1. We will denote V(x, y) = V1(x) + V2(y) + W(x, y) and abbreviate the Lebesgue spaces for the different variables x, y, and (x, y) by

The initial data ψ0(x, y) in (2) are products of functions ${\varphi }_{0}^{x}\in {L}_{x}^{2}$ and ${\varphi }_{0}^{y}\in {L}_{y}^{2}$, that are square-integrable and typically, Schwartz class, see below.

2.1. Assumptions on regularity and growth of the potentials

We choose a very classical set of assumptions on the regularity and the growth of the potential, since our focus is more on finding appropriate ways to approximate the solution in a standard framework than on treating specific situations.

Assumption 1. All the potentials that we consider are smooth, real-valued, and at most quadratic in their variables:

and, for $\alpha \in {\mathbb{N}}^{n}$, $\beta \in {\mathbb{N}}^{d}$,

We also assume that ∇y WL, but note that this condition can easily be relaxed, see example 3. All the initial date we consider are smooth and rapidly decaying, that is, Schwartz class functions:

Under the above assumption, it is well-known that Hx , Hy and H are essentially self-adjoint on ${L}^{2}({\mathbb{R}}^{N})$, with N = n, d and n + d, respectively (as a consequence of Faris–Lavine theorem, see e.g. [26, theorem 10.38]).

Example 1. Since assumption 1 involves similar properties for V1, V2 or W, we present examples for V1 only, which can readily be adapted to V2 and W. For instance, we can consider

with ωj ⩾ 0 (possibly anisotropic harmonic potential), ${\beta }_{j}\in \mathbb{R}$, ${A}_{j}\in {\mathbb{R}}^{n\times n}$ positive definite symmetric matrices (Gaussian potential), and Vper a smooth potential, periodic along some lattice in ${\mathbb{R}}^{n}$.

The assumptions on the growth and smoothness of the potentials and the regularity of the initial data call for comments.

Remark 1. Concerning the growth of V1, V2 and W, the assumption that they are at most quadratic concerns the behavior at infinity and could be relaxed, up to suitable sign assumptions. Local behavior is rather free, for example a local double well is allowed, as long as it is not too confining at infinity. We choose to stick to the at most quadratic case, since bounded second order derivatives simplify the presentation.

Remark 2. Concerning the smoothness of our potentials, most of our results still hold assuming only smoothness of W, as long as the operators Hx and Hy are essentially self-adjoint on an adequate domain included in ${L}^{2}({\mathbb{R}}^{N})$, with N = n, d. For example, V1 and V2 could both present Coulomb singularities, and the results of proposition 1 would still hold. In the semiclassical régime, we can also allow a Coulomb singularity for V1 and prove propositions 4 and 5.

Remark 3. Concerning the smoothness and the decay of the initial data, most of our results still hold, if the initial data are contained in one of the spaces ${{\Sigma}}^{k}({\mathbb{R}}^{N})$ containing functions f whose norms

Equation (4)

are bounded. Note that $\mathcal{S}({\mathbb{R}}^{N})={\cap }_{k\in \mathbb{N}}{{\Sigma}}^{k}$. For example, proposition 1 still holds for initial data in Σ1, while proposition 4 requires initial data in a semiclassically scaled Σ3 space.

2.2. System–bath Hamiltonians

An important class of coupled quantum systems are described by system–bath Hamiltonians [1, 25].

These are naturally given in the format required by (26). In the present discussion, we specify that the bath is described by a harmonic oscillator, ${V}_{\mathrm{b}}(y)=\frac{1}{2}{k}_{2}^{0}\vert y{\vert }^{2}$ (or a set of harmonic oscillators in more than one dimension) and the system–bath coupling Vsb(x, y) = W(x, y) is of cubic form, such that we obtain in the notation of (26),

where ${k}_{2}^{0} > 0$ and $\overrightarrow{\eta }\in {\mathbb{R}}^{d}$. The cubic, anharmonic coupling W(x, y) is a non-trivial case, which is employed, e.g. in the description of vibrational dephasing [27, 28] and Fermi resonances [29]. It is natural to assume smoothness and subquadratic growth for Vs(x). However, the coupling potential W(x, y) clearly fails to satisfy the growth condition of assumption 1. Moreover, it is not clear that in such a framework the total Hamiltonian H is essentially self-adjoint. On the other hand, adding a quartic confining potential,

guarantees that H is essentially self-adjoint. Indeed, Young's inequality for products yields

so choosing ${c}_{0}={k}_{4}^{0}$, we have, for the total potential V(x, y) = Vs(x) + Vb(y) + W(x, y),

for some constants C1, C2 ⩾ 0. Hence, the Faris–Lavine theorem implies that Hx , Hy and H are essentially self-adjoint. In the following, we will therefore also provide slight extensions of our estimates to accommodate this specific, but interesting type of coupling (see remark 8).

3. Partially flat coupling

In this section, we present error estimates that reflect partial flatness properties of the coupling potential W(x, y) in the sense, that quantities like ${\Vert}{\nabla }_{y}W{{\Vert}}_{{L}^{\infty }}$ or ${\Vert}{\nabla }_{x}{\nabla }_{y}W{{\Vert}}_{{L}^{\infty }}$ are small. Depending on the regularity of the initial data, the smallness of these norms could also be relaxed to the smallness of ${\Vert}{\left\langle x\right\rangle }^{-{\sigma }_{x}}{\langle y\rangle }^{-{\sigma }_{y}}{\nabla }_{y}W{{\Vert}}_{{L}^{\infty }}$ for some σx , σy ⩾ 0, see example 3. We investigate two approximation strategies, one that is based on brute-force collocation, the other one on spatial averaging. In each case, we prove that the coupling in (x, y) is negligible at leading order with respect to ∇y W. Throughout this section, ψ = ψ(t, x, y) denotes the solution to the initial value problem (1) and (2).

3.1. Brute-force approach

We consider the uncoupled system of equations

Equation (5)

In view of assumption 1, these equations have unique solutions ${\varphi }^{x}\in C(\mathbb{R};{L}_{x}^{2})$, ${\varphi }^{y}\in C(\mathbb{R};{L}_{y}^{2})$, and higher regularity is propagated, ${\varphi }^{x}\in C(\mathbb{R};{{\Sigma}}_{x}^{k})$, ${\varphi }^{y}\in C(\mathbb{R};{{\Sigma}}_{y}^{k})$, for all $k\in \mathbb{N}$, where we recall that Σk has been defined in (4). The plain product solves

This is not the right approximation, since the residual term is not small: even if W varies very little in y, then W(x, y) − W(x, 0) − W(0, y) ≈ W(x, 0) − W(x, 0) − W(0, 0) = −W(0, 0). This term is removed by considering instead

It satisfies the equation

with approximative potential Wapp(x, y) = W(x, 0) + W(0, y) − W(0, 0). The last term Σψ controls the error ψψapp, as we will see more precisely below. Saying that the coupling potential W is flat in y means that ∇y W is small, and we write

This suggests that partial flatness of W implies smallness of the approximation error.

