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

Numerical solution of the linear semiclassical Schrödinger equation on the real line

Arieh Iserles
Department of Applied Mathematics and Theoretical Physics
Centre for Mathematical Sciences
University of Cambridge
Wilberforce Rd, Cambridge CB4 1LE
United Kingdom
   Karolina Kropielnicka
Institute of Mathematics
Polish Academy of Sciences
Antoniego Abrahama 18, 81-825 Sopot
Poland
   Katharina Schratz
Laboratoire Jacques-Louis Lions
Sorbonne Université
4 place Jussieu, 75252 Paris
France
   Marcus Webb
Department of Mathematics
University of Manchester
Alan Turing Building
Manchester M13 9PL
United Kingdom
Abstract

The numerical solution of a linear Schrödinger equation in the semiclassical regime is very well understood in a torus 𝕋d\mbox{\Bbb T}^{d}. A raft of modern computational methods are precise and affordable, while conserving energy and resolving high oscillations very well. This, however, is far from the case with regard to its solution in d\mbox{\Bbb R}^{d}, a setting more suitable for many applications. In this paper we extend the theory of splitting methods to this end. The main idea is to derive the solution using a spectral method from a combination of solutions of the free Schrödinger equation and of linear scalar ordinary differential equations, in a symmetric Zassenhaus splitting method. This necessitates detailed analysis of certain orthonormal spectral bases on the real line and their evolution under the free Schrödinger operator.

1 Introduction

1.1 Why the real line?

This paper is concerned with the numerical solution of the linear Schrödinger equation in the semiclassical regime, describing the motion of an electron in a quantum system,

iεut=ε2Δu+V(𝒙)u,t0,𝒙d,{\mathrm{i}}\varepsilon\frac{\partial u}{\partial t}=-\varepsilon^{2}\Delta u+V(\mbox{\boldmath$x$\unboldmath})u,\qquad t\geq 0,\;\mbox{\boldmath$x$\unboldmath}\in\mbox{\Bbb R}^{d}, (1.1)

where the initial condition u(𝒙,0)=u0(𝒙)L2(d)u(\mbox{\boldmath$x$\unboldmath},0)=u_{0}(\mbox{\boldmath$x$\unboldmath})\in\mathrm{L}_{2}(\mbox{\Bbb R}^{d}) for all 𝒙d\mbox{\boldmath$x$\unboldmath}\in\mbox{\Bbb R}^{d} is given. The semiclassical parameter ε>0\varepsilon>0 is a small number which describes the square root of the ratio between the mass of an electron and the total mass of the system, and V:dV:\mbox{\Bbb R}^{d}\rightarrow\mbox{\Bbb R} is the interaction potential which is assumed to be smooth for the purposes of this paper. Since |u(𝒙,t)|2|u(\mbox{\boldmath$x$\unboldmath},t)|^{2} gives the probability density of the electron residing at 𝒙x at time tt, the system is required to satisfy,

d|u(𝒙,t)|2d𝒙1,\int_{\mbox{\sBbb R}^{d}}|u(\mbox{\boldmath$x$\unboldmath},t)|^{2}\,\mathrm{d}\mbox{\boldmath$x$\unboldmath}\equiv 1, (1.2)

and any physically relevant numerical solution must be consistent with this conservation law. To read more, (?) is an excellent, up to date review of both the equation (1.1) and its numerical solution.

Respecting the unitarity property (1.2) underlies the importance of geometric numerical integration methodologies in this context and has been central to modern treatment of the linear Schrödinger equation in the semiclassical, 0<ε10<\varepsilon\ll 1, and the atomistic, ε=1\varepsilon=1, regimes alike (?????). However, all these publications are focussed on a subtly different problem: instead of being defined in d\mbox{\Bbb R}^{d}, the equation (1.1) is set on a torus, typically 𝕋d\mbox{\Bbb T}^{d}, with periodic boundary conditions. This is of crucial importance to splitting techniques, a common denominator to all these methodologies, because the free Schrödinger equation

iut=εΔu,{\mathrm{i}}\frac{\partial u}{\partial t}=-\varepsilon\Delta u, (1.3)

given with periodic boundary conditions, can be approximated very rapidly, affordably and precisely by means of the Fast Fourier Transform (FFT).

