Brought to you by:
Paper

How do wave packets spread? Time evolution on Ehrenfest time scales

, and

Published 10 May 2012 © 2012 IOP Publishing Ltd
, , Citation Roman Schubert et al 2012 J. Phys. A: Math. Theor. 45 215307 DOI 10.1088/1751-8113/45/21/215307

1751-8121/45/21/215307

Abstract

We derive an extension of the standard time-dependent WKB theory, which can be applied to propagate coherent states and other strongly localized states for long times. It in particular allows us to give a uniform description of the transformation from a localized coherent state to a delocalized Lagrangian state, which takes place at the Ehrenfest time. The main new ingredient is a metaplectic operator that is used to modify the initial state in a way that the standard time-dependent WKB theory can then be applied for the propagation. We give a detailed analysis of the phase space geometry underlying this construction and use this to determine the range of validity of the new method. Several examples are used to illustrate and test the scheme and two applications are discussed. (i) For scattering of a wave packet on a barrier near the critical energy, we can derive uniform approximations for the transition from reflection to transmission. (ii) A wave packet propagated along a hyperbolic trajectory becomes a Lagrangian state associated with the unstable manifold at the Ehrenfest time; this is illustrated with the kicked harmonic oscillator.

Export citation and abstract BibTeX RIS

1. Introduction

The semiclassical propagation of wave packets is of fundamental importance in many applications of quantum mechanics and has therefore been studied extensively in the literature [19]. Time-dependent WKB approximations apply to extended initial states of the form

Equation (1)

where the phase function S(x) is real valued and smooth, and the amplitude A(x) is smooth and non-oscillating. The time evolution of such states can be expressed in terms of transport along a family of classical trajectories determined by initial positions x and initial momenta p = ∇S(x); see, e.g., [2, 5].

A different class of initial states is given by coherent states, see, e.g., [3, 7], which are strongly localized, and therefore, their propagation can be described up to certain times using only one trajectory and the linearized motion around it. A Gaussian coherent state is a function of the form

Equation (2)

where $Z=(p,q)\in \mathbb{R}^{2n}$ is a set of parameters, which determine where the state is localized in phase space and B is a complex symmetric n × n matrix with Im B > 0. The latter condition ensures that the state is localized around x = q. The roles of B and Z become clearer if we look at the Wigner function of such a state, which is a Gaussian of the form

Equation (3)

with phase space coordinates $z\in \mathbb{R}^n\times \mathbb{R}^n$ and where G is a positive-definite symmetric and symplectic matrix determined by B. As was described by Hepp and Heller [1013], the time evolution of a coherent state can in the semiclassical limit be described by using just the classical trajectory Z(t) = (p(t), q(t)) through Z and the linearized flow around it which is a time-dependent 2n × 2n symplectic matrix ${\mathcal S}(t)$. This simple description of the semiclassical propagation of coherent states made them a very useful tool in many applications, in particular in chemistry due to the work of Heller and coworkers and their appearance in initial value representations like the Herman–Kluk propagator [3, 4, 1318]

We will be using time-dependent WKB methods developed for states of the form (1) to understand the propagation of the states (2) for long times. In order to identify the relevant time scales, it is useful to consider the Wigner function of the time-evolved coherent state, which in leading order is given by

Equation (4)

where $G(t)= {\mathcal S}^{\dagger }(t)G{\mathcal S}(t)$ and ${\mathcal S}(t)$ is the linearized flow around the central trajectory Z(t). The basic idea leading to this approximation is that since a coherent state is localized around a point in phase space, one can approximate the Hamiltonian by its quadratic Taylor expansion around the classical trajectory Z(t). The resulting Schrödinger equation can then be solved explicitly. This approximation can only be expected to be accurate as long as the propagated state stays localized, and from (4), we see that this is the case as long as λmin[G(t)]/ℏ ≫ 1, where we denote by λmin/max[G(t)] the smallest or the largest eigenvalue of the matrix $G(t)={\mathcal S}^{t}(t)G{\mathcal S}(t)$, respectively. Since G(t) is symmetric and positive, we have λmin[G(t)] > 0, and that G(t) is symplectic implies that 1/λmin[G(t)] = λmax[G(t)], so the localization condition becomes

Equation (5)

In order to relate this more explicitly to the properties of the classical dynamics, we use the estimate $\lambda _{\rm max}[G(t)]=\Vert G(t)\Vert \le C \Vert {\mathcal S}(t)\Vert ^2$ and so the condition becomes $\Vert {\mathcal S}(t)\Vert \sqrt{\hbar }\ll 1$. Here, we use for a matrix A the norm induced by the Euclidean norm $\Vert x\Vert =\sqrt{ x\cdot x}$, i.e. $\Vert A\Vert :=\sup _{\Vert x\Vert =1} \frac{\Vert Ax\Vert }{\Vert x\Vert }$. The time at which the width of a propagated coherent state reaches order 1 is called the Ehrenfest time TE(ℏ), and by the previous discussion, we see that it is determined by the condition

Equation (6)

Hence, the Ehrenfest time depends on the dynamical properties of the classical system near the trajectory Z(t). If the trajectory Z(t) is hyperbolic, with the largest Liapunov exponent λ > 0, then ||S(t)|| ∼ eλt and so

Equation (7)

On the other hand, if the dynamics is integrable, we typically have $\Vert {\mathcal S}(t)\Vert \sim t$ for large t and then

Equation (8)

Ehrenfest showed in his classical paper [19] how one can recover classical mechanics from quantum dynamics by considering expectation values in propagated localized states. His construction breaks down if the state becomes delocalized, and is hence named Ehrenfest time for the time scale at which this occurs. In more recent years, the behaviour of time-evolved coherent states for long times has attracted the attention of mathematicians and rigorous estimates for long times have been derived; in particular, in [20, 21], it was shown that the remainder terms remain small for times satisfying $\vert t\vert < \frac{1}{3}T_E$ in the hyperbolic case. One should emphasize however that the breakdown of Ehrenfest's argument at the Ehrenfest time does not mean that the quantum to classical correspondence breaks down, it only means that the approximate propagation of a wave packet based on one central trajectory breaks down. If one includes more trajectories, one can use semiclassical propagation well beyond the Ehrenfest time as has been demonstrated in [22, 23], although from a mathematical point of view this is still a challenging problem.

The principal aim of this work is to present a propagation scheme for coherent states, which works at times of the order of the Ehrenfest time tTE and is able to describe the transition from a localized state (2) to an extended state (1) at the Ehrenfest time in a uniform way. In a previous paper [25], it was demonstrated that at the Ehrenfest time a coherent state that is transported along a hyperbolic trajectory becomes effectively a WKB state of the type (1) associated with the unstable manifold of that trajectory; see also [24]. Here, we will give the theoretical foundation for that claim by presenting a new propagation scheme, which is a combination of a time-dependent WKB approximation and a metaplectic correction.

The qualitative change in the nature of the state at the Ehrenfest time was described in [26] for the cat map and in [27] for the propagation of coherent states centred on unstable fixed points in one-dimensional multiple well potentials. In contrast to the present approach, that work was not based on direct propagation of the state in the position representation but on the Wigner function and the Egorov theorem from [28]. That means in particular that we can give a much more detailed description of the state at and beyond the Ehrenfest time.

The plan of the paper is as follows. In section 2, we recall time-dependent WKB theory for the propagation of states of the form (1) and rewrite it in an exact form where we explicitly separate the classical transport and the quantum dispersion. The main idea of this paper is then developed in section 3 where we show how to approximate the action of the dispersive part on a coherent state using a simple metaplectic operator and combine this with the time-dependent WKB approximation to obtain a method that allows us to describe the propagation of coherent states at and beyond the Ehrenfest time. In section 4, we develop the phase space geometry underlying the metaplectically extended WKB method. This allows us to determine its range of validity, in particular its dependence on dynamical properties of the classical system. We then develop the geometric picture even further in section 5 to allow more general Hamiltonians and the inclusion of caustics after the Ehrenfest time; the price we have to pay is that remainder estimates become less explicit now. The last two sections are devoted to examples. In section 6, we consider a couple of simple explicitly or almost explicitly solvable examples, which demonstrate in some detail how our method works in practice. We show in particular that for a potential barrier we can describe the transition from transmission to reflection near the critical energy in a uniform way. In section 7, we reconsider the situation from [25] and demonstrate how our method reproduce the results. As a particular application, this shows using [29] that in a strongly chaotic system coherent states become semiclassically equidistributed beyond the Ehrenfest time. Finally, we summarize our results in the conclusions. In the appendix, we collect some more technical results from the semiclassical analysis that we use in the main body of the work.