Remark 4. For choosing another collocation point than the origin, one might use the matrix

The Taylor expansion

Equation (6)

implies for (x, y) ≈ (x0, y0) that

which corresponds to the standard normal mode expansion. Hence, choosing (x0, y0) such that the maximal singular value of M(x0, y0) is minimized, we minimize the error of the brute-force approach.

3.2. Mean-field approach

Instead of pointwise evaluations of the coupling potential, we might also use partial averages for an approximation. We consider

Equation (7)

where we have denoted

Equation (8)

Equation (9)

where we have used the fact that the L2-norms of ϕx and ϕy are independent of time, since W is real-valued. Note that (7) is the nonlinear system of equations of the time-dependent Hartree approximation. Contrary to the brute-force approach, L2 regularity does not suffice to define partial averages in general. In view of assumption 1, a fixed point argument (very similar to the proof of e.g. [30, lemma 13.10]) shows that this system has a unique solution $({\phi }^{x},{\phi }^{y})\in C(\mathbb{R};{{\Sigma}}_{x}^{1}\times {{\Sigma}}_{y}^{1})$, and higher regularity is propagated, ${\phi }^{x}\in C(\mathbb{R};{{\Sigma}}_{x}^{k})$, ${\phi }^{y}\in C(\mathbb{R};{{\Sigma}}_{y}^{k})$, for all k ⩾ 2. The approximate solution is then

with the phase given by the full average

It solves the equation

Remark 5. The correcting phase $\mathrm{exp}(\mathrm{i}{\int }_{0}^{t}\left\langle W\right\rangle \mathrm{d}s)$ seems to be crucial if we want to compute the wave function. On the other hand, since it is a purely time dependent phase factor, it does not affect the usual quadratic observables. The same applies for the phase exp(itW(0, 0)) of the brute-force approximation.

3.3. Error estimates for wave functions

We begin with an approximation result at the level of L2-norms only. For its proof, see section 4.

Proposition 1. Under assumption 1, we have the following error estimates:

Brute-force approach: for ψapp(t, x, y) = φx (t, x)φy (t, y)exp(itW(0, 0)) defined by (5),

Mean-field approach: for ${\phi }_{\mathrm{app}}(t,x,y)={\phi }^{x}(t,x){\phi }^{y}(t,y)\mathrm{exp}(\mathrm{i}{\int }_{0}^{t}\left\langle W\right\rangle \mathrm{d}s)$ defined by (7),

We see that the smallness of ${\Vert}{\nabla }_{y}W{{\Vert}}_{{L}^{\infty }}$ controls the error between the exact and the approximate solution in both approaches.

Example 2. An important class of examples consists of those where W is slowly varying in y: W(x, y) = w(x, ηy) with η ≪ 1 and w bounded, as well as its derivatives. In that case ${\Vert}{\nabla }_{y}W{{\Vert}}_{{L}^{\infty }}=\eta {\Vert}{\nabla }_{y}w{{\Vert}}_{{L}^{\infty }}$.

Example 3. In the case W(x, y) = W1(x)W2(y), the averaged potentials satisfy

with

In this special product case, the crucial source term takes the form

and proposition 1 can be augmented by the gradient-free estimate

Equation (10)

The L-norms, that provide the upper bounds in proposition 1, separate as

and it is ${\Vert}{\nabla }_{y}{W}_{2}{{\Vert}}_{{L}^{\infty }}$ that controls the estimates. Suppose we have W2(y) = η|y|2 with η small: ∇W2 is not bounded, but we can adapt the proof of proposition 1 to get

that is, the extra power of y is transferred to the ϕy term.

Remark 6. In the spirit of the last observations of example 3, in terms of $\eta {:=}{\Vert}{\left\langle x\right\rangle }^{-{\sigma }_{x}}{\langle y\rangle }^{-{\sigma }_{y}}{\nabla }_{y}W{{\Vert}}_{{L}^{\infty }}$, for some σx , σy ⩾ 0, we get in proposition 1:

See appendix B for details of the argument.

Remark 7. If V1 is confining, V1(x)  ≳  |x|2 for |x| ⩾ R (for instance, V1(x) = c|x|2k , c > 0 and k a positive integer, a typical case where V1 may be super-quadratic while Hx and H remain self-adjoint), then we can estimate ${\Vert}x{\phi }^{x}{{\Vert}}_{{L}_{x}^{2}}$ uniformly in time. If V1 = 0, or more generally if V1(x) → 0 as |x| → , we must expect some linear growth in time ${\Vert}x{\phi }^{x}(t){{\Vert}}_{{L}_{x}^{2}}\lesssim \left\langle t\right\rangle $, and the order of magnitude in t is sharp, corresponding to a dispersive phenomenon.

Remark 8. The framework of a cubic system–bath coupling $W(x,y)=\frac{1}{2}\overrightarrow{\eta }\cdot x\vert y{\vert }^{2}$ as described in section 2.2 is recovered by taking σx = σy = 1 in example 3. In addition, in the presence of a quartic confinement with ${k}_{4}^{0} > 0$, in view of remark 7, we also know that ${\Vert}\langle y\rangle \vert y\vert {\phi }^{y}(t){{\Vert}}_{{L}_{y}^{2}}$ is bounded uniformly in t.

Adding control on the gradients of the functions φx (t), φy (t) respectively ϕx (t), ϕy (t), allows also error estimates at the level of the kinetic energy. For a proof, see appendix B.2.

Proposition 2. Under assumption 1, there exists a constant C > 0 depending on second order derivatives of the potentials such that we have the following error estimates:

Brute-force approach: for ψapp(t, x, y) = φx (t, x)φy (t, y)exp(itW(0, 0)) defined by (5),

and

Mean-field approach: for ${\phi }_{\mathrm{app}}(t,x,y)={\phi }^{x}(t,x){\phi }^{y}(t,y)\mathrm{exp}(\mathrm{i}{\int }_{0}^{t}\left\langle W\right\rangle \mathrm{d}s)$ defined by (7), analogous estimates for ψ(t) − ϕapp(t) hold.