Our contention is that the periodic setting imposes unwelcome limitations on the solution, which might lead to altogether false outcomes, and this becomes problematic once a solution over a long time interval is sought (e.g. in quantum control). The underlying reason is the tension arising from the nature of the differential equation and of the initial condition, both predicated by quantum-mechanical considerations. The differential equation itself is dispersive: different waves travel at different speeds, dependent on their wavelengths, which can span a very wide range, all the way from 𝒪(1){\cal O}\!\left(1\right) to 𝒪(ε1){\cal O}\!\left(\varepsilon^{-1}\right). The initial condition is typically a linear combination of highly localised (and rapidly oscillating) wave packets. Recall that |u(𝒙,t)|2|u(\mbox{\boldmath$x$\unboldmath},t)|^{2} represents the probability of a particle residing at 𝒙x in time tt: while it is a central tenet of quantum mechanics that a particle cannot be completely localised, typically |u(𝒙,t)|2|u(\mbox{\boldmath$x$\unboldmath},t)|^{2} is a linear combination of narrowly-concentrated Gaussian-like structures. These Gaussian-like structures travel at different speeds and, provided the equation is solved for sufficiently long time, some of them eventually reach the boundary. At this very moment periodicity becomes a foe because the wave packet reaches the boundary and ‘pops out’ at the other end — this is not physical!

An alternative to periodic boundary conditions is to impose zero Dirichlet or zero Neumann boundary conditions. However, the following argument shows that this approach is also problematic. Consider an initial condition u0H01(0,1)u_{0}\in\mathrm{H}^{1}_{0}(0,1) and potential VH1(0,1)V\in\mathrm{H}^{1}(0,1). Now consider the following two initial boundary value problems, the first of which has zero Dirichlet boundary conditions, the second of which has periodic boundary conditions:

iεvt\displaystyle{\mathrm{i}}\varepsilon\frac{\partial v}{\partial t} =\displaystyle= ε22vx2+V(x)v,x[0,1],\displaystyle-\varepsilon^{2}\frac{\partial^{2}v}{\partial x^{2}}+V(x)v,\qquad x\in[0,1], (1.4)
v(0,t)\displaystyle v(0,t) =\displaystyle= 0,v(1,t)=0,t>0,\displaystyle 0,\qquad v(1,t)=0,\qquad t>0,
v(x,0)\displaystyle v(x,0) =\displaystyle= u0(x),x[0,1],\displaystyle u_{0}(x),\qquad x\in[0,1],

and,

iεwt\displaystyle{\mathrm{i}}\varepsilon\frac{\partial w}{\partial t} =\displaystyle= ε22wx2+V(|x|)w,x[1,1],\displaystyle-\varepsilon^{2}\frac{\partial^{2}w}{\partial x^{2}}+V(|x|)w,\qquad x\in[-1,1], (1.5)
w(1,t)\displaystyle w(-1,t) =\displaystyle= w(1,t),xw(1,t)=xw(1,t),t>0,\displaystyle w(1,t),\qquad\partial_{x}w(-1,t)=\partial_{x}w(1,t),\qquad t>0,
w(x,0)\displaystyle w(x,0) =\displaystyle= sign(x)u0(|x|),x[1,1].\displaystyle\mathrm{sign}(x)u_{0}(|x|),\qquad x\in[-1,1].

The relationship between v(x,t)v(x,t) and w(x,t)w(x,t) for x[0,1]x\in[0,1] and t>0t>0 is rather simple. Clearly the oddness of w(x,0)w(x,0) is preserved since the second derivative and multiplication by V(|x|)V(|x|) preserve oddness. Combining oddness with periodicity implies that w(0,t)=0=w(1,t)w(0,t)=0=w(1,t) for all time. It therefore follows from uniqueness of solution to (1.4) that w(x,t)=v(x,t)w(x,t)=v(x,t) for x[0,1]x\in[0,1] and t>0t>0. So now let us return to the notion of a wave packet moving towards the boundary, but this time with zero Dirichlet boundary conditions imposed. The solution to the odd extension implies that this wave packet will be reflected back and its sign reversed — while this physically happens in the case of an infinite potential barrier, it is not the correct behaviour when posed in free space! A similar construction can be made for Neumann boundary conditions.

Refer to caption
Figure 1.1: Top: We plot the evolution of (1.4) with u0(x)u_{0}(x) an approximate wave packet (so that zero boundary conditions are satisfied) and V(x)=0V(x)=0. The wave packet moves rightwards towards the right boundary until time t=10t=10, after which it moves leftwards, returning unscathed by the encounter. Such “reflections” contradict the behaviour of a wave packet in free space. Bottom: We plot the evolution of the corresponding extension in (1.5). We see that the reflection behaviour for the Dirichlet initial boundary value problem can be explained by the periodic behaviour of this one.

We hope this has convinced the reader: no matter what we do, and no matter how rapidly and accurately we can solve Schrödinger’s equation posed on a bounded set, the result of truncating the domain from d\mbox{\Bbb R}^{d} to such a set destroys the physics of the problem over a large enough time interval. This is the raison d’être for this paper: solve (1.1) without compromising its setting in d\mbox{\Bbb R}^{d}. Throughout the remainder of the paper we assume that (1.1) is presented in a single space dimension, d=1d=1. A generalisation to a modest number of space dimensions can be accomplished with tensor products along the lines of (?), while generalisation to a large number of dimensions would require a raft of additional techniques and is beyond the scope of the current paper.