2. Time-dependent WKB theory

Our method is an extension of the well-known time-dependent WKB method for real-valued phase functions. So we start by recalling this method, and to keep the discussion simple, we restrict ourselves first to the case of the standard Schrödinger equation with the potential V in $\mathbb{R}^n$, which in appropriate units reads

Equation (9)

In section 5, we will discuss a more general setting. We want to solve this equation for an initial state of the form $\psi =A_0\mathrm{e}^{\frac{\mathrm{i}}{\hbar }S_0}$, where S0 is real valued. Inserting the usual WKB ansatz

Equation (10)

with S(t, x) being real valued, into the Schrödinger equation gives an expression that we separate, following the standard treatment [5], into the following two equations:

Equation (11)

Equation (12)

If A does not depend on ℏ, then this is a splitting into different powers of ℏ. Equation (11) is the Hamilton–Jacobi equation and its solutions are described using the propagation of the Lagrangian manifolds defined by p = ∇S(t, q):

Equation (13)

which are transported by the Hamiltonian flow associated with the classical Hamiltonian $H(p,q)=\frac{1}{2}p^2+V(q)$. We will analyse this further in section 4. The phase function S(t, x) often exists only on a finite time interval and develops singularities if caustics appear, then one has to use a refined ansatz. Because of the association with Lagrangian submanifolds WKB states of the form (10) are often called Lagrangian states in the mathematical literature [30, 31].

If we ignore in (12) the term of order ℏ on the right-hand side, then this is a transport equation that describes how the amplitude A is transported along the Lagrangian manifold Λt. Let us define the transport operator T(t) as the solution to the transport equation

Equation (14)

where the operator in brackets on the right-hand side is self-adjoint, and hence T(t) is unitary. Hence, if we make for A(t) an ansatz

Equation (15)

with an operator D(t) that is to be determined, and insert this into (12), then we obtain for D(t) the equation

Equation (16)

where

Equation (17)

and with the initial condition D(0) = I. Since Δ(t) is self-adjoint, D(t) is unitary as well. From a more detailed analysis of the operator T(t) in section 4.2, we will learn that Δ(t) is a generalized Laplacian of the form

Equation (18)

see (49). Collecting all terms we have a solution to the original Schrödinger equation of the form

Equation (19)

So the time evolution is described by three parts: (i) propagation of the Lagrangian manifold Λt, (ii) classical transport of the amplitude by the unitary operator T(t) and (iii) quantum dispersion described by D(t). Note that both T(t) and D(t) depend on the initial state via the initial phase function S0.

The main condition we need in order that (19) holds on a time interval [0, T] is that the projection of Λt to position space is smooth for all t ∈ [0, T], i.e. Λt does not develop caustics. Under this condition, the representation (19) is exact, and S(t) and T(t) are both determined by transport along classical trajectories in phase space, the only contribution not yet linked to classical dynamics is the dispersive part D(t). We obtain the usual WKB result by neglecting the dispersion, i.e. by replacing D(t) by the identity I. To see what kind of error we make by doing this, we can use Duhamel's principle that we will also state in the general form for later use [20]. Let $\hat{H}(t)$ and $\hat{H}_1(t)$ be two time-dependent self-adjoint operators, and U(t) and U1(t) be the time evolution operators generated by them with the initial conditions U(0) = U1(0) = I; then,

Equation (20)

This formula allows us to estimate how close the time evolutions generated by two different Hamiltonians are to each other. If we choose $\hat{H}(t)=-\frac{\hbar ^2}{2}\Delta (t)$ and $\hat{H}_1=0,$ then U(t) = D(t) and U1(t) = I, and (20) gives

Equation (21)

Since D is unitary, we then directly obtain

Equation (22)

So if the second-order derivatives of A0 are bounded and the coefficients of Δ(t) are not growing with t, then the solution to Schrödinger's equation satisfies:

Equation (23)

which is the standard time-dependent WKB result. Furthermore, Duhamel's formula can be iterated and gives an expansion with an error term of order O((ℏ|t|)N) if the first 2N derivatives of A0 are bounded and the coefficients of Δ(t) do not grow with t [20].

3. Metaplectic extension of WKB

We would like to apply (19) to a coherent state, i.e. a state of the form (2). With S0(x) = p · (xq) and $A_0(x)=(\pi \hbar )^{-n/4}\,\mathrm{e}^{-\frac{1}{2\hbar }(x-q)\cdot B(x-q)}$, this state is of the form that we can use in (19), but now

Equation (24)

so (22) does not allow us to use the standard WKB approximation D(t) ≈ I.

The way to solve this problem is to approximate the action of D(t) on A0 not by the identity, as in the WKB method, but to borrow from the standard propagation of coherent states and approximate the generator of D(t) by its Taylor expansion around the centre of A0 up to second order. Since D(t) acts on a state that is concentrated in position space at x = q and in momentum space at p = 0, this means that we freeze the coefficients of Δ(t) at x = q and approximate D(t) by the operator generated by it.

To formalize this idea, we introduce for $q\in \mathbb{R}^n$ the unitary operator Lq as

Equation (25)

which shifts by q and then rescales in ℏ, so that (Lqa)(x) is concentrated around x = q, e.g., if a(x) = πn/4 exBx/2, then A0 = Lqa is the amplitude of the coherent state (2) considered above. Now inserting A = Lqa into (16) gives for a the equation

Equation (26)

and we will approximate the operator on the right-hand side by

Equation (27)

From (18), one finds that Δq(t) = ∑αij(t, q)∂ij and

Equation (28)

in the sense that $\hat{Q}_q(t)a(x)=O_t(\sqrt{\hbar })$ if a is smooth and has bounded derivatives. So the operator Δq(t) is indeed obtained from Δ(t) by freezing the coefficients at x = q and furthermore discarding the first-order terms. In (52), we will obtain more explicit bounds on (28) in terms of the classical flow.

Therefore, if we define the operator Dq(t) by

Equation (29)

then we expect that D(t)LqaLqDq(t)a holds. To check this, we use again Duhamel's principle (20)

Equation (30)

and, since D(t) and Lq are unitary, we obtain from (28)

Equation (31)

The t-dependence in the remainder term will be governed by the behaviour of $\hat{Q}_q(t)$. In the next section, we give conditions on the initial state and on the classical dynamics under which the coefficients of $\hat{Q}_q(t)$ stay bounded, see (51) and (52).

To summarize our results so far, if $\psi _0=A_0\,\mathrm{e}^{\frac{\mathrm{i}}{\hbar }S_0}$, where A0 = Lqa is concentrated around q, then the time-evolved state is given by

Equation (32)

so we can use the standard time-dependent WKB to propagate coherent states if we include the additional operator Dq(t). This operator has a quadratic generator and is therefore a metaplectic operator, see [3, 32, 33], and so we call it the metaplectic correction to the standard time-dependent WKB method. In order to understand the behaviour of Dq(t) in some more detail, we have to look at the propagation of the Lagrangian manifold Λ by the classical dynamics, and the associated geometric interpretation of our extended time-dependent WKB scheme. This will be the subject of the next section.

We finally note that it is sometimes useful to work with Mq(t) := LqDq(t)L*q, because then LqDq(t)a = Mq(t)Lqa = Mq(t)A0 and we can allow for A0 to be of a more general form than the one induced by the specific scaling with Lq, as long as it is concentrated around x = q. Note that Mq satisfies the equation

Equation (33)

and the time evolution of an initial state $\psi _0=A_0\mathrm{e}^{\frac{\mathrm{i}}{\hbar }S_0}$, where A0(x) is concentrated around x = q, can be approximated by

Equation (34)

4. Geometric interpretation

In order to understand the scope and the accuracy of the metaplectic extension of time-dependent WKB theory, which we outlined in the previous section, we have to understand some of the geometry underlying it. We will see that in particular the time dependence of the remainder term is governed by the way in which the classical flow transports the initial Lagrangian manifold Λ. Furthermore, a proper understanding of the geometry will allow us to extend the scheme beyond caustics.

4.1. Transport of the Lagrangian manifold

Let us first return to the solutions of the Hamilton–Jacobi equation (11), which we will discuss for a general Hamiltonian H(ξ, x):

Equation (35)

As is well known, there are two main ingredients involved in the solution; see, e.g., [5, 30, 31]. The first is the transport of the initial Lagrangian manifold

Equation (36)

where $U\subset \mathbb{R}^n$ is an open set that contains the support of the amplitude A(x). Let us denote by Φt(ξ, x), where $(\xi ,x)\in \mathbb{R}^n\times \mathbb{R}^n$, the Hamiltonian flow generated by H, i.e. Φt(ξ, x) = (p(t), q(t)), where (p(t), q(t)) are the solutions to Hamilton's equation $\dot{p}=-\nabla _qH(p,q)$ and $\dot{q}=\nabla _p H(p,q)$ with initial conditions p(t = 0) = ξ and q(t = 0) = x, respectively. Using this flow, we can transport the Lagrangian manifold Λ and obtain a family of Lagrangian manifolds

Equation (37)

We will call the initial Lagrangian manifold Λ non-contracting (with respect to the flow Φt) if there exists a C > 0 such that for all t > 0

Equation (38)

Here, $\mathrm{d}\Phi ^t|_{\Lambda }(z):T_z\Lambda \rightarrow T_{\Phi ^t(z)}\Lambda _t$ is the restriction of the linearized flow at z = (p, q) to Λ, which is given in local coordinates by the matrix of the derivatives of the components of Φt with respect to the coordinates on Λ. This condition means that, roughly speaking, trajectories starting nearby on Λ do not coalesce into each other. An example for a non-contracting submanifold is the unstable manifold of a hyperbolic trajectory, and more generally, if a system is hyperbolic, then any submanifold that is transversal to the stable foliation is non-contracting. Stable manifolds are then of course examples of manifolds which are not non-contracting. In an integrable system, any manifold that lies in the regular part of the foliation of phase space into invariant tori is non-contracting.