Remark 9. The strategy used to prove proposition 2 can be iterated to infer error estimates in Sobolev spaces of higher order, ${H}^{k}({\mathbb{R}}^{n+d})$ for k ⩾ 2, provided that we consider momenta of the same order k, which explains the interest in the functional spaces Σk . Error estimates in such spaces can also be obtained by first proving that ψ and the approximate solution(s) remain in Σk , and then interpolating with the L2 error estimate from proposition 1.

3.4. Error estimates for quadratic observables

For obtaining quadratic estimates, we consider observables such as the energy or the momenta, that is, operators that are differential operators of order at most 2 with bounded smooth coefficients. These differential operators have their domain in ${H}^{2}({\mathbb{R}}^{n+d})$, as the operator H. More generally, we could consider pseudo-differential operators B = op(b) associated with a smooth real-valued function b = b(Z) with $Z=(z,\zeta )\in {\mathbb{R}}^{2(n+d)}$, whose action on functions $f\in \mathcal{S}({\mathbb{R}}^{n+d})$ is given by

We assume that b satisfies the Hörmander condition

Equation (11)

that is, b is a symbol of order 2, see e.g. [31, chapter I.2]. We shall also consider observables that depend only on the variable x or the variable y. The following estimates are proven in appendix B.3.

Proposition 3. Under assumption 1, for $b\in {C}^{\infty }({\mathbb{R}}^{n+d})$ satisfying 11 and B = op(b), there exists a constant Cb > 0 such that we have the following error estimates:

Brute-force approach: for ψapp(t, x, y) = φx (t, x)φy (t, y)exp(itW(0, 0)), defined by (5), the error

satisfies

where Nn+d > 0 depends on n + d, M(x, y) = ∂x y W(x, y), while

Mean-field approach: for ${\phi }_{\mathrm{app}}(t,x,y)={\phi }^{x}(t,x){\phi }^{y}(t,y)\mathrm{exp}(\mathrm{i}{\int }_{0}^{t}\left\langle W\right\rangle \mathrm{d}s)$ defined by (7), the error ⟨ψ(t), (t)⟩ − ⟨ϕapp(t), app(t)⟩ satisfies a similar estimate.

Remark 10. The averaging process involved in the action of an observable on a wave function allows to prove estimates like the one in proposition 3, that are more precise than the standard ones stemming from norm estimates,

Remark 11. We point out that the error is governed by derivatives of second order in W, involving a derivative in the y variable that is supposed to be small. Besides, note that the direct use of an estimate on the wave function itself would have involved H2 norms of ψapp(s), while this estimate only requires H1 norms. This first improvement is due to the averaging process present in Egorov theorem.

3.5. Energy conservation

The error estimates of propositions 13, do not allow to distinguish between the brute-force single point collocation and the mean-field Hartree approach. However, in computational practice most of the employed methods are of mean-field type. Why? Our previous analysis, that specifically addresses the coupling of quantum systems, does not allow for an answer, and we resort to a more general point of view. Both approximations, the brute-force and the mean-field one, are norm-conserving. However, the mean-field approach is energy-conserving with the same energy as 1. At first sight, this is surprising, since the mean-field Hamiltonian Hmf(t) depends on time. In a more general framework, where the time-dependent Hartree approximation is considered as application of the time-dependent Dirac–Frenkel variational principle on the manifold of product functions, energy conservation is immediate, see [32, section 3.2] or [33, chapter II.1.5].

Lemma 1. Under assumption 1 and considering the mean-field approach ${\phi }_{\mathrm{app}}(t,x,y)={\phi }^{x}(t,x){\phi }^{y}(t,y)\mathrm{exp}(\mathrm{i}{\int }_{0}^{t}\left\langle W\right\rangle (s)\mathrm{d}s)$ defined by (7), we have

where the mean-field Hamiltonian is given by

Below in appendix B.5 we give an elementary ad-hoc proof of lemma 1.

Remark 12. In the brute-force case, the approximate Hamiltonian Hbf = Hx + Hy + W(x, 0) + W(0, y) − W(0, 0) is time-independent, and we have

However this conserved value does not correspond to the exact energy of (1), but only to an approximation of it.

4. An exemplary proof

Here we discuss our basic proof strategy and apply it for the norm estimate of proposition 1. The norm estimates of example 3, propositions 2 and 4 and the observable estimates of propositions 3 and 5 are carried out in appendices B and C.

Lemma 2. Let N ⩾ 1, A be self-adjoint on ${L}^{2}({\mathbb{R}}^{N})$, and ψ solution to the Cauchy problem

where ${\psi }_{0}\in {L}^{2}({\mathbb{R}}^{N})$ and ${\Sigma}\in {L}_{\mathrm{loc}}^{1}({\mathbb{R}}^{+};{L}^{2}({\mathbb{R}}^{N}))$. Then for all t ⩾ 0,

This standard lemma is our main tool for proving norm estimates. It will be applied with either h = 1 or h = ɛ as parameter. Its proof is given in appendix A. Now we present the proof of proposition 1.

Proof. Denote by rψ = ψψapp and rϕ = ψϕapp the errors corresponding to each of the two approximations presented in sections 3.1 and 3.2, respectively. They solve

Equation (12)

We note that both approximations and their components are norm-conserving for all times t ⩾ 0, that is,

  • In the case of the brute-force approach, we consider the Taylor expansions (6) and derive the estimates
    Equation (13)
  • In the mean-field approach, we note that for $(t,x,y)\in \mathbb{R}\times {\mathbb{R}}^{n+d}$,
    Like we did in the brute-force approach, we may use either of the estimates

In the first case, we come up with

Now we have

where we have used Cauchy–Schwarz inequality for the last term. We infer

where we have used Young inequality (α + β)2 ⩽ 2(α2 + β2), hence

and finally

Equation (14)

In the case of the second type approximation for δW, we similarly find

Proposition 1 then follows from lemma 2 with h = 1. □

5. Dimension reduction via semiclassical analysis

In this section, we consider coupled systems, where one part is governed by a semiclassically scaled Hamiltonian, that is, ${H}_{y}={H}_{y}^{\varepsilon }$ with

First we motivate such a partial semiclassical scaling in the context of system–bath Hamiltonians and introduce wave packets as natural initial data for the semiclassical part of the system. We explore partial semiclassical wave packet dynamics guided by classical trajectories and by trajectories with averaged potentials. Thus, the partially highly-oscillatory evolution of a PDE in dimension n + d is reduced to a less-oscillatory PDEs in dimensions n, and ODEs in dimension d. The corresponding error estimates in section 5.5 compare the true and the approximate product solution in norm and with respect to expectation values.