To achieve this aim, we will extend the framework of symmetric Zassenhaus splittings, which has been developed for (1.1) on the torus 𝕋 (??), to (1.1) posed on the whole real line. This is not a straightforward exercise, because we cannot use special properties of the Fourier basis. In Section 2 we derive these Zassenhaus splittings under more general assumptions, allowing for bases other than the Fourier basis to be used. In Section 3, we discuss the solution of the free Schrödinger equation (1.3), focusing on two bases which are orthonormal on the real line: Hermite functions and Malmquist–Takenaka functions. In Section 5 we demonstrate how these pieces can be put together to construct practical numerical solvers on the real line.

2 Splitting techniques

For the clarity of exposition we write x2\partial_{x}^{2} instead of Δ\Delta as in (1.1). The simplest splitting methodology is to separate the potential and kinetic parts in (1.1), iεx2uiε1V(x)u{\mathrm{i}}\varepsilon\partial_{x}^{2}u-{\mathrm{i}}\varepsilon^{-1}V(x)u, building upon the fact that separate solutions of

ut=iε1V(x)uandut=iεx2u\frac{\partial u}{\partial t}=-{\mathrm{i}}\varepsilon^{-1}V(x)u\qquad\mbox{and}\qquad\frac{\partial u}{\partial t}={\mathrm{i}}\varepsilon\partial_{x}^{2}u

are (at least in a torus or a parallelepiped) much less expensive to compute than those of the full problem. We abuse notation for the exponential and write

u(x,t)=eitε1V(x)u(x,0)andu(x,t)=eitεx2u(x,0)u(x,t)={\mathrm{e}}^{-{\mathrm{i}}t\varepsilon^{-1}V(x)}u(x,0)\qquad\mbox{and}\qquad u(x,t)={\mathrm{e}}^{{\mathrm{i}}t\varepsilon\partial_{x}^{2}}u(x,0)

for their respective solutions. Splitting methods produce a sequence of functions u0(x)u^{0}(x), u1(x)u^{1}(x), u2(x)u^{2}(x),\ldots, intended to satisfy uk(x)u(x,kh)u^{k}(x)\approx u(x,kh) where hh is the time-step parameter. These functions of xx can be discretised by any approach, for example by a spectral method.

The two simplest splitting methods are the Lie–Trotter formula

uk+1(x)=eiεhx2eihε1V(x)uk(x),u^{k+1}(x)={\mathrm{e}}^{{\mathrm{i}}\varepsilon h\partial_{x}^{2}}{\mathrm{e}}^{-{\mathrm{i}}h\varepsilon^{-1}V(x)}u^{k}(x), (2.6)

and

uk+1(x)=eihε1V(x)/2eiεhx2eihε1V(x)/2uk(x).u^{k+1}(x)={\mathrm{e}}^{-{\mathrm{i}}h\varepsilon^{-1}V(x)/2}{\mathrm{e}}^{{\mathrm{i}}\varepsilon h\partial_{x}^{2}}{\mathrm{e}}^{-{\mathrm{i}}h\varepsilon^{-1}V(x)/2}u^{k}(x). (2.7)

Of course, the role of ε1V(x)\varepsilon^{-1}V(x) and εx2\varepsilon\partial_{x}^{2} can be reversed. The latter approach, advocated in (?) in tandem with spectral methods, is the famous Strang splitting (known also as Strang–Marchuk splitting in Russian literature).

Formally, the Strang splitting is known to produce time-stepping methods bearing an error of 𝒪(h3){\cal O}\!\left(h^{3}\right). However, this is misleading because the error constant is of size 𝒪(ε1){\cal O}\!\left(\varepsilon^{-1}\right), as we show below using Theorem 3. A more effective measure of error should incorporate the small parameter ε\varepsilon, which may be even smaller in magnitude than the time-step hh. To calculate the effective error of the splitting (2.7), where the error constant does not depend on the small semiclassical parameter ε\varepsilon, let us have a closer look at symmetric Baker–Campbell–Hausdorff formula (?, Sec. III.4.2),

e12τAeτBe12τA=esBCH(τA,τB){\mathrm{e}}^{\frac{1}{2}\tau A}{\mathrm{e}}^{\tau B}{\mathrm{e}}^{\frac{1}{2}\tau A}={\mathrm{e}}^{\mathrm{sBCH}(\tau A,\tau B)} (2.8)

where A=ε1VA=-\varepsilon^{-1}V, B=εx2B=\varepsilon\partial_{x}^{2} and τ=ih\tau={\mathrm{i}}h with