The second ingredient needed to determine S(t, x) is the projection of Λt to position space along the momentum directions, i.e. we take the projection $\pi :\mathbb{R}_p^n\times \mathbb{R}_q^n\rightarrow \mathbb{R}_q^n$ defined by π(p, q) = q and restrict it to Λt. If this map,

Equation (39)

has no singularities, i.e. π(Λt) = Ut does not contain any caustics of Λt, then there exists, at least locally, a phase function S(t, x) such that

Equation (40)

for some open set $U_t\in \mathbb{R}^n$. By assumption, $\pi _{\Lambda _0}:\Lambda _0\rightarrow U_0\subset \mathbb{R}^n$ is smooth so we can invert it, with inverse given by $\pi _{\Lambda _0}^{-1}(x)=(\nabla S(0,x),x)$, and hence, we can form

Equation (41)

which is a map from UUt, and if $\pi _{\Lambda _t}$ is smooth, it is a smooth and invertible map. But note that ϕΛ(t) is not a flow. Analogously to the above notion of non-contracting, (38), we will say that ϕΛ(t) is non-contracting on $U\subset \mathbb{R}^n$ if there exist C > 0 such that

Equation (42)

independent of t ⩾ 0, where ϕ'Λ(t, x) denotes the matrix of first-order derivatives of the components of ϕΛ(t, x). A necessary condition for ϕΛ to be non-contracting is that Λ0 is non-contracting, but in addition we need that the projection $\pi :\Lambda _t\rightarrow \mathbb{R}^n$ is not singular. This implies that we have to stay away from caustics.

4.2. The transport operator

Another description of the map ϕΛ(t, x) is as follows: take the trajectory (p(t), q(t)) = Φt(ξ, x), which starts at t = 0 at q(t = 0) = x with the initial momentum ξ = ∇S(x); then, the map ϕΛ(t, x) is the q component of this trajectory, ϕΛ(t, x) = q(t). Since by Hamilton's equations $\dot{q}=\nabla _p H(p,q)$ and on Λt we have p = ∇S(t, q), we find that ϕΛ(t, x) is the unique solution to

Equation (43)

Note that we allowed here for a general Hamiltonian, and not only one of the simple form kinetic plus potential energy.

We will now show that the transport operator T(t) defined in (14) can be written using this map as

Equation (44)

By comparing this formula with appendix A.2 and using (43), we see that the operator in (44) satisfies

Equation (45)

where $\hat{K}(t)$ is the Weyl quantization of the function

Equation (46)

In particular, we see that for the special case $H=\frac{1}{2}p^2+V(q),$ we have from appendix A.2

Equation (47)

and comparing with the transport equation (12) proves our claim.

4.3. The metaplectic correction

We can use this observation to give a more detailed description of the operator

Equation (48)

This is an application of Egorov's theorem and the technical details are given in appendix A.3.2, where we show that the Weyl symbol of this operator is given by

Equation (49)

where ${\mathcal A}(t,x)$ is a symmetric matrix given by

Equation (50)

and (ϕ−1Λ)i'' is the matrix of second derivatives of the ith component of ϕ−1Λ. Bounds on the coefficients of the operator Δ(t) play an important role in estimating the accuracy of the time-dependent WKB propagation and the metaplectic extension, because Δ(t) appears in the error terms (22) and (31). This is where the non-contraction condition that we introduced above becomes important; from (42), we have that if ϕΛ(t) is non-contracting, then

Equation (51)

and hence, the operator Δ(t) has bounded coefficients. We can now also give a more explicit description of the operator Δq(t). A short calculation shows that for a phase space function H(ξ, x), which is quadratic in the momentum ξ, we have $\hbar L_q^*\hat{H}L_q =\hat{H}_q$, with $H_q(\xi , x)=H(\xi , q+\sqrt{\hbar }\, x)$ and so $ H_q(\xi , x)=H_q(\xi , q)+O(\sqrt{\hbar })$. Applying this to $\delta H(t,\xi ,x)=\frac{1}{2}\xi {\mathcal A}(t,x) \xi +O(\hbar ^2)$ gives that

Equation (52)

If ϕΛ(t) is non-contracting, then the remainder is bounded uniformly in time.

Since we have now an explicit expression for Δq(t), we can compute the action of the metaplectic operator Dq(t), defined in (29), on a function a. Set

Equation (53)

and let $\hat{a}(\xi )=\int \mathrm{e}^{-\mathrm{i}x\xi } a(x)\, \mathrm{d}x$ be the Fourier transform of a, and then,

Equation (54)

Similarly, we find for the rescaled operator Mq(t) that integrating (33) gives

Equation (55)

where $\hat{A}_{\hbar }(\xi )=\int \mathrm{e}^{-\frac{\mathrm{i}}{\hbar } x\xi } A(x)\, \mathrm{d}x$. The operators Dq(t) and Mq(t) act as the Fourier multipliers with Gaussian functions. For an initial amplitude of the form $A(x)=(\pi \hbar )^{-n/4}\,\mathrm{e}^{-\vert x\vert ^2/(2\hbar )}$, one finds in particular

Equation (56)

4.4. The role of the initial manifold

The localized initial states that we consider are of the form $\psi =(L_q a)\,\mathrm{e}^{\frac{\mathrm{i}}{\hbar } S}$, and the phase function determines the initial Lagrangian manifold, which is crucial for the semiclassical propagation scheme that we presented. The question we want to consider now is if the state ψ determines the Lagrangian manifold uniquely, or in other words, if there might be other functions $\tilde{a}$ and $\tilde{S}$, such that $\psi =(L_q\tilde{a})\,\mathrm{e}^{\frac{\mathrm{i}}{\hbar }\tilde{S}}$. In that case, we would have some freedom in the choice of the initial Lagrangian manifold we use for propagation.

The condition $(L_q a)\,\mathrm{e}^{\frac{\mathrm{i}}{\hbar } S}=(L_q\tilde{a})\,\mathrm{e}^{\frac{\mathrm{i}}{\hbar }\tilde{S}}$ gives after multiplication by $\mathrm{e}^{-\frac{\mathrm{i}}{\hbar }\tilde{S}}$ and application of L*q that

Equation (57)

and by the Taylor expansion, we see that

Equation (58)

where R(ℏ, q, x) is a smooth real-valued function. The first term on the right-hand side gives just a constant phase factor, so if

Equation (59)

then $\tilde{a}$ defines a nice smooth function. But this condition means that the two Lagrangian submanifolds Λ and $\tilde{\Lambda }$, generated by S and $\tilde{S}$, respectively, intersect at (p, q) = (∇S(q), q). We therefore conclude that in order to propagate a state localized around a phase space point (p, q), we can use any Lagrangian manifold through that point that is non-contracting. This freedom of choice can be used to select, for instance, initial manifolds for which the propagation becomes particularly easy, e.g., some for which no caustics develop.

In the case that the trajectory through (p, q) is hyperbolic, any Lagrangian submanifold that is transversal to the stable directions is non-contracting, and hence can be used for propagation. Although it seems most natural to take the unstable manifold, any other manifold that is sufficiently transversal to stable direction will converge exponentially fast to the unstable manifold and hence should work with comparable efficiency. We test this with the example in section 7.

4.5. Relation to the standard coherent state propagation and time scales

For times that are short compared to the Ehrenfest time, our approach should reproduce the standard results on coherent state propagation. For the general class of initial amplitudes of the form (25), the propagation is reviewed in [34]. A propagated coherent state is of the form

Equation (60)

where (p(t), q(t)) is a classical trajectory, l(t) is an action-type phase that also includes the Maslov terms, and $a_{\hbar }(t,x)=a_0(t,x)+\sqrt{\hbar }\, a_{1}(t,x)+\cdots$ with leading term given as the solution to

Equation (61)

This means that the centre of the state is propagated along the trajectory (p(t), q(t)) and its shape changes according to the linearized dynamics around the centre.