5.1. Semiclassical scaling

We reconsider the system–bath Hamiltonian with cubic coupling of section 2.2, now formulated in physical coordinates (X, Y), that is,

where the coordinates X and Y of the system and the bath part are prescaled, resulting in the single mass parameters μ1μ2 for each subsystem and one single harmonic frequency ω2 for the bath (noting that, alternatively, several harmonic bath frequencies ω2,j could be introduced, without modifying the conclusions detailed below). The corresponding time-dependent Schrödinger equation reads

We perform a local harmonic expansion of the potential Vs(X) around the origin X = 0 and assume that it is possible to determine a dominant frequency ω1. We then define the natural length scale of the system as

Rescaling coordinates as $(x,y)=\frac{1}{a}(X,Y)$, we obtain

where we have introduced the dimensionless parameters

and denoted ${V}_{1}(x)=\frac{1}{\hslash {\omega }_{1}}{V}_{\mathrm{s}}({a}_{1}x)$ and ${\overrightarrow{\eta }\;}^{\prime }=\frac{{a}_{1}}{{\mu }_{1}{\omega }_{1}^{2}}\enspace \overrightarrow{\eta }$. The rescaling of the system potential Vs and the coupling vector $\overrightarrow{\eta }$ do not alter their role in the Hamiltonian, whereas the two dimensionless parameters ɛ and ϖ deserve further attention. We now consider the régime where both the mass ratio ɛ between system and bath and the frequency ratio ϖ between bath and system are small, that is, where the system is viewed as 'light' and 'fast' when compared to the 'heavy' and 'slow' bath.

Example 4. For the hydrogen molecule H2, where the electrons are considered as the quantum subsystem while the interatomic vibration is considered as the classical subsystem, we have μ1 = me and μ2 = 918.6me. Further, the characteristic electronic energy is of the order of ℏω1 = 1Eh while the first vibrational level is found at ℏω2 = 0.020 05Eh. Hence the dimensionless parameters are both small, ɛ = 0.032 99 and ϖ = 0.020 05.

Example 5. As a second example, we consider coupled molecular vibrations, exemplified by the H2 molecule in a 'bath' of rare-gas atoms, here chosen as krypton (Kr) atoms. The H2 vibration is now considered as a quantum system interacting with weak intermolecular vibrations. The reduced masses are given as μ1 (H–H) = 0.5u = 911.44me (where u refers to atomic mass units), μ2 (Kr–Kr) = 41.9u = 76.379 × 103 me, and μ3 (H2–Kr) = 1.953u = 3560.10me. The vibrational quanta associated with these vibrations are ℏω1(H–H) = 4159.2 cm−1 = 0.0189Eh, ℏω2(Kr–Kr) = 21.6 cm−1 = 9.82 × 10−5 Eh, and ℏω3(H2–Kr) = 26.8 cm−1 = 1.22 × 10−4 Eh (see references [34, 35]). The resulting dimensionless mass ratios are given as ${{\epsilon}}_{12}=\sqrt{{\mu }_{1}/{\mu }_{2}}=0.109$ and ${{\epsilon}}_{13}=\sqrt{{\mu }_{1}/{\mu }_{3}}=0.51$, and the corresponding frequency ratios are ϖ12 = ω2/ω1 = 0.005 and ϖ13 = ω3/ω1 = 0.006. In the case of the H2–Kr relative motion, note that the frequency ratio ϖ13 is indeed small whereas the mass ratio is epsilon13 ∼ 0.5; this shows that the quantum–classical boundary is less clearly defined than in the first example of coupled electronic-nuclear motions. In such cases, different choices can be made in defining the quantum–classical partitioning.

In an idealized setting, where ɛ is considered as a small positive parameter whose size can be arbitrarily small, we would say that

and view the system–bath Hamiltonian Hsb as an instance of a partially semiclassical operator

whose potentials V1(x) and V2(y) are independent of the semiclassical parameter ɛ and satisfy the growth conditions of assumption 1. As emphasized in section 2.2, the cubic coupling potential W(x, y) does not satisfy the subquadratic estimate, but can be controlled by additional moments of the approximate solution. A corresponding rescaling of time, t = ɛω1 τ, translates the time-dependent Schrödinger equation to its semiclassical counterpart

Equation (15)

where the physical and the rescaled wave functions are related via

Remark 13. Criteria for justifying a semiclassical description are somewhat versatile in the literature. Our scaling analysis shows, that for system–bath Hamiltonians with cubic coupling the obviously small parameter ɛ, that describes a ratio of reduced masses, has to be complemented by an equally small ratio of frequencies ϖ. Otherwise, the standard form of an ɛ-scaled Hamiltonian, as it is typically assumed in the mathematical literature, does not seem appropriate.

5.2. Semiclassical initial data and ansatz

As before, the initial data separate scales,

Equation (16)

where we now assume that g ɛ is a semiclassically scaled wave packet,

Equation (17)

with $({q}_{0},{p}_{0})\in {\mathbb{R}}^{2d}$, a smooth and rapidly decreasing, i.e. $a\in \mathcal{S}({\mathbb{R}}^{d};\mathbb{C})={\cap }_{k\in \mathbb{N}}{{\Sigma}}^{k}$. In the typical case, where the bath is almost structureless (say, near harmonic), the amplitude a could be chosen as a complex Gaussian, but not necessarily. We now seek an approximate solution of the form ${\psi }_{\mathrm{app}}^{\varepsilon }(t,x,y)={\psi }_{1}^{\varepsilon }(t,x){\psi }_{2}^{\varepsilon }(t,y)$, where ${\psi }_{2}^{\varepsilon }$ is a semiclassically scaled wave packet for all time,

Equation (18)

Here, $(q(t),p(t))\in {\mathbb{R}}^{2d}$, the phase $S(t)\in \mathbb{R}$, and the amplitude ${u}_{2}(t)\in \mathcal{S}({\mathbb{R}}^{d},\mathbb{C})$ must be determined.

Remark 14. We note that our approximation ansatz differs from the adiabatic one, that would write the full Hamiltonian as ${H}^{\varepsilon }=-\frac{{\varepsilon }^{2}}{2}{{\Delta}}_{y}+{H}_{\mathrm{f}}(y)$, where

is an operator, that parametrically depends on the 'slow' variable y and acts on the 'fast' degrees of freedom x. From the adiabatic point of view, one would then construct an approximate solution as ${\psi }_{\text{bo}}^{\varepsilon }(t,x,y)={\Phi}(x,y){\psi }_{2}^{\varepsilon }(t,y)$, where Φ(x, y) is an eigenfunction of the operator Hf(y); here, the subscript 'bo' stands for Born–Oppenheimer. The result of corollary 2 emphasizes the difference between these two points of view.