sBCH(τA,τ\displaystyle\mathrm{sBCH}(\tau A,\tau B)=τA+τBτ3124[[B,A],A]τ3112[[B,A],B]\displaystyle B)=\tau A+\tau B-\tau^{3}\frac{1}{24}[[B,A],A]-\tau^{3}\frac{1}{12}[[B,A],B]
+τ575760[[[[B,A],A],A],A]+τ571440[[[[B,A],A],A],B]\displaystyle+\tau^{5}\frac{7}{5760}[[[[B,A],A],A],A]+\tau^{5}\frac{7}{1440}[[[[B,A],A],A],B]
+τ51180[[[[B,A],A],B],B]+τ51720[[[[B,A],B],B],B]\displaystyle+\tau^{5}\frac{1}{180}[[[[B,A],A],B],B]+\tau^{5}\frac{1}{720}[[[[B,A],B],B],B]
+τ51480[[[B,A],A],[B,A]]τ51360[[[B,A],B],[B,A]]+h.o.t.\displaystyle+\tau^{5}\frac{1}{480}[[[B,A],A],[B,A]]-\tau^{5}\frac{1}{360}[[[B,A],B],[B,A]]+{\rm h.o.t.}

Given that AA and BB are unbounded operators and also contain powers of ε\varepsilon, we now proceed to clarify the meaning of “h.o.t.” (higher order terms).

2.1 A new analysis of the sBCH formula for the semiclassical Schrödinger equation

As it was shown in (?) Schrödinger equations in semiclassical regime produce oscillations in space of frequency 𝒪(ε1){\cal O}\!\left(\varepsilon^{-1}\right), which places restrictions on the discretisation in space depending on which basis is used, because we must employ sufficiently fine discretisation to resolve these oscillations. If the spatial variable is discretised using the Fourier basis then this necessitates 𝒪(ε1)\mathcal{O}(\varepsilon^{-1}) basis elements, which in turn, leads to the conclusion that after discretisation, operators of type xn\partial_{x}^{n} have a spectral radius which scales like 𝒪(εn){\cal O}\!\left(\varepsilon^{-n}\right). As we discuss in Section 3, for other bases it is not necessarily the case that 𝒪(ε1)\mathcal{O}(\varepsilon^{-1}) basis elements can resolve spatial oscillations of frequency 𝒪(ε1){\cal O}\!\left(\varepsilon^{-1}\right) (indeed the Fourier basis is the optimal basis for resolving periodic oscillations). As such, we will not make assumptions about the number of basis elements, but rather, make assumptions directly on the spectral radius of the partial derivative operator (an assumption which holds in both of our examples discussed in Section 3)

Assumption 1

Throughout this paper we will assume that after spatial discretisation, the operator x\partial_{x} has spectral radius 𝒪(ε1)\mathcal{O}(\varepsilon^{-1}).

Since the potential V(x)V(x) can in principle be an unbounded function on the real line, we must be careful that our expansions be treated locally in xx.

Assumption 2

The potential V:V:\mbox{\Bbb R}\to\mbox{\Bbb R} is infinitely differentiable, which we write VCloc()V\in\mathrm{C}^{\infty}_{\mathrm{loc}}(\mbox{\Bbb R}). As a result, all derivatives are locally bounded in .

We can now make sense of “h.o.t.” in the sBCH formula by bounding the magnitude element of the Hall basis for the free Lie algebra generated by A=ε1VA=-\varepsilon^{-1}V and B=εx2B=\varepsilon\partial_{x}^{2} (i.e. AA, BB, [A,B][A,B], [[B,A],A][[B,A],A], [[B,A],B][[B,A],B] \ldots (??)).

Theorem 3

Let A=ε1VA=-\varepsilon^{-1}V and B=εx2B=\varepsilon\partial_{x}^{2} assume they have been discretised following Assumption 1, and stipulate Assumption 2. Then all terms CC of the Hall basis constructed of letters AA and BB either vanish (i.e. C0C\equiv 0) or are 𝒪(ε1)\mathcal{O}(\varepsilon^{-1}).

Before we proceed with the proof of the theorem, note that all elements of Hall basis (?) of commutators constructed of letters AA and BB live in the set

G={k=0Kyk(x)xk:K+,y0,,yKCloc()},\mbox{\gothic G}=\left\{\sum_{k=0}^{K}y_{k}(x)\partial_{x}^{k}\,:K\in\mbox{\Bbb Z}_{+},y_{0},\ldots,y_{K}\in\mathrm{C}^{\infty}_{\mathrm{loc}}(\mbox{\Bbb R})\right\},

by applying the product rule (for differentiation). For example,

[B,A]=[x2,V]=(V(2)+2V(1)x+Vx2Vx2)=V(2)2V(1)x\displaystyle[B,A]=-[\partial_{x}^{2},V]=-\left(V^{(2)}+2V^{(1)}\partial_{x}+V\partial_{x}^{2}-V\partial_{x}^{2}\right)=-V^{(2)}-2V^{(1)}\partial_{x}
[[B,A],A]=ε1[[x2,V],V]=ε12(V(1))2\displaystyle[[B,A],A]=\varepsilon^{-1}[[\partial_{x}^{2},V],V]=\varepsilon^{-1}2(V^{(1)})^{2}
[[B,A],B]=ε[[x2,V],x2]=ε(V(4)+4V(3)x+4V(2)x2),\displaystyle[[B,A],B]=-\varepsilon[[\partial_{x}^{2},V],\partial_{x}^{2}]=\varepsilon\left(V^{(4)}+4V^{(3)}\partial_{x}+4V^{(2)}\partial_{x}^{2}\right),
[[[B,A],A],A]=0,\displaystyle[[[B,A],A],A]=0,