If we expand S(t, x) around x = q(t) up to second order, then we find that we can write our approximation (32) in the form (60) with

Equation (62)

Now, using the operator identities

Equation (63)

Equation (64)

Equation (65)

Equation (66)

which all follow by direct computation, we obtain with $\nabla S(t,q(t))=p(t)=\dot{q}(t)$ that

Equation (67)

where S and its derivatives are all evaluated at x = q(t). If we furthermore use $\Delta ( b(t,x)\mathrm{e}^{\frac{\mathrm{i}}{2}x\cdot S^{\prime \prime }x})=[\Delta b(t,x)+2\mathrm{i}\nabla b(t,x)\cdot S^{\prime \prime } x + (-\vert S^{\prime \prime }x\vert ^2+\mathrm{i}\Delta S)b(t,x)]\,\mathrm{e}^{\frac{\mathrm{i}}{2}x\cdot S^{\prime \prime }x}$ and combine this with (67), we find

Equation (68)

Finally, from the Hamilton–Jacobi equation (11), we obtain $-x\cdot \dot{S}^{\prime \prime } x-\vert S^{\prime \prime } x\vert ^2/2=x\cdot V^{\prime \prime } x$, and hence, the leading term a0(t, x) satisfies (61).

This was a formal computation that showed that our result reproduces the previously existing results, but did not gave much insight as to where the standard approximation might break down. To see this more clearly let us just look at the amplitude T(t)Lqb with b = Dqa. By the Taylor expansion around x = q(t) = ϕΛ(t, q), we have ϕ−1Λ(x) − q ≈ [ϕΛ'(t, q(t))]−1(xq(t)), and hence,

Equation (69)

From this expression, we see that the state will be localized around x = q(t) if

Equation (70)

and since ϕΛ(t, x) is a projection of the flow from phase space, this is satisfied if tTE(ℏ). On the other hand, if $\sqrt{\hbar }\Vert \phi _{\Lambda }^{\prime }(t,q(t))\Vert \approx 1$, then the amplitude is no longer localized around a point, and the state becomes a Lagrangian or a WKB state.

5. A generalized propagation scheme and caustics

So far we have avoided the discussion of caustics, but although there are situations where no caustics occur, in many interesting situations we expect caustics to develop. A caustic is the image in position space $\mathbb{R}^n$ of the singularities of the projection $\pi _{\Lambda _t}:\Lambda _t\rightarrow \mathbb{R}^n$, which means that at a caustic the tangent plane TzΛt to Λt at a point z = (p, q) ∈ Λt becomes vertical in the sense that TzΛtVz ≠ {0}, where $V_z=\lbrace (\xi ,q)\,\, ;\, \xi \in \mathbb{R}^n\rbrace$ denotes the vertical subspace. In a neighbourhood of a caustic, the manifold Λt can no longer be represented by a generating function S(t, x) as in (40), and so the simple time-dependent WKB propagation we presented above can no longer work. This problem is usually solved by switching to a different representation of the quantum state, e.g., taking the Fourier transform switches from the position representation to the momentum representation, and in this new representation, we have to consider instead the projection of Λt to momentum space.

We want to use a method that allows for a bit of flexibility so that we can accommodate all possible different representations. To this end, we look at how we can in general approximate the full Hamiltonian H by a simpler Hamiltonian H1 in a way that the propagation of Lagrangian states initially associated with Λ can be approximated using H1. The crucial observation that will guide our conditions on H1 is that if B(p, q) and A(x) are sufficiently smooth functions and $\hat{B}$ is the Weyl quantization of B, then

Equation (71)

where BΛ(x) = B(∇S(x), x); this is a classical result, see, e.g, [31]. This means that the Lagrangian state is concentrated in phase space on Λ; in particular, if BΛ = 0, then $\hat{B} \psi =O(\hbar )$. More generally, if B vanishes of order N on Λ, i.e. (∂αB)|Λ = 0 for |α| ⩽ N, then

Equation (72)

Since a propagated Lagrangian state is concentrated near the propagated Lagrangian manifold Λt = Φt(Λ), where Φt is the Hamiltonian flow generated by H, we expect that if H1 is close to H near Λt, then it will provide an accurate approximation for the purpose of propagating states concentrated near Λt. The conditions for a function H1 to be a general first-order approximation of H near Λt are

Equation (73)

These conditions ensure that the Hamiltonian vector fields generated by H and H1 agree on Λt; hence, if we denote by Φt1 the time t map generated by H1, then

Equation (74)

i.e. the classical dynamics agree when restricted to Λ.

Let us now turn to the quantizations of H and H1. Let U(t) and U1(t) be the time evolution operators generated by $\hat{H}$ and $\hat{H}_1$, respectively; then, we expect from (74) that U(t)ψ ≈ U1(t)ψ for a Lagrangian state associated with Λ. To verify this, we make an ansatz

Equation (75)

and inserting this into the Schrödinger equation for U(t) gives for V(t) the equation

Equation (76)

where

Equation (77)

By Egorov's theorem, see appendix A.3, the operator $\widehat{\delta H}(t)$ is, in leading order in ℏ, the quantization of HH1 transported along the map Φt1:

Equation (78)

This implies from (73) that

Equation (79)

and therefore, from (72), we have for an initial state of the form $\psi =A_0\,\mathrm{e}^{\frac{\mathrm{i}}{\hbar }S_0}$, where A0 has bounded derivatives, that

Equation (80)

Thus, the generator of V(t) is small when applied to ψ, and therefore, Duhamel's principle (20) gives

Equation (81)

and hence, U1(t)ψ is a good approximation for U(t)ψ if ψ is a Lagrangian state.

This is a generalization of the standard time-dependent WKB method presented in section 2. The choice of H1 corresponding to the results in section 2 is

Equation (82)

which we obtain from H(p, q) = H(∇S(t, x) + ξ, x) as a first-order approximation in the shifted momentum ξ = p − ∇S(t, x). For a Hamiltonian of the form $H=\frac{1}{2}p^2+V(q),$ one finds using appendix A.2 that $\hat{H}_1=H(\nabla S(t,x),x)-\mathrm{i}\hbar \nabla S(t,x)\cdot \nabla -\frac{\mathrm{i}\hbar }{2}\Delta S(t,x)$, and so by comparing this with (14) and (11), we see that in this case we have

Equation (83)

So how does a caustic affect this relation? In defining H1 we used coordinates (ξ, x) near Λt defined by p = ∇S(t, x) + ξ and q = x, so that Λt = {ξ = 0}, but at a caustic this coordinate system becomes singular. In order to repair this we can just use a different canonical coordinate system near Λt, one which does not become singular, and choose for H1 the first-order Taylor approximation in the transversal coordinates.

The only problem we might encounter with this more general version of time-dependent WKB theory is that the control of the time dependence of the remainder term (80) becomes more involved. In section 2, we were able to take advantage of the very explicit version of Egorov's theorem from appendix A.3.2 to reduce this problem to the non-contractiveness of the transport map ϕΛ(t). For different choices of H1, this problem can become much harder and we reserve a more thorough investigation for the future. But this question concerns our ability to obtain explicit error estimates; it does not prevent us from using, after suitable testing, the method for explicit numerical propagation in concrete problems.

Now, if the initial state ψ0 is localized in phase space around a point (p, q) ∈ Λ0 so that the estimate (80) no longer holds, then it is natural to approximate V(t) by taking only the behaviour of its generator δH(t, ξ, x) near (p, q) into account. From (73), we have δH(t, p, q) = 0 and ∇δH(t, p, q) = 0, and hence, the simplest nontrivial approximation is quadratic, with z = (ξ, x) and z0 = (p, q), and we have

Equation (84)

The time evolution $\tilde{M}_{p,q}(t)$ generated by $\widehat{\delta H}_2(t)$ via $\mathrm{i}\hbar \partial _t M_{p,q}(t)=\widehat{\delta H}_2(t) M_{p,q}(t)$ and Mp, q(0) = I is a metaplectic operator since the generator $\widehat{\delta H}(t)$ is the quantization of a quadratic function on phase space. Therefore, for ψ strongly localized around (p, q), we expect V(t)ψ ≈ Mp, qψ, and hence, we have arrived now at the more general version of (34), which reads

Equation (85)

if ψ is concentrated in phase space at (p, q) as in (25).

The relation of Mp, q(t) to Mq(t) follows from (83); if H1 is of the form (82), we find that

Equation (86)

where p = ∇S0(q) and $S_0^{(2)}(x)=p(x-q)+\frac{1}{2}(x-q)S_0^{\prime \prime }(q)(x-q)$ is the quadratic part of the Taylor expansion of S0 around x = q.

Since Mp, q(t) has a quadratic generator, it is the quantization of a linear symplectic map Pp, q(t) on $T_{p,q}(\mathbb{R}^n\times \mathbb{R}^n),$ and by construction, this map can be expressed in terms of the linearizations of the maps Φt and Φt1. Since V(t) = U−11(t)U(t), we can view V(t) as a quantization of the map (Φt1)−1○Φt, and Mp, q(t) is then the quantization of the linearization of that map around p, q:

Equation (87)

Since Φt and Φt1 are identical on Λ, we have that $P_{p,q}(t)|_{T_{p,q}\Lambda _0}=I$, and hence, Pp, q is a shear relative to the tangent space Tp, qΛ0 of the initial Lagrangian manifold at (p, q).