We denote by

Equation (19)

the part of the approximate solution that just contains the amplitude. With this notation,

The analysis developed in the next two sections allows to derive two different approximations, based on ordinary differential equations governing the semiclassical wave packet part, which are justified in section 5.5 (see proposition 4).

5.3. Approximation by partial Taylor expansion

Plugging the expression of ${\psi }_{\mathrm{app}}^{\varepsilon }(t,x,y)$ into (15) and writing $y=q(t)+z\sqrt{\varepsilon }$ in combination with the Taylor expansions

we find:

where the argument of ${u}_{\mathrm{app}}^{\varepsilon }$ and its derivatives are taken in $z=\frac{y-q(t)}{\sqrt{\varepsilon }}$. To cancel the first four terms in the $\sqrt{\varepsilon }$ line, it is natural to require

Equation (20)

Now cancelling the first four terms in the first line of the right-hand side yields

Equation (21)

In other words, (q(t), p(t)) is the classical trajectory in y, and S(t) is the associated classical action. At this stage, we note that the term $z\cdot {\nabla }_{y}W(x,q){u}_{\mathrm{app}}^{\varepsilon }$ is not compatible with decoupling the variables x and z (or equivalently, x and y). Using that ${\Vert}{\nabla }_{y}W{{\Vert}}_{{L}^{\infty }}$ is assumed to be small, the above computation becomes

In view of (19), we set

Equation (22)

Equation (23)

Equation (23) is a Schrödinger equation with a time-dependent harmonic potential: it has a unique solution in L2 as soon as $a\in {L}^{2}({\mathbb{R}}^{d})$. In addition, since a ∈ Σk for all $k\in \mathbb{N}$, ${u}_{2}\in C(\mathbb{R};{{\Sigma}}_{z}^{k})$ for all $k\in \mathbb{N}$. The validity of this approximation is stated in proposition 4 below. Note that if a is a Gaussian state, then u2 too and its (time-dependent) parameters—width matrix and centre point—can be computed by solving ODEs (see e.g. [30, 33, 36, 37] and references therein).

5.4. Approximation by partial averaging

Following e.g. [38, 39] or [37, section 2], we write

where the averages are with respect to $\vert {\psi }_{2}^{\varepsilon }(t,y){\vert }^{2}\mathrm{d}y$. For example,

Equation (24)

where we anticipate the fact that the ${L}_{y}^{2}$-norm of ${\psi }_{2}^{\varepsilon }(t)$ is independent of time. We almost literally repeat the previous argument and find that

with ${\tilde{r}}_{j}^{\varepsilon }={\mathrm{v}}_{j}{u}_{\mathrm{app}}^{\varepsilon }$, j = 1, 2, and z is taken as $z=(y-q(t))/\sqrt{\varepsilon }$. The parameters satisfy the equations of motion

We see that we can now define the approximate solution by:

Equation (25)

Since the matrix ${\langle {\nabla }^{2}{V}_{2}\rangle }_{y}(t)$ is real-valued, we infer that the ${L}_{z}^{2}$-norm of u2(t) is independent of time, hence ${\Vert}{\psi }_{2}(t){{\Vert}}_{{L}_{y}^{2}}={\Vert}{u}_{2}(t){{\Vert}}_{{L}_{z}^{2}}={\Vert}a{{\Vert}}_{{L}^{2}}$. The equation in u2 is now nonlinear, and can be solved in Σ1, since ∇V2 is at most linear in its argument: ${u}_{2}\in C(\mathbb{R};{{\Sigma}}_{z}^{1})$, and higher Σk regularity is propagated. Here again, if a is a Gaussian, then so is u2 and its width and centre can be computed by solving ODEs (see [30, 33]). Note also that, differently from the previous setting, u2 is now ɛ-dependent via the quantity ${\langle {\nabla }^{2}{V}_{2}\rangle }_{y}(t)$ (see (24)). However, this dependence is very weak since a Taylor expansion in (24) shows that u2 is close in any Σk  norm from the solution of the equation (23). For this reason, we do not keep memory of this ɛ-dependence and write u2. By contrast, the ɛ-dependence of ${\psi }_{1}^{\varepsilon }$ is strong since it involves oscillatory features in time.

5.5. The approximation results

The main outcome of the approximations can be stated as follows, and is proved in appendix C.1:

Proposition 4. Let ψɛ be the solution to (15) and (16), with g ɛ given by (17). Then with ${\psi }_{\mathrm{app}}^{\varepsilon }$ given either like in section 5.3 or like in section 5.4, there exist constants K0, K1 independent of ɛ such that for all t ⩾ 0,

Corollary 1. Assume $\eta {:=}{\Vert}{\nabla }_{y}W{{\Vert}}_{{L}^{\infty }}\ll \sqrt{\varepsilon }$, then for all T > 0,

Remark 15. Using the same techniques as in appendix B.2, one can prove estimates on higher regularity norms, using ɛ-derivatives in y and standard ones in x. For example, if a ∈ Σ4, then there exists K0, K1 independent of ɛ such that

We refer to [40] (see also [30, chapter 12]) for more detailed computations.

Note that, in both approximations, the evolution of u2 corresponds to the standard quadratic approximation. In particular, if a is Gaussian, then u2 is Gaussian at all time, and solving the equation in u2 amounts to solving ordinary differential equations. However, the equation (22) solved by ψ1(t) is still quantum, such that a reduction of the total space dimension of the quantum system has been made from n + d to n.

Let us now discuss the approximation of observables that we choose as acting only in the variable y. Due to the presence of the small parameter ɛ, we choose semiclassical observables and associate with $b\in {C}_{c}^{\infty }({\mathbb{R}}^{2d})$ (b smooth and compactly supported) the operator opɛ (b) whose action on functions $f\in \mathcal{S}({\mathbb{R}}^{d})$ is given by

As before, the error estimate is better for quadratic observables than for the wave functions. More specifically, the following result, that is proved in appendix C.2, improves the error estimate from proposition 4 by a factor $\sqrt{\varepsilon }$.

Proposition 5. Let ψɛ be the solution to (15) and (16), with g ɛ given by (17). Then with $b\in {C}_{c}^{\infty }({\mathbb{R}}^{2d})$ and ${\psi }_{\mathrm{app}}^{\varepsilon }$ given either like in section 5.3 or in section 5.4, there exist a constant K independent of ɛ such that for all t ⩾ 0,

Remark 16. Of course, we could have considered a mixed setting consisting of pseudodifferential operators as in section 3.4 in the variable x, and semiclassical as above in the variable y. One would then obtain estimates mixing those of propositions 3 and 5.