where V(k)=xkVV^{(k)}=\partial_{x}^{k}V. We define the height of the commutator CC as the largest index of non-zero coefficient yK(x)y_{K}(x):

ht(C)=ht(k=0Kyk(x)xk)=K, where yK(x)0,{\rm ht}(C)={\rm ht}\!\left(\sum_{k=0}^{K}y_{k}(x)\partial_{x}^{k}\right)=K,\text{ where }y_{K}(x)\not\equiv 0,

One can observe, that ht(A)=0{\rm ht}(A)=0, ht(B)=2{\rm ht}(B)=2, ht([B,A])=1{\rm ht}([B,A])=1, ht([[B,A],A])=0{\rm ht}([[B,A],A])=0 and ht([[B,A],B])=2{\rm ht}([[B,A],B])=2.

In the proof we will also refer to the formula elaborated in (?)

[i=0nfi(x)xi,j=0mgj(x)xj]=\displaystyle\left[\sum_{i=0}^{n}f_{i}(x)\partial^{i}_{x},\sum_{j=0}^{m}g_{j}(x)\partial^{j}_{x}\right]= i=0nj=0m=0i(i)fi(x)(xigj(x))x+j\displaystyle\sum_{i=0}^{n}\sum_{j=0}^{m}\sum_{\ell=0}^{i}\binom{i}{\ell}f_{i}(x)\left(\partial^{i-\ell}_{x}g_{j}(x)\right)\partial^{\ell+j}_{x}
j=0mi=0n=0j(j)gj(x)(xjfi(x))x+i.\displaystyle\mbox{}-\sum_{j=0}^{m}\sum_{i=0}^{n}\sum_{\ell=0}^{j}\binom{j}{\ell}g_{j}(x)\left(\partial^{j-\ell}_{x}f_{i}(x)\right)\partial^{\ell+i}_{x}. (2.9)

Proof (of Theorem 3) Let us assume, that a certain non-zero commutator CC in Hall basis is built of NAN_{A} letters AA and NBN_{B} letters BB. We show by induction on NA+NBN_{A}+N_{B}, that

ht(C)NBNA+1.{\rm ht}(C)\leq N_{B}-N_{A}+1. (2.10)

The cases in which NA+NB=1N_{A}+N_{B}=1 are obtained explicitly as ht(A)=0{\rm ht}(A)=0 and ht(B)=2{\rm ht}(B)=2, thus (2.10) is satisfied for the generators of the free Lie algebra. Now let us assume that a given non-zero commutator CC satisfies (2.10), so can be written as

C=k=0Kyk(x)xk,C=\sum_{k=0}^{K}y_{k}(x)\partial_{x}^{k},

where 0KNBNA+10\leq K\leq N_{B}-N_{A}+1 and yK0y_{K}\not\equiv 0. Then by (2.9),

[A,C]\displaystyle[A,C] =ε1k=0K1(j=kK(jk)yj(x)V(jk)(x))xk,\displaystyle=\varepsilon^{-1}\sum_{k=0}^{K-1}\left(\sum_{j=k}^{K}\binom{j}{k}y_{j}(x)V^{(j-k)}(x)\right)\partial_{x}^{k},
[B,C]\displaystyle[B,C] =εk=0Kyk′′(x)xk+2yk(x)xk+1,\displaystyle=\varepsilon\sum_{k=0}^{K}y_{k}^{\prime\prime}(x)\partial_{x}^{k}+2y_{k}^{\prime}(x)\partial_{x}^{k+1},

Therefore, ignoring the cases where these commutators vanish identically, we see that (2.10) is satisfied for [A,C][A,C] and [B,C][B,C] by the inductive hypothesis. This, in fact, completes the induction step for the entire Hall basis, because any commutator in the Hall basis can be written as a linear combination of words of the form

[a1,[a2,[,[an1,an]]]],[a_{1},[a_{2},[\dots,[a_{n-1},a_{n}]\dots]]],

where ak{A,B}a_{k}\in\{A,B\} for all kk, by the Jacobi identity (this is known as the Dynkin basis).