In case that H1 is given in (82), the map Pp, q(t) can be described in some more detail. Note that since H1 in (82) contains only a linear term in the momentum, we find that Hamilton's equations generated by H1 give for x(t) the equation

Equation (88)

and since the right-hand side does not depend on the momentum ξ, the trajectory x(t) does not depend on the initial momentum ξ. Hence, Φt1 maps vertical subspaces into vertical subspaces, i.e. $\mathrm{d}\Phi ^t_1 (V_{z_0})\subset V_{z(t)}$, where for z = (p, q), we set $V_z=\lbrace (\xi ,q)\, ,\xi \in \mathbb{R}^n\rbrace$. Therefore, the map Pp, q(t) satisfies

Equation (89)

and as we show in appendix B, this implies that the knowledge of the Lagrangian subspace $(\mathrm{d}\Phi ^t)^{-1}(V_{\Phi ^t(p,q)})$ determines the map Pp, q(t)−1 and hence Pp, q(t) uniquely.

Let us look at two examples where we can determine the long time limit of $(\mathrm{d}\Phi ^t)^{-1}(V_{\Phi ^t(p,q)})$ from dynamical conditions.

  • (a)  
    Assume we have a one-dimensional system with an x-independent Hamiltonian H(p), which satisfies ∂2H/∂p2 > 0. Then, the velocity increases with p and hence $ (\mathrm{d}\Phi ^t)^{-1}(V_{\Phi ^t(p,q)})\rightarrow \lbrace (0,q), q\in \mathbb{R}\rbrace$ for t. Therefore, if $T_{(p,q)}\Lambda _0\ne \lbrace (0,q), q\in \mathbb{R}\rbrace ,$ we can use the result from appendix B and see that for large t the map Pp, q(t) tends to a limit Pp, q, and hence, the corresponding metaplectic correction Mp, q(t) tends to a limit, too.
  • (b)  
    If the trajectory through (p, q) is hyperbolic and the vertical subspaces Vz are transversal to the unstable subspaces Vuz along the trajectory, then $ (\mathrm{d}\Phi ^t)^{-1}(V_{\Phi ^t(p,q)})$ tends to the stable subspace Vsp, q for large t, and at an exponential rate if the hyperbolicity is uniform. So in this situation, we find that for large t the map Pp, q(t) tends to a limit Pp, q, which is uniquely defined by the conditions that Pp, q|Tp, qΛ0 = I and Pp, q(Vsp, q) = Vp, q. Therefore, in this situation, the metaplectic correction tends for large t to a limit, too, and exponentially fast if the hyperbolicity is uniform.

6. Examples

In order to illustrate how the theory that we developed in the previous sections works, it is very instructive to look at a couple of simple examples. They will allow us to understand in more detail the interplay among the classical dynamics, the choice of an initial Lagrangian manifold and the analytical constructions we developed.

6.1. The free particle

The first example, we look at is the free particle in one dimension, with the Hamilton operator $-\frac{\hbar ^2}{2}\Delta$. We want to propagate a state that is initially localized at $(p,q)\in \mathbb{R}^2$, where p ⩾ 0. As an initial phase function, we choose

Equation (90)

where $\alpha \in \mathbb{R}$ is a real parameter. The corresponding Lagrangian manifold is

Equation (91)

which is a line through (p, q) with slope α. The corresponding classical Hamilton function is $H(\xi )=\frac{1}{2}\xi ^2$ and the Hamiltonian flow it generates is given by Φt(ξ, x) = (ξ, x + tξ). Applying the flow to Λ0 gives $\Lambda _t=\Phi ^t(\Lambda _0)=\lbrace (p+\alpha x, q+x +t(p+\alpha x)), x\in \mathbb{R}\rbrace ,$ and by replacing x with x/(1 + αt), this can be rewritten as

Equation (92)

which is a line through (p, q(t)) = Φt(p, q) with the slope α(t). Note that if α < 0, i.e. if the initial line has a negative slope, then α(t) → for t → −1/α; this means that the line Λt turns vertical, and hence we have a caustic. But if α ⩾ 0, then α(t) ⩽ α for all t ⩾ 0, and no caustics occur. The phase function that generates Λt and satisfies the Hamilton–Jacobi equation with the initial condition S0 is given by

Equation (93)

To find the transport operator, we have to find the map $\phi _{\Lambda }(t,x)=\pi _{\Lambda _t}\Phi ^t\big(\pi ^{-1}_{\Lambda _0}(x)\big)$. To this end, we notice that $\pi ^{-1}_{\Lambda _0}(x)=(\nabla S_0(x),x)=(p+\alpha (x-q),x)$, and hence $\Phi ^t\big(\pi ^{-1}_{\Lambda _0}(x)\big)=(p+\alpha (x-q),x+t(p+\alpha (x-q)),$ and the final projection just takes the second component, therefore,

Equation (94)

Since ϕ'Λ(t, x) = (1 + αt), the condition α ⩾ 0 guarantees that ϕΛ is non-contracting and we obtain $\phi _{\Lambda }^{-1}(t,x)=\frac{1}{1+\alpha t} (x-t(p-\alpha q))$. Therefore, the action of the transport operator reads

Equation (95)

which for α ⩾ 0 is well defined for all t ⩾ 0. In order to compute the metaplectic correction, which is generated in (49), we note that ${\mathcal A}(t,x)=(1+\alpha t)^{-2}$ is independent of x, and hence, we have

Equation (96)

Then, the metaplectic correction (55) is

Equation (97)

where the time-dependent factor in the exponent comes from the integration of the time dependence in Δ(t), $\int _0^t\frac{1}{(1+\alpha t^{\prime })^2}\, \mathrm{d}t^{\prime }=\frac{t}{1+\alpha t}$. We see that in particular, as we observed at the end of section 5, that the operator tends to a limit for large t, $ M_q(t)\rightarrow M_q(\infty )=\mathrm{e}^{\frac{\mathrm{i}\hbar }{2} \frac{1}{\alpha }\Delta }$.

We computed all the elements in our extended time-dependent WKB propagation scheme for the free particle. Let us apply this to an initial state $\psi _0(x)=A_0(x)\mathrm{e}^{\frac{\mathrm{i}}{\hbar }[p(x-q)+\frac{\alpha }{2}(x-q)^2]}$ with A0 = Lqa0 for some $a_0\in S(\mathbb{R})$. Then, Mq(t)A0 = Lqat with at = Dq(t)a0 and since Dq(t) tends to a limit for large t, we have that $a_t\in S(\mathbb{R})$ with bounds uniform in $t\in \mathbb{R}^+$. Hence, we find

Equation (98)

with

Equation (99)

Since the Hamiltonian is quadratic, this expression is actually exact, and the remainder terms are zero. Note that the behaviour of the amplitude A(t, x) depends on the size of ℏ1/2(1 + αt), which determines the Ehrenfest time scale (8). If ℏ1/2(1 + αt) → 0, the amplitude becomes concentrated, but if ℏ1/2(1 + αt) ∼ 1, then the amplitude A(t, x) is a smooth function and ψ(t) actually becomes a Lagrangian state.

Let us note that for the special case that our initial state is a Gaussian coherent state,

Equation (100)

where $b\in \mathbb{C}$ with Im b > 0, our result is identical to the one obtained by the standard propagation of coherent states, since the Hamiltonian is quadratic. The only difference is that we derived it in a different way. But it was still instructive to go through the derivation since it highlights a few important points. We can rewrite the state using the phase function S0 from (90) as

Equation (101)

and then applying (98) and (99) gives us the propagated state. But since the initial state is independent of α, the final result is independent of the choice of α, too, i.e. of the initial Lagrangian submanifold used to propagate the state. But the intermediate steps depend on α, first of all, we needed to choose α ⩾ 0 in order to avoid caustics, and then the choice of α determined how we split the evolution into transport, implemented by the operator T(t), and dispersion, represented by D(t) = M0(t). The simplest choice would have been α = 0; then, Λ0 is horizontal, and actually invariant under the classical flow. In this case, T(t) is defined by transport along one central trajectory, and all the dispersion is described by D(t). If α > 0, then Λ0 is transversal to the flow and the action of T(t) actually takes a family of trajectories into account. Since these trajectories move with different speed, this already accounts for part of the dispersion of the wave packet, and the operator M(t) has only to add the dispersion transversal to Λt. This is reflected by the fact that the coefficients of Δ(t) tend to 0 for large t, and fast enough so that the limit

Equation (102)

actually exists if α > 0. This means that for large t we can replace M0(t) by its limit and still obtain a good approximation of the propagated state. This is why α > 0 is preferable to α = 0, in particular if we extend these constructions to more general systems.

In addition, the fact that for α > 0 the coefficients of Δ(t) tend to 0 for large t reflects the property that the map ϕΛ is expanding; this also has consequences for the size of the remainder term if we add higher order terms to the Hamiltonian. To study this effect will be the main focus of the next example.

6.2. Integrable systems

Let us look now at a more general Hamiltonian that is a function of the momentum only, but no longer necessarily quadratic, so H = H(ξ), and we will assume that for ξ > 0 we have H'(ξ) > 0 and H''(ξ) ⩾ 0. This is the form a Hamiltonian of an integrable system takes in action–angle coordinates.