6. A numerical example

For an illustrative numerical application, we consider a system–bath type Hamiltonian with cubic coupling $W(x,y)=\frac{1}{2}\eta x{y}^{2}$, as developed in section 2.2,

in dimension d = n = 1,

Equation (26)

The mass ratio between the system and the bath is moderately small, μ1/μ2 = 0.25. The system potential is a quartic double well, while the bath potential is harmonic,

The length scale = 4a1 of the double well is a multiple of the system's natural harmonic unit a1. The initial data ${\psi }_{0}(x,y)={\varphi }_{0}^{x}(x){\varphi }_{0}^{y}(y)$ are the ground state of the bivariate harmonic oscillator, that results from the harmonic approximation of Vs(x) + Vb(y) around the left well (x, y) = (0, 0). The coupling constant η < 0 is negative to ensure that the total Hamiltonian's ground state is localised in the right well (x, y) = (2, 0), providing a setup with pronounced non-equilibrium dynamics. For such a system–bath model, the gradient-free error estimate (10) of example 3 is given by

Equation (27)

Figure 1 presents the results from the following numerical experiment: (see the supplmentary material for computational details https://stacks.iop.org/A/54/414002/mmedia) we identify a frequency ratio ϖref = 1/100 between bath and system and a coupling parameter ${\eta }_{\text{ref}}=-{k}_{2}^{0}/(3{a}_{1}\ell )$ as generating 'reference' system–bath dynamics. For this parameter choice, the mean-field approximation, when compared with a numerically converged MCTDH approximation, results in roughly a 0.1% error after 10 units of the natural harmonic time scale t1 = 1/ω1 of the system (blue curve), see figure 1(a). Hence, the Hartree approximation is excellent on the time scale under study. Decreasing the frequency ratio by a factor of four, roughly halves the error (red curve). And, as expected, an increase in the coupling strength also increases the error (grey curve), while decreasing the coupling also decreases the error (yellow curve). The corresponding plot of figure 1(b) illustrates that the upper bound of the theoretical error estimate correctly captures the initial slope of all four error curves, while slightly over-estimating the actual error as time evolves. A more detailed assessment of the error estimate (27), in particular of its long-time behaviour (up to 150 ps), when the mean-field approximation goes up to errors of the order of 1%, and a more complete screening of physically relevant parameter régimes are work in progress for a numerical companion paper to the present theoretical study.

Figure 1.

Figure 1. Error of the mean-field approximation for a family of system–bath models with cubic coupling $W(x,y)=\frac{1}{2}\eta x{y}^{2}$ as a function of time. The left plot shows the time evolution of the norm of the error for four different variations of the parameter values as given in table 1. The right plot shows the corresponding upper bound of the error estimate (27).

Standard image High-resolution image

Table 1. Parameters defining the four numerically simulated variations of the system–bath Hamiltonian (26). The blue model uses the coupling parameter ${\eta }_{\text{ref}}=-{k}_{2}^{0}/(3{a}_{1}\ell )$ and the frequency ratio ϖref = 1/100. The red model varies the frequency ratio, the grey and yellow models the coupling strength.

Model variation ϖ = ω2/ω1 η (coupling)
Blue ϖref ηref
Red $\frac{1}{4}{\varpi }_{\text{ref}}$ ηref
Grey ϖref $\frac{9}{8}{\eta }_{\text{ref}}$
Yellow ϖref $\frac{3}{4}{\eta }_{\text{ref}}$

7. Conclusion and outlook

We have presented quantitative error bounds for the approximation of quantum dynamical wave functions in product form. For both considered approaches, a brute-force single point collocation and the conventional mean-field Hartree approximation, we have obtained similar error estimates in L2-norm (proposition 1, example 3), in H1-norm (proposition 2), and for quadratic observables (proposition 3). To our knowledge, such general estimates, that quantify decoupling in terms of flatness properties of the coupling potential, are new. The corresponding analysis for semiclassical subsystems (propositions 4 and 5) confirms the more general finding, that error estimates for quadratic observables provide smaller bounds than related norm estimates. The single product analysis, as presented here, provides a well-posed starting point for the investigation of more elaborate approximation methods. If the initial data satisfy

then we may invoke the linearity of (1) to write $\psi (t,x,y)={\sum }_{j=1}^{J}\enspace {\psi }_{j}(t,x,y)$, where each ψj solves i∂t ψj = j , with ${\psi }_{j\vert t=0}={\varphi }_{0j}^{x}\otimes {\varphi }_{0j}^{y}$. We approximate each ψj ψj,app individually in one of the ways discussed in the present paper and use the triangle inequality for

where the norm can be an L2 or an energy norm, for instance. However, working on each ψj instead of ψ directly, seems to prevent control of the limit J. Multi-configuration methods therefore use ansatz functions of the form

where the families ${({\varphi }_{j}^{(x)}(t))}_{j\geqslant 1}$ and ${({\varphi }_{\ell }^{(y)}(t))}_{\ell \geqslant 1}$ satisfy orthonormality or rank conditions, while gauge constraints lift redundancies for the coefficients ${a}_{k\ell }(t)\in \mathbb{C}$. We view our contributions here as an important first step for a systematic assessment of such multi-configuration approximations in the context of coupled quantum systems. A numerical companion paper, that explores the dynamics of system–bath Hamiltonians with cubic coupling on multiple time-scales with respect to various parameter régimes, is currently in preparation.

Acknowledgments

Adhering to the prevalent convention in mathematics, we chose an alphabetical ordering of authors. We thank Loïc Joubert-Doriol, Gérard Parlant, Yohann Scribano, and Christof Sparber for stimulating discussions in the context of this paper. We acknowledge support from the CNRS 80|Prime project AlgDynQua and the visiting professorship program of the I-Site Future.

Data availability statement

All data that support the findings of this study are included within the article (and any supplementary files).

Appendix A.: General estimation lemmas

Here we provide the proof of the standard energy estimate, lemma 2.

Proof. In view of the self-adjointness of A, we have

and therefore, by the Cauchy–Schwarz inequality, $\frac{d}{\mathrm{d}t}{\Vert}\psi (t){\Vert}\leqslant \frac{1}{h}{\Vert}{\Sigma}(t){\Vert}$. Integrating in time, we obtain

In the context of observables, refined error estimates will follow from the application of the following lemma.