Next we show that every non-zero commutator CC in the Hall basis scales like 𝒪(ε1){\cal O}\!\left(\varepsilon^{-1}\right). Indeed, when CC is made up of NAN_{A} letters A=ε1VA=-\varepsilon^{-1}V and NBN_{B} letters B=εx2B=\varepsilon\partial_{x}^{2}, the linearity of commutators implies the equality of commutators CC and εNBNAC¯\varepsilon^{N_{B}-N_{A}}\bar{C}, where C¯\bar{C} has the same structure as CC, but with A¯=V\bar{A}=-V and B¯=x2\bar{B}=\partial_{x}^{2} instead of AA and BB. Obviously ht(C)=ht(C¯){\rm ht}(C)={\rm ht}(\bar{C}). Now by Assumption 1, x\partial_{x} scales like ε1\varepsilon^{-1} after discretisation and by Assumption 2 we have that all variable coefficients yky_{k} lie in Cloc()\mathrm{C}^{\infty}_{\mathrm{loc}}(\mbox{\Bbb R}) (so all derivatives are locally bounded). Therefore C¯=𝒪(εht(C))\bar{C}=\mathcal{O}(\varepsilon^{-{\rm ht}(C)}). Since for non-zero CC, we have ht(C)NBNA+1{\rm ht}(C)\leq N_{B}-N_{A}+1, we conclude that,

C=εNBNAC¯=𝒪(εNBANht(C¯))=𝒪(εNBNA(NBNA+1))=𝒪(ε1),C=\varepsilon^{N_{B}-N_{A}}\bar{C}=\mathcal{O}(\varepsilon^{N_{B}-A_{N}-{\rm ht}(\bar{C})})=\mathcal{O}(\varepsilon^{N_{B}-N_{A}-(N_{B}-N_{A}+1)})=\mathcal{O}(\varepsilon^{-1}),

which concludes the proof of the theorem.        \Box

An immediate consequence of Theorem 3 and (2.8) is that

e12τAeτBe12τA=eτ(A+B)+𝒪(h3ε1)=eτ(A+B)+𝒪(h3ε1).{\mathrm{e}}^{\frac{1}{2}\tau A}{\mathrm{e}}^{\tau B}{\mathrm{e}}^{\frac{1}{2}\tau A}={\mathrm{e}}^{\tau(A+B)+{\cal O}\!\left(h^{3}\varepsilon^{-1}\right)}={\mathrm{e}}^{\tau(A+B)}+{\cal O}\!\left(h^{3}\varepsilon^{-1}\right).

This means that taking the time step size h=𝒪(ε)h={\cal O}\!\left(\varepsilon\right) in the Strang splitting (2.7) yields a local truncation error of 𝒪(h2)\mathcal{O}(h^{2}) or equivalently, 𝒪(ε2)\mathcal{O}(\varepsilon^{2}). However, a time step h=𝒪(ε)h={\cal O}\!\left(\varepsilon\right) is overly expensive. If instead, one took a more reasonable h=𝒪(ε1/2)h={\cal O}\!\left(\varepsilon^{1/2}\right), then the local truncation error is effectively 𝒪(h)\mathcal{O}(h) or equivalently, 𝒪(ε1/2)\mathcal{O}(\varepsilon^{1/2}). In summary, unless the time step is unacceptably reduced, the effective error of the Strang splitting is larger than that suggested by an analysis which ignores the smallness of ε\varepsilon.

2.2 Symmetric Zassenhaus splittings

This order reduction for the Strang splitting in the case of Hamiltonians in a semi-classical setting motivates the quest for higher order splittings. A systematic approach is to calculate higher order symmetric Zassenhaus splittings, first proposed in (?). Using this methodology we will derive two splittings for the solution operator exp(τ(A+B))\exp(\tau(A+B)) where A=ε1VA=-\varepsilon^{-1}V, B=εx2B=\varepsilon\partial_{x}^{2} and τ=ih\tau={\mathrm{i}}h, of order 𝒪(h3ε1){\cal O}\!\left(h^{3}\varepsilon^{-1}\right) and 𝒪(h5ε1){\cal O}\!\left(h^{5}\varepsilon^{-1}\right) respectively, in the family of symmetric Zassenhaus splittings.

  1. 1.

    To derive the first symmetric Zassenhaus splitting, we apply the sBCH formula in the following way.

    e12τAeτA+τBe12τA=esBCH(τA,τA+τB){\mathrm{e}}^{-\frac{1}{2}\tau A}{\mathrm{e}}^{\tau A+\tau B}{\mathrm{e}}^{-\frac{1}{2}\tau A}={\mathrm{e}}^{\mathrm{sBCH}(-\tau A,\tau A+\tau B)} (2.11)

    where

    sBCH(τA,τA+τB)=\displaystyle\mathrm{sBCH}(-\tau A,\tau A+\tau B)= (2.12)
    =τB+τ3124[[B,A]A]+τ3112[[B,A],B]\displaystyle=\tau B+\tau^{3}\frac{1}{24}[[B,A]A]+\tau^{3}\frac{1}{12}[[B,A],B]
    τ51720[[[[B,A],A],B],B]τ51720[[[[B,A],B],B],B]\displaystyle-\tau^{5}\frac{1}{720}[[[[B,A],A],B],B]-\tau^{5}\frac{1}{720}[[[[B,A],B],B],B]
    τ51480[[[B,A],A],[B,A]]τ51240[[[B,A],B],[B,A]]+𝒪(h7ε1).\displaystyle-\tau^{5}\frac{1}{480}[[[B,A],A],[B,A]]-\tau^{5}\frac{1}{240}[[[B,A],B],[B,A]]+{\cal O}\!\left(h^{7}\varepsilon^{-1}\right).