The classical dynamics is given by Φt(ξ, x) = (ξ, x + tH'(ξ)), and by the condition H''(ξ) ⩾ 0, the velocity H'(ξ) cannot decrease with increasing ξ. This implies that if we take again an initial phase function of the form S0(x) = px + αx2/2, then the corresponding Lagrangian submanifold does not develop caustics for t ⩾ 0 if α ⩾ 0, We find $\Lambda _t=\Phi ^t(\Lambda _0)=\lbrace (p+\alpha x, x+tH^{\prime }(p+\alpha x)),x\in \mathbb{R}\rbrace$ and therefore

Equation (103)

and so if α > 0 and H'' > 0, then ϕΛ(t, x) is actually expanding. Since the flow is now no longer linear, Λt is no longer a straight line and we do not attempt to find an explicit expression for S(t, x), but rather take it as defined by the Hamilton–Jacobi equation. But we observe that since Λt = {(∇S(t, ϕΛ(t, x)), ϕΛ(t, x))}, we have ∇S(t, ϕΛ(t, x)) = p + αx = ∇S0(x), which is of course a consequence of momentum conservation.

Having the phase function and the map ϕΛ(t), and therefore, the transport operator T(t), we have the ingredients of the standard WKB propagation, and we now want to compute the metaplectic correction for an initial state localized at (p, 0). Here, we have to use the generalized formulation from section 5 since the Hamiltonian is no longer a sum of a quadratic kinetic energy term and a potential energy term. The transport operator is the quantization of the map Φt1(ξ, x) = ([ϕ'Λ(t, x)†]−1ξ, ϕΛ(t, x)) and so the operator $\widehat{\delta H}$ has the symbol δH(t, ξ, x) = δH(0)(t, Φt1(ξ, x)) + Ot(ℏ2), where δH(0)(t, ξ, x) = H(∇S(t, x) + ξ) − H(∇S(t, x)) − ξ · ∇H(∇S(t, x)) and the Ot(ℏ2) term comes from Egorov's theorem. Therefore, using ∇S(t, ϕΛ(t, x)) = ∇S0(x), we find

Equation (104)

Since the point (p, 0) corresponds to ξ = 0, x = 0, we find using (103) that the generator of the metaplectic correction is $\widehat{\delta H_2}(t)=-\frac{\hbar ^2}{2}\frac{H^{\prime \prime }(p)}{(1+\alpha t H^{\prime \prime }(p))^2} \Delta$ and hence

Equation (105)

We observe the same phenomenon as in the free case, and for α > 0, the large t behaviour has a limit. But now this is also reflected in the behaviour of the remainder term, which is determined by the size of

Equation (106)

and since the time integral of this quantity is bounded, we obtain for a coherent state ψ0 concentrated at (p, 0) that

Equation (107)

Here, we made the assumption that we can also control the remainder term Ot(ℏ2) from the application of Egorov's theorem, which seems likely since the transport operator has still a simple form. But we leave a detailed investigation to a future publication.

This is in contrast to the standard coherent state propagation that can only work if $\vert t\vert \sqrt{\hbar }\ll 1$, since $T_E=1/\sqrt{\hbar }$ for an integrable system. The physical reason for this is that for $t\sqrt{\hbar }\sim 1$ the state U1(t)M(t0 is no longer a coherent state but has become a Lagrangian state, i.e. there is a qualitative change in the nature of the state at the Ehrenfest time.

6.3. A parabolic barrier

Let us now look at a case with a hyperbolic trajectory, let $H(\xi ,x)=\frac{1}{2}\xi ^2-\frac{V_0}{2}x^2$; this Hamiltonian describes the motion in a parabolic barrier if V0 > 0. The classical flow is given by

Equation (108)

where $\lambda :=\sqrt{V_0}>0$ is the Liapunov exponent. The origin is a hyperbolic fixed point with stable and unstable manifolds given by

Equation (109)

Let us first choose an initial state localized on the fixed point, i.e. at (0, 0). An initial phase function S0(x) = α2x2/2 corresponds to $\Lambda _0=\lbrace (\alpha x,x)\, ,\, x\in \mathbb{R}\rbrace$ and the transported manifold Λt = Φt0) can be written as

Equation (110)

The corresponding phase function is S(t, x) = α(t)x2/2, and furthermore, we find

Equation (111)

and therefore,

Equation (112)

The manifold Λt does not develop caustics if α ⩾ −λ. The case α = −λ is special since then Λt = Vs, i.e. Λt = Λ0 is invariant and equal to the stable manifold. In this case, the manifold is contracting, but for all α > −λ the manifold is non-contracting. Since the case α = λ is also special, Λt = Vu for all t, and so the initial manifold is the unstable manifold; in all other cases, ΛtVu for large t. These different cases are also reflected in the behaviour of T(t) and M(t). For α = −λ, we have ϕΛ(x) = e−λtx, and hence,

Equation (113)

The transport operator T(t) is squeezing at an exponential rate, and the metaplectic correction has to make up for this at an exponential rate, too. In contrast, if α > −λ, we have

Equation (114)

and so T(t)A(x) = [ϕ'Λ(t)]−1/2AΛ−1(t, x)) is stretching at an exponential rate. The metaplectic correction is given by

Equation (115)

and tends exponentially fast to a limit. This dichotomy between the cases α = −λ and α > −λ is analogous to the one between α = 0 and α > 0 that we found for the free particle. For α > −λ, the metaplectic correction saturates in time, whereas for α = −λ, its contribution grows exponentially. Since the Hamiltonian is quadratic, both cases give exact solutions, and the different initial manifolds only correspond to different ways to split the time evolution. But as in the integrable case, if we move to a perturbation, the contracting case α = −λ gives remainder terms that blow up, whereas the expanding cases α > −λ give remainder terms that remain bounded by $\sqrt{\hbar }$ independent of time.

So far we have looked at an initial state that is concentrated on top of the barrier; let us now look at a initial state localized at (p, q) with q < 0 and p > 0, hence an incoming state. Then the question of interest is if this state gets transmitted or reflected, and how one can describe the transition between reflection and transmission uniformly if one changes (p, q). We chose an initial phase function S0(x) = p(xq) + α(xq)2/2, which again for α > −λ gives an expanding initial manifold Λ0. For simplicity, we will choose α = λ, i.e. Λ0 = (p, q) + Vu; the general case α > −λ can be treated similarly. From computing Φt(∇S0(x), x), we find

Equation (116)

where q(t) = λ−1psinh (λt) + qcosh (λt), and therefore,

Equation (117)

For the phase function, we find

Equation (118)

and $\mathcal {L}(t)=\int _0^t (p\dot{q}-H(p,q))\, \mathrm{d}t$. We see that M(t) again tends with exponential speed to a limit $M^{\infty }= \mathrm{e}^{\frac{\mathrm{i}}{2}\frac{1}{2\lambda } \Delta }$. The behaviour of T(t)Lqa depends crucially on $\mathrm{e}^{-\lambda t}/\sqrt{\hbar }$; for small time, this parameter is large, and hence, the state is localized, but for

Equation (119)

i.e. on Ehrenfest time scales, the function T(t)Lqa is no longer strongly localized and the transported wave packet is actually a Lagrangian state associated with

Equation (120)

Where this state moves depends on q(t), and for long times, we have

Equation (121)

So the behaviour of q(t) depends on the value of p + λq. The case p + λq = 0 corresponds to initial conditions (p, q) in the stable manifold Vs of the fixed point at (0, 0), and so the trajectory runs into the fixed point and the Lagrangian state is for long times located on the unstable manifold Vu. The case p + λq > 0 gives a wave packet that is transmitted over the barrier and the case p + λq < 0 corresponds to a wave packet which is reflected. The strength of our approach is that we can describe the transition between these different cases in a uniform way given in (117). We studied a simple example here, but the method works for general potential barriers and allows us to describe the transition between reflection and transmission of time-dependent wave packets in a uniform way.

Such processes are of great importance in the theory of chemical reactions and we expect that our method could be of great use in the theoretical and quantitative description of time resolved chemical reactions, which are experimentally accessible in femto and atto chemistry [35, 36]. A first step will be to analyse the transport of wave packets over barriers, or more generally through bottlenecks in phase space, using the method of quantum normal forms recently developed in [37, 38].

7. Transport along a hyperbolic trajectory and its unstable manifold

In this section, we want to discuss the case that the initial coherent state is concentrated in a point z = (p, q) on a hyperbolic trajectory z(t) = Φt(z). As a test system, we choose the kicked harmonic oscillator

Equation (122)

where K is the chaoticity parameter that we will choose to be K = 2. This system has also been used in [25]. In figure 1(a), we show a portrait of the classical phase space, the system has an unstable fixed point at the origin and we display the unstable manifold. The Liapunov exponent of the unstable fixed point is λ = 0.83... and we use ℏ = 0.0008, so the Ehrenfest time is

Equation (123)
Figure 1.