Lemma 3. Let N ⩾ 1, A1, A2, B be self-adjoint on ${L}^{2}({\mathbb{R}}^{N})$, and ψ(1), ψ(2), solutions to the homogeneous Cauchy problems

where ${\psi }_{0}\in {L}^{2}({\mathbb{R}}^{N})$. Then, for all t ⩾ 0

with $\rho (s,t)=\left\langle {\psi }^{(1)}(s),\left[\mathrm{exp}(\mathrm{i}\;{A}_{2}(t-s)/h)B\enspace \mathrm{exp}(-\mathrm{i}\;{A}_{2}(t-s)/h),{A}_{1}-{A}_{2}\right]{\psi }^{(1)}(s)\right\rangle $, where we have denoted by $[\mathcal{A},\mathcal{B}]=\mathcal{AB-BA}$ the standard commutator.

Proof. We denote the unitary evolution operators by Uj (t) = exp(−iAj t/h) and calculate

Appendix B.: Proof of error estimates: partially flat coupling

In this section, we prove error estimates in L2-norm for general potentials W (remark 6), in H1-norm (proposition 2), and for quadratic observables (proposition 3).

B.1. Proof of remark 6

Proof. To prove the estimates of example 3, recall that we have denoted

The fundamental theorem of calculus yields

so we have

and we replace the pointwise estimate of δW with

The estimate on Σϕ becomes

and we conclude by resuming the same estimates as above:

and, in view of the inequality

we find

B.2. Proof of proposition 2

To prove error estimates in ${H}^{1}({\mathbb{R}}^{n+d})$, we differentiate (12) in space, and two aspects must be considered: (i) in our framework, the operator ∇x,y does not commute with H. (ii) We must estimate ∇x,y Σψ and ∇x,y Σϕ . Indeed, we compute

and

In the typical case where V1 is harmonic, ∇x V1 is linear in x, and so xrψ appears as a source term. Note that in the general setting of assumption 1, $\vert {\nabla }_{x}{V}_{1}(x)\vert \lesssim \left\langle x\right\rangle $.

Remark 17. If ∇x V1 and ∇x W are bounded, then lemma 2 yields

The term ${\Vert}{r}_{\psi }(s){{\Vert}}_{{L}^{2}}$ is estimated in proposition 1, and ${\Vert}{\nabla }_{x}{{\Sigma}}_{\psi }(s){{\Vert}}_{{L}^{2}}$ is estimated below.

Multiplying (12) by x, we find similarly

Energy estimates provided by lemma 2 applied to the equation for ∇x rψ and xrψ then yield a closed system of estimates:

where we have used the estimate |∇x V1 + ∇x W| ⩽ C(1 + |x|), and the uncertainty principle (uncertainty in x, Cauchy–Schwarz in y),

The Gronwall lemma then yields

for some C > 0. We compute

The first term in controlled by $\vert y\vert {\Vert}{\nabla }_{x}{\nabla }_{y}W{{\Vert}}_{{L}^{\infty }}\vert {\psi }_{\mathrm{app}}\vert $. The second term is controlled like in section 3.3, by replacing ψapp with ∇x ψapp. We can of course resume the same approach when considering ∇y rψ , and the analogue of the above first term is now controlled by $\vert x\vert {\Vert}{\nabla }_{x}{\nabla }_{y}W{{\Vert}}_{{L}^{\infty }}\vert {\psi }_{\mathrm{app}}\vert $. Finally, in the case of rϕ , computations are similar (we do not keep track of the precise dependence of multiplicative constants here).

B.3. Proof of proposition 3

Proof. We use lemma 3 for the operators H and the approximate Hamiltonian Hbf,

Equation (B.1)

to obtain

where

By Egorov theorem, see [41, theorem 11.1], the operator B(σ) is also a pseudodifferential operator, that is, B(σ) = op(b(σ)) for some function b(σ) that satisfies the growth condition (11). We have

with the notations of appendix B.1. Then, by the direct estimate of lemma 4 below,

Equation (B.2)

where Cb(σ) > 0 depends on derivative bounds for the function b(σ) and

We therefore obtain

Using the rectangular n × d matrix M(x, y) introduced in remark 4, the gradient of δW(x, y) can be written as

We estimate the Sobolev norm by

so that integration in time provides

where the constant Cb = maxσ∈[0,t] Cb(σ) depends on derivatives of b. In the mean-field case, the approximate Hamiltonian is time-dependent,

Equation (B.3)

The difference of the Hamiltonians is also a function, which is now time-dependent, HHmf(t) = W + ⟨W⟩(t) − ⟨Wx (t) − ⟨Wy (t). However, it is easy to check that a similar estimate can be performed, leading to an analogous conclusion. □

B.4. Commutator estimate

We now explain the commutator estimate used in the previous subsection:

Lemma 4. Let N ⩾ 1 and b = b(z, ζ) be a smooth function on ${\mathbb{R}}^{2N}$ satisfying the Hörmander growth condition (11). Let δW be a smooth function on ${\mathbb{R}}^{N}$ with bounded derivatives. Then, there exist constants Cb > 0 and MN > 0 such that

for all $\psi \in {H}^{1}({\mathbb{R}}^{N})$.

Proof. We explicitly write the commutator as

We Taylor expand the function δW(z) around the point z', so that

with

Corresponding to the above decomposition, we write [op(b), δW]ψ(z) = f1(z) + f2(z) and estimate the two summands separately. We observe that (zz')exp(iζ ⋅ (zz')) = −i∇ζ  exp(iζ ⋅ (zz')) and perform an integration by parts to obtain

Therefore,

where the constant Cb > 0 depends on derivative bounds of the function b. For the remainder term of the above Taylor approximation we write

and obtain that

where Cb' > 0 depends on derivative bounds of b, and MN > 0 depends on the dimension N. □

B.5. Proof of energy conservation

Here we provide an elementary ad hoc proof for energy conservation of the time-dependent Hartree approximation, lemma 1.

Proof. A first observation is that

Indeed,

with ${W}_{\mathrm{app}}(t)={\langle W\rangle }_{y}(t)+{\langle W\rangle }_{x}(t)-\left\langle W\right\rangle (t).$ We deduce

where we have used the self-adjointness of Hmf(t) and norm-conservation in the multiplicative components. Secondly, since

the approximate energy coincides with the actual energy, and we obtain the aimed for result. □

Appendix C.: Proof of error estimates in the semiclassical régime

Here we present the proofs of the semiclassical estimates given in propositions 4 and 5.

C.1. Error estimates for the wave function

In this section, we prove proposition 4, and comment on the constants K0, K1, which may be analyzed more explicitly in some cases.