    Substituting (2.12) into (2.11) results in the first symmetric Zassenhauss splitting, which coincides with Strang splitting,

    eτ(A+B)\displaystyle{\mathrm{e}}^{\tau(A+B)} =e12τAesBCH(τA,τA+τB)e12τA\displaystyle={\mathrm{e}}^{\frac{1}{2}\tau A}{\mathrm{e}}^{\mathrm{sBCH}(-\tau A,\tau A+\tau B)}{\mathrm{e}}^{\frac{1}{2}\tau A} (2.13)
    =e12τAeτBe12τA+𝒪(h3ε1).\displaystyle={\mathrm{e}}^{\frac{1}{2}\tau A}{\mathrm{e}}^{\tau B}{\mathrm{e}}^{\frac{1}{2}\tau A}+{\cal O}\!\left(h^{3}\varepsilon^{-1}\right).
  2. 2.

    To derive the second symmetric Zassenhaus splitting, we split the inner term of (2.13) by the same approach as above, that is

    e12τBesBCH(τA,τA+τB)e12τB\displaystyle{\mathrm{e}}^{-\frac{1}{2}\tau B}{\mathrm{e}}^{\mathrm{sBCH}(-\tau A,\tau A+\tau B)}{\mathrm{e}}^{-\frac{1}{2}\tau B} =esBCH(τB,sBCH(τA,τA+τB))\displaystyle={\mathrm{e}}^{\mathrm{sBCH}(-\tau B,\mathrm{sBCH}(-\tau A,\tau A+\tau B))}

    which leads to,

    eτA+τB=e12τAe12τBesBCH(τB,sBCH(τA,τA+τB))e12τBe12τA,{\mathrm{e}}^{\tau A+\tau B}={\mathrm{e}}^{\frac{1}{2}\tau A}e^{\frac{1}{2}\tau B}{\mathrm{e}}^{\mathrm{sBCH}(-\tau B,\mathrm{sBCH}(-\tau A,\tau A+\tau B))}{\mathrm{e}}^{\frac{1}{2}\tau B}e^{\frac{1}{2}\tau A}, (2.14)

    where

    sBCH(τB,sBCH(τA,τB+τA))\displaystyle\mathrm{sBCH}(-\tau B,\mathrm{sBCH}(-\tau A,\tau B+\tau A))
    =124τ3[[B,A],A]+112τ3[[B,A],B]\displaystyle=\frac{1}{24}\tau^{3}[[B,A],A]+\frac{1}{12}\tau^{3}[[B,A],B]
    192880τ5[[[[B,A],A],B],B]171440τ5[[[[B,A],B],B],B]\displaystyle-\frac{19}{2880}\tau^{5}[[[[B,A],A],B],B]-\frac{17}{1440}\tau^{5}[[[[B,A],B],B],B]
    τ51480[[[B,A],A],[B,A]]τ51240[[[B,A],B],[B,A]]+𝒪(h7ε1).\displaystyle-\tau^{5}\frac{1}{480}[[[B,A],A],[B,A]]-\tau^{5}\frac{1}{240}[[[B,A],B],[B,A]]+{\cal O}\!\left(h^{7}\varepsilon^{-1}\right).

    Observe that by Theorem 3, the first two commutators (which involve three letters) scale like 𝒪(h3ε1){\cal O}\!\left(h^{3}\varepsilon^{-1}\right) and the remainder scales like 𝒪(h5ε1){\cal O}\!\left(h^{5}\varepsilon^{-1}\right). Therefore, these first two terms are what will appear in this Zassenhaus splitting. However, the commutator,