Figure 1. (a) We show a phase space portrait of the KHO dynamics, i.e. equation (122), for K = 2. The origin is a hyperbolic fixed point, and we also displayed the corresponding unstable manifold. (b) We plot the Wigner function of the evolved state (124) at t = 4, which is of the order of the Ehrenfest time TE = 4.29.... It evolves along the unstable manifold (grey line) and starts resembling a WKB state. For comparison in the upper left corner, the Wigner function at t = 0 is shown; note that the vertical axis has been rescaled compared to (a).

Standard image

As an initial state, we choose a Gaussian wave packet centred on the fixed point at the origin:

Equation (124)

In figure 1(b), we display the Wigner function of the time-evolved state at t = 4, which can be seen to be stretched along the unstable manifold. As expected at the Ehrenfest time, the state becomes a WKB-type state associated with the unstable manifold. For comparison, the real part of the time-evolved wavefunction ψ(t) is plotted in figure 2(c).

Figure 2.

Figure 2. In the top panels, we test (130): the initial state (124) is propagated forward using the full propagator to t = 4 and propagated backwards using just time-dependent WKB, the result (symbol) is compared with the metaplectic approximation (full line): (a) amplitude and (b) phase derivative. We also show the initial state for reference (dotted lines in (a)). In the bottom panels, we compare the metaplectically extended WKB scheme with exact quantum propagation (circles). (d) A zoom of the tail of the wavefunction in (c) to show that agreement extends down to this scale and into the most nonlinear region.

Standard image

The prediction of the theory in section 3 is that, after a suitable metaplectic correction of the initial amplitude, the state can be propagated using the standard time-dependent WKB approximation. This means that if we decompose the initial state as

Equation (125)

then the evolved state is from (19) and (34) given by

Equation (126)

Here, S(t, x) is the solution of the Hamilton–Jacobi equation satisfying S(0, x) = S0(x). The operators T(t), D(t) and M0(t) also depend on the choice of S0(q). Even if according to section 4.4 this choice is, to some extent, arbitrary, the efficiency of the method may be sensitive to S0(x). We shall only consider the quadratic case $S_0(x)=\frac{\tan \theta }{2}\, x^2$, corresponding to a linear Lagrangian manifold p = tan (θ) q. We shall see that there is a wide range of initial manifolds that work well.

The metaplectic correction M0(t) is calculated by considering the linearized dynamics at the fixed point, i.e. we approximate cos (q) by 1 − q2 in (122). Within this approximation, the initial manifold remains a straight line through the origin for all times and the transport operator T(t) (14) becomes a scaling transformation:

Equation (127)

with κ(t) being a time-dependent factor (which can be calculated by following the linear evolution of the initial manifold). Accordingly, the generalized Laplacian (17) reads

Equation (128)

and the metaplectic correction is just

Equation (129)

Note that equation (126) is equivalent to saying that the sequence of operations

Equation (130)

must produce approximately a Gaussian state centred at 0, for a Gaussian initial amplitude a0(x). This is the first test we shall carry out, with the choice S0 = 0 corresponding to the initial Lagrangian manifold p = 0. First, we propagate the state (125) exactly during some time, and then propagate it back using time-dependent WKB. Figures 2(a) and (b) show the comparison of the exact result D(t) L0a0 with the metaplectic approximation M0(t) L0a0. In the case of the modulus, the agreement between the exact DL0a0 and M0L0a0 is perfect within visual resolution. There is a small difference between the phases, but this occurs in a region where the amplitude is very small. We also show a direct comparison between the exact and the metaplectically extended WKB propagated states in figures 2(c) and (d). The agreement is excellent.

We will now investigate how robust the method is with respect to changing the choice of the (linear) initial WKB manifold. We have changed the slope θ in a range of almost 90°, from θ = −0.30π/2 to θ = 0.65π/2. The comparison between exact and metaplectically extended WKB at t = 4 is shown in figure 3. Except for very slight differences (blue triangles), the agreement is still excellent for these 'large' slopes.

Figure 3.

Figure 3. We compare the effect of different initial Lagrangian manifolds p = tan θ q on the propagation, we considered for θ the following fractions of π/2: −0.30, 0.35, 0.65. (a) We display |DL0a0|(q) (circles) and the metaplectic approximations |M0L0a0|(q) (lines). (b) We show the evolved state at t = 4 (real part). We compare metaplectically extended WKB (symbols: black dots and blue triangles) with exact quantum propagation (line), two slopes where considered: 0.65π/2 and −0.30π/2.

Standard image

As we increase the slope, the amplitude concentrates in narrower regions. This happens because, as the slope is large, the map induced in the q-coordinate by the WKB manifold is more expansive and T* is more contracting.

We tested the theory also for different values for ℏ and found good agreement with the expected behaviour (not shown).

8. Conclusions

In this paper, we derived an extension of the standard time-dependent WKB method that can be applied to highly localized states like coherent states. It allows us to describe in a uniform way the transition in time from a semiclassically highly localized coherent state to a delocalized Lagrangian state, which takes place at the Ehrenfest time.

The main idea on which this extension of time-dependent WKB theory is built is an exact decomposition of the time evolution of an initial state of the form $\psi _0(x)=A_0(x)\,\mathrm{e}^{\frac{\mathrm{i}}{\hbar }S_0(x)}$, where S0 is real valued, into several parts

Equation (131)

Here, S(t) is a solution of the Hamilton–Jacobi equation and is hence related to the transport of the Lagrangian manifold Λ0 generated by S0 through phase space. The unitary operator T(t) transports functions in position space along the projections of the phase space trajectories emanating from Λ0. Finally, D(t) is the propagator generated by the time-dependent Hamiltonian −ℏ2T*(tT(t)/2. S(t) and T(t) are defined purely in terms of transport along classical trajectories and D(t) takes into account the dispersive effect of quantum mechanics.

The standard time-dependent WKB approximation is obtained by approximating D(t) ≈ I; this works fine if the amplitude A0 is sufficient flat, i.e. has bounded derivatives. But for a coherent state, A0 is strongly localized around a point x = q; then, it is more natural to freeze the coefficients of the generator of D(t) at x = q, and the resulting operator is a metaplectic operator Mq(t) whose action on functions can be computed quite easily. The main result of this paper is thus

Equation (132)

which is valid for amplitudes A0 that are strongly localized around x = q. In order to justify the validity of this approximation for long times, particularly on Ehrenfest time scales, we had to analyse the underlying classical dynamics more carefully. We introduced a non-contraction condition on the position space trajectories emanating from a neighbourhood of the initial state with momenta p = ∇S(x), and if this condition holds our propagation scheme is effective. The non-contraction condition excludes caustics, but we have a large freedom in the choice of the initial phase function S(x), which in many cases allows us to avoid caustics, at least until the state becomes delocalized.

For times shorter than the Ehrenfest time, our scheme reproduces the standard coherent state propagation results based on a Taylor expansion of the Hamiltonian around the centre trajectory. But for times of the order of the Ehrenfest time, we find that the state becomes extended and a Lagrangian state, which in the chaotic case, is supported by the unstable manifold of the centre trajectory. From that time onwards, the standard time-dependent WKB theory applies, as has been observed in [25]. In particular, if the system is hyperbolic and mixing, one can apply the results from [29] to conclude that the state becomes equidistributed after the Ehrenfest time.

In order to extend the results to more general Hamiltonians, and to be able to include caustics, we noted that the standard time-dependent WKB approximation can be viewed as the exact quantum time evolution generated by a quantization of a first-order Taylor approximation of the Hamilton function around the Lagrangian manifold associated with the time-evolved WKB state. Based on this insight, we could now choose different first-order approximations, which remained valid for general Hamiltonians and at caustics. The price one has to pay is a more complicated and less explicit formalism. In addition, all the previous results were in principle rigorous, although we refrained from stating them in the form of theorems, but here we have to rely on an assumption that certain special cases of Egorov's theorem remain valid on Ehrenfest time scales. But we get a further benefit from this more general way to look at the time-dependent WKB approximation, we can show that the classical map associated with the metaplectic correction Mq(t) is a shear map on phase space and that it furthermore in many cases converges to a limit for large t, which can be determined from simple geometric considerations. This implies that in many cases

Equation (133)

and we can determine M()q up to a phase from simple geometric considerations. If the centre trajectory is hyperbolic, then the limit in (133) is reached exponentially fast.

We illustrated and tested the theory with several examples. For the free particle, and more generally, for integrable systems, the dynamics can be computed quite explicitly. The Ehrenfest time is of order TE ∼ ℏ−1/2 and we see a qualitative transition in the nature of the state from a localized coherent state to an extended WKB state at this time scale. Similarly, for a parabolic barrier, the dynamics can be computed explicitly and the formalism developed in this work gives explicit formulae, which describe in a uniform way the transition from reflection to transmission of a wave packet when one varies the energy near the critical energy. We then carried out detailed numerical tests on the kicked harmonic oscillator for an initial state localized on a hyperbolic fixed point. These showed impressive agreement between the metaplectic extension of the WKB method and the exact quantum propagation. These tests also illustrated the simplicity of the method; the metaplectic correction is in fact a very simple operator, and if one has implemented the time-dependent WKB method, then adding the metaplectic correction is easy and at once allows one to propagate a much larger class of states.