C.1.1. Approximation by partial Taylor expansion

Proof. Section 5.3 defines an approximate solution of the from

with

and

It solves the equation

where the remainder ${r}_{1}^{\varepsilon }$ is due to the Taylor expansion in V2, and satisfies the pointwise estimate

while the remainder ${r}_{2}^{\varepsilon }$ is due to the Taylor expansion in W, and satisfies the pointwise estimate

This implies for the L2-norm,

Lemma 2 then yields, with now h = ɛ,

According to the signature of ∇2 V2(q(t)), the quantities ${\Vert}\vert z{\vert }^{3}{u}_{2}(s){{\Vert}}_{{L}_{z}^{2}}$ and ${\Vert}\vert z\vert {u}_{2}(s){{\Vert}}_{{L}_{z}^{2}}$ may be bounded uniformly in s ⩾ 0 or not. For instance, they are bounded if ∇2 V2 is uniformly positive definite, or at least uniformly positive definite along the trajectory q. On the other hand, we always have an exponential bound, even if it may not be sharp,

for some constants C0, C1 > 0. This control is sharp in the case where ∇2 V2 is uniformly negative definite. See e.g. [30, lemma 10.4] for a proof of the exponential control, and [30, section 10.5] for a discussion on its optimality. In particular, for bounded time intervals, the (relative) error is small if ${\Vert}{\nabla }_{y}W{{\Vert}}_{{L}^{\infty }}\ll \sqrt{\varepsilon }\ll 1$. □

Remark 18. If ∇y W is not bounded, e.g. ${\nabla }_{y}W(x,y)=\eta {\left\langle x\right\rangle }^{\gamma }$, then we can replace the previous error estimate with

In other words, the cause for the unboundedness of ∇y W is transferred to a weight for ${\psi }_{1}^{\varepsilon }$. Similarly, if ∇y W is unbounded in y, we may change the weight in the terms ${\Vert}\vert z{\vert }^{k}{u}_{2}{{\Vert}}_{{L}_{z}^{2}}$, after substituting y with $q+z\sqrt{\varepsilon }$.

C.1.2. Approximation by partial averaging

Proof. The semiclassical approximation obtained by partial averaging reads:

with

and

To estimate the size of ${\tilde{r}}_{1}$ and ${\tilde{r}}_{2}$ introduced in section 5.4, we might argue again via Taylor expansion. Indeed,

where

Hence, we have for all averages f = V2, ∇V2, ∇2 V2, W(x, ⋅) that

where the error constant depends on moments of |u2|2. In particular, if u2(0) is Gaussian, the odd moments of |u2(t, z)|2 vanish, and the above estimate improves to $\mathcal{O}(\varepsilon )$. Hence, the L2-norm of ${\tilde{r}}_{1}$ is $\mathcal{O}(\sqrt{\varepsilon })$ close to the L2-norm of r1, and the L2-norm of ${\tilde{r}}_{2}$ is $\mathcal{O}(\eta \sqrt{\varepsilon })$, $\eta ={\Vert}{\nabla }_{y}W{{\Vert}}_{{L}^{\infty }}$, close to the L2-norm of r2 (with each time an extra $\sqrt{\varepsilon }$ gain in the above mentioned Gaussian case). In particular, the order of magnitude for the difference between exact and approximate solution is the same as in the previous subsection, only multiplicative constants are affected. We emphasize that the constants C0 and C1 from the previous subsection are in general delicate to assess. On the other hand, in specific cases (typically when u2 is Gaussian and ∇2 V2 is known), they can be computed rather explicitly. □

C.2. Error estimates for quadratic observables

The proof of proposition 5 is discussed in the next two sections.

C.2.1. Approximation by Taylor expansion

Proof. Taylor expansion yields a time-dependent Hamiltonian ${H}_{\mathrm{app}}^{\varepsilon }={H}_{\text{te}}^{\varepsilon }$ with

where q = q(t). In particular, the difference ${H}^{\varepsilon }-{H}_{\text{te}}^{\varepsilon }$ is a function,

In view of lemma 3, if B = opɛ (b) with $b\in {C}_{c}^{\infty }({\mathbb{R}}^{2d})$, it yields (a posteriori estimate)

where

By Egorov theorem [41, theorem 11.1], B(σ) = ɛ  opɛ (b(σ)) for a function $b(\sigma )\in {C}_{c}^{\infty }({\mathbb{R}}^{d})$. Therefore, by semiclassical calculus,

where ${\Vert}{\mathrm{o}\mathrm{p}}_{\varepsilon }({r}^{\varepsilon }(s,t)){{\Vert}}_{\mathcal{L}({L}^{2})}$ is bounded uniformly in ɛ, whence the estimate of proposition 5. □

C.2.2. Approximation by partial averaging

Proof. The time-dependent Hamiltonian is ${H}_{\mathrm{app}}^{\varepsilon }={H}_{\text{pa}}^{\varepsilon }$ with

where q = q(t). In particular, as in the preceding case, the difference ${H}^{\varepsilon }-{H}_{\text{pa}}^{\varepsilon }$ is a time-dependent function

and the arguments developed above also apply. □

C.3. Time-adiabatic approximation

The evolution equations for the quantum part of the system, equations (22) and (25), can be written as an adiabatic problem:

where $\mathfrak{h}(t)$ is one of the time-dependent self-adjoint operators on ${L}^{2}({\mathbb{R}}^{n})$

We assume here that $\mathfrak{h}(t)$ has a compact resolvent and thus, that its spectrum consists in a sequence of time-dependent eigenvalues

We also assume that some eigenvalue Λj (t) is separated from the remainder of the spectrum for all $t\in \mathbb{R}$ and that the initial datum ${\varphi }_{0}^{x}$ is in the eigenspace of $\mathfrak{h}(0)$ for the eigenvalue Λj (0):

Equation (C.1)

Then adiabatic theory as developed by Kato [42] states that ψ1(t) stays in the eigenspace of Λj (t) on finite time, up to a phase.

Proposition 6 (Kato [42]). Assume we have (C.1) and that Λj (0) is a simple eigenvalue of $\mathfrak{h}(0)$ such that there exists δ0 > 0 for which

Denote by ${{\Phi}}_{j}^{x}(t)$ a family of normalized eigenvectors of $\mathfrak{h}(t)$ such that

Then, for all T > 0, there exists a constant CT > 0 such that

In contrast to the Born–Oppenheimer point of view recalled in remark 14, we obtain the following time-adiabatic extension for our wave-packet approximation:

Corollary 2. In the setting of propositions 4 and 6, we obtain the following approximate solution

Footnotes

Please wait… references are loading.