There are many areas where the results from this paper should be of interest. In many physical systems, the Ehrenfest time for a realistic set of parameters is actually quite short, e.g., in quantum billiards coherent states spread out after a few bounces [22, 23], and the metaplectic extension of WKB should be able to describe this transition. The scattering of a wave packet of a barrier near the critical energy is another example; the Ehrenfest time is the time when the wave packet reaches the barrier, i.e. the time where the physically interesting processes start to occur. Metaplectically extended WKB allows us to describe this process explicitly and uniformly in ℏ and t. A large class of chemical reactions is described in the framework of transition state theory by the crossing of a barrier on a high-dimensional energy surface and modern experimental methods in atto and femto chemistry allow us to study the dynamics of such chemical reactions with impressive precision, [35, 36]. The metaplectic extension to WKB together with the normal form approach to transition state theory [37, 38] should allow us to give an efficient theoretical description of chemical processes on such time scales.

Appendix A.: The Wigner–Weyl correspondence

In this appendix, we collect some material on Weyl quantization, which we use in the main part. Much of it is a development of standard material and only the result on Egorov's theorem seems to be new. For the material we present, see [30, 31].

A.1. Weyl quantization

Let $\hat{W}(\xi , x):=\mathrm{e}^{\frac{\mathrm{i}}{\hbar } (\xi \hat{q}+x\hat{p})}$ be the Weyl operators that represent translations in phase space; then, we can define for a function A(p, q) on phase space an operator, its Weyl quantization, by

Equation (A.1)

where $\mathcal {F}A(\xi ,x)=\int\!\!\!\int A(p,q)\,\mathrm{e}^{-\frac{\mathrm{i}}{\hbar }[p\xi +qx]}\, \mathrm{d}p\,\mathrm{d}q$ denotes the Fourier transform of A. The function A(p, q) is called the Weyl symbol of $\hat{A}$, and the properties of $\hat{A}$ can often be determined from properties of A, e.g., if $A\in S^{\prime }(\mathbb{R}^n\times \mathbb{R}^n)$, then $\hat{A} :S(\mathbb{R}^n)\rightarrow S^{\prime }(\mathbb{R}^n)$. If $\hat{A}=|\psi \rangle \langle \psi |$ is the projection onto a state ψ, then A(ξ, x) is proportional to the Wigner function of the state ψ.

The product of operators can be expressed in terms of the symbols, and one has $\hat{A}\hat{B}=\widehat{A\sharp B}$, where

Equation (A.2)

Here, {A, B}(p, q) denote the Poisson bracket, and the arrows over the derivatives indicate if they act on the function on the left or on the right. Suitable conditions on the functions A and B under which this expansion holds can be found in [30].

A.2. Dynamics generated by first-order operators

Let K(t, x, ξ) = X(t, x) · ξ, where X(t, x) is a time-dependent vector field. We are interested in the time evolution T(t, s) generated by

Equation (A.3)

Let ϕ(t, s, x) be the family of maps generated by the vector field X(t, x):

Equation (A.4)

Then,

Equation (A.5)

A.3. Egorov's theorem

Egorov's theorem is one way to formulate the correspondence principle; it gives a general relation between classical and quantum dynamics. Let H(t) be a real-valued smooth phase space function, $\hat{H}(t)$ be its Weyl quantization, and Φt and U(t) be the classical and quantum time evolution generated by H(t) and $\hat{H}(t)$, respectively. Then, Egorov's theorem states that

Equation (A.6)

where At = A○Φt, and A and H have to satisfy some conditions on their smoothness and growth at infinity, see [28]. Since At is the classical time evolution of A, this means that quantum and classical evolution are close for small ℏ. The remainder term Ot(ℏ2) does depend on time, and the best general estimates are of the form Ot(ℏ2) ≪ ℏ2 eΓt for some constant Γ > 0, which depends on H. We want to discuss two cases in which one has better control over the remainder.

One approach to Egorov's theorem is based on writing Heisenberg's equation of motion for $\hat{A}(t)=U^*(t)\hat{A}U(t)$, i.e. $\mathrm{i}\hbar \partial _t\hat{A}(t)=[\hat{H},\hat{A}]$, in terms of the symbols using (A.2), which gives

Equation (A.7)

The leading order term is just the Liouville equation and gives AA○Φt, the main problem is then to control the higher order terms.

A.3.1. Quadratic Hamiltonians and metaplectic operators

If H(p, q) is a quadratic function of p and q, with possibly time-dependent coefficients, then the ℏ expansion in (A.7) terminates after the leading term, and hence, the evolution equation for A is just the classical Liouville equation. So in this case

Equation (A.8)

where At = A○Φt, and one says Egorov's theorem is exact. The corresponding classical maps Φt are linear; hence, the operators U(t) for all quadratic H form a quantization of the symplectic group, which turns out to be a double cover of the symplectic group called the metaplectic group. The operators U(t) are often referred to as metaplectic operators [3, 32, 33].

A.3.2. Conjugation by a flow

The second case we need is that U(t) = T(t, 0), i.e. H(t) is linear in p and given by H(t, p, q) = K(t, p, q) = X(t, q) · p. The classical map Φt1(p, q) generated by H is given by the solutions to

Equation (A.9)

with the initial conditions ξ(0) = p and x(0) = q. We can express Φt1 in terms of the map ϕ(t, 0, x) as

Equation (A.10)

where [ϕ'(0, t, x)]† is the transpose of the inverse of the matrix ϕ'(t, 0, x).

Let us consider first the case that A is linear in p, i.e. A = b(q) · p for some vector-valued function b(q); then, only the leading order term in (A.7) is non-zero, and hence

Equation (A.11)

without any remainder terms; hence, Egorov's theorem is exact again. Using this result, we can discuss the case we will need, namely the case that $\hat{A}$ is a second-order differential operator of the form

Equation (A.12)

so that $T^*\hat{A}T=T^*{\mathcal B}(q)\nabla T\cdot T^* {\mathcal B}(q)\nabla T,$ and if we denote the rows of ${\mathcal B}$ by bj(q), we can use the previous result (A.11) and (A.2) to obtain $T^*(t,0)\hat{A} T(t,0)=\hat{A}(t)$ with

Equation (A.13)

where we have used (A.2) . This can be further simplified, but we will restrict ourselves to the case ${\mathcal B}=I$; hence, bi = ei and then we find

Equation (A.14)

which is the symbol of T*(t, 0)ΔT(t, 0). The term of order ℏ2 contains second derivatives of the inverse of ϕ(t, 0; q) and so if ϕ(t, 0; q) is non-contracting, then these derivatives stay bounded.

Appendix B.: A lemma on symplectic maps

The results in this appendix are used at the end of section 5 to show that the classical map associated with the metaplectic correction is uniquely determined by the dynamics of the tangent space to the initial Lagrangian submanifold and the vertical subspace.

Lemma B.1. Let $L_1,L_2\subset \mathbb{R}^n\times \mathbb{R}^n$ be Lagrangian subspaces with L1L2 = {0} and let L be another Lagrangian subspace with LL1 = {0}. Then, there exists a unique linear symplectic map T with $T|_{L_1}=I$ and T(L2) = L.

To show this, we will use mainly the following two standard facts from linear symplectic geometry; see, e.g., [31].

  • (a)  
    By the Darboux theorem, there exist symplectic coordinates $(v,w)\in \mathbb{R}^n\times \mathbb{R}^n$ such that L1 = {w = 0} and L2 = {v = 0}.
  • (b)  
    If L3 is a Lagrangian subspace with L3L1 = {0}, then there exists a unique symmetric matrix A, such that L3 = {v = Aw}.

Let us now prove the lemma. The condition $T|_{L_1}=I$ implies that in the coordinates from (a) the matrix representing T is of the form $\mathcal {M}_T=\big(\begin{array}{@{}cc@{}}{\ssty I} & {\ssty A}\\ [-5pt] {\ssty 0} &{\ssty B}\end{array}\big)$, where A and B are n × n matrices. Now this matrix must also be symplectic, i.e. MtTΩMT = Ω with $\Omega =\big(\begin{array}{@{}cc@{}}{\ssty 0} & {\ssty -I}\\ [-5pt] {\ssty I} & {\ssty 0}\end{array}\big)$ and this gives B = I and A = At. Therefore,

Equation (B.1)

but from (b) the condition T(L2) = L then uniquely determines A. So $M_T=\big(\begin{array}{@{}cc@{}} {\ssty I} & {\ssty A}\\ [-4pt] {\ssty 0} &{\ssty I}\end{array}\big)$ is the unique shear relative to L1, which maps L2 to L.

Please wait… references are loading.
10.1088/1751-8113/45/21/215307