    [[B,A],B]=[[εx2,ε1V],εx2]\displaystyle[[B,A],B]=[[\varepsilon\partial_{x}^{2},-\varepsilon^{-1}V],\varepsilon\partial_{x}^{2}] =ε(V(4)+4V(3)x+4V(2)x2),\displaystyle=\varepsilon\left(V^{(4)}+4V^{(3)}\partial_{x}+4V^{(2)}\partial_{x}^{2}\right),

    will not be skew-Hermitian after discretisation (which would result in loss of unitarity of the method), and therefore cannot be substituted into (2.14). For this reason, as proposed in (?), we use a substitution rule of the following kind:

    y(x)x=12[x0xy(s)ds]x212xy(x)+12x2[x0xy(s)ds],y(x)\partial_{x}=-\frac{1}{2}\left[\int_{x_{0}}^{x}y(s){\rm d}s\right]\partial_{x}^{2}-\frac{1}{2}\partial_{x}y(x)+\frac{1}{2}\partial_{x}^{2}\left[\int_{x_{0}}^{x}y(s){\rm d}s\,\cdot\right],

    and obtain terms that remain skew-Hermitian after discretisation:

    sBCH(τB,sBCH(τA,τB+τA))\displaystyle\mathrm{sBCH}(-\tau B,\mathrm{sBCH}(-\tau A,\tau B+\tau A))
    =τ3ε1112(V(1))2+τ3ε112V(4)+τ3ε13V(3)x+τ3ε13V(2)x2+𝒪(h5ε1)\displaystyle=\tau^{3}\varepsilon^{-1}\frac{1}{12}(V^{(1)})^{2}+\tau^{3}\varepsilon\frac{1}{12}V^{(4)}+\tau^{3}\varepsilon\frac{1}{3}V^{(3)}\partial_{x}+\tau^{3}\varepsilon\frac{1}{3}V^{(2)}\partial_{x}^{2}+{\cal O}\!\left(h^{5}\varepsilon^{-1}\right)
    =\displaystyle= τ3ε1112(V(1))2+16τ3ε{V(2)x2+x2[V(2)]}𝒪(ε2)112τ3εV(4)+𝒪(h5ε1).\displaystyle\tau^{3}\varepsilon^{-1}\frac{1}{12}(V^{(1)})^{2}+\frac{1}{6}\tau^{3}\varepsilon\underbrace{\left\{V^{(2)}\partial_{x}^{2}+\partial_{x}^{2}\left[V^{(2)}\cdot\right]\right\}}_{{\cal O}\!\left(\varepsilon^{-2}\right)}-\frac{1}{12}\tau^{3}\varepsilon V^{(4)}+{\cal O}\!\left(h^{5}\varepsilon^{-1}\right).

    In the final form of the splitting (2.14) the small 𝒪(h3ε){\cal O}\!\left(h^{3}\varepsilon\right) term involving V(4)V^{(4)} can be discarded.

Summing up these two derivations, we have the splittings,

uk+1(x)=e0e21e0uk(x)+𝒪(h3ε1)u^{k+1}(x)={\mathrm{e}}^{\mathcal{R}_{0}}{\mathrm{e}}^{2\mathcal{R}_{1}}{\mathrm{e}}^{\mathcal{R}_{0}}u^{k}(x)+{\cal O}\!\left(h^{3}\varepsilon^{-1}\right) (2.15)

and

uk+1(x)=e0e1e22e1e0uk(x)+𝒪(h5ε1),u^{k+1}(x)={\mathrm{e}}^{\mathcal{R}_{0}}{\mathrm{e}}^{\mathcal{R}_{1}}{\mathrm{e}}^{2\mathcal{R}_{2}}{\mathrm{e}}^{\mathcal{R}_{1}}{\mathrm{e}}^{\mathcal{R}_{0}}u^{k}(x)+{\cal O}\!\left(h^{5}\varepsilon^{-1}\right), (2.16)

where, letting τ=ih\tau={\mathrm{i}}h,

0\displaystyle\mathcal{R}_{0} =\displaystyle= 12τε1V,\displaystyle-\frac{1}{2}\tau\varepsilon^{-1}V,
1\displaystyle\mathcal{R}_{1} =\displaystyle= 12τεx2,\displaystyle\frac{1}{2}\tau\varepsilon\partial_{x}^{2},
2\displaystyle\mathcal{R}_{2} =\displaystyle= 112τ3ε{x2[V(2)]+V(2)x2}+124τ3ε1(V(1))2.\displaystyle\color[rgb]{0,0,0}\frac{1}{12}\tau^{3}\varepsilon\left\{\partial_{x}^{2}[V^{(2)}\,\cdot\,]+V^{(2)}\partial_{x}^{2}\right\}+\frac{1}{24}\tau^{3}\varepsilon^{-1}(V^{(1)})^{2}.

Note that 0=𝒪(hε1)\mathcal{R}_{0}={\cal O}\!\left(h\varepsilon^{-1}\right), 1=𝒪(hε1)\mathcal{R}_{1}={\cal O}\!\left(h\varepsilon^{-1}\right), 2=𝒪(h3ε1)\mathcal{R}_{2}={\cal O}\!\left(h^{3}\varepsilon^{-1}\right).

It is also possible to derive even higher order methods, such as

un+1(x)=e0e1e2e23e2e1e0un(x)+𝒪(h7ε1),u^{n+1}(x)={\mathrm{e}}^{\mathcal{R}_{0}}{\mathrm{e}}^{\mathcal{R}_{1}}{\mathrm{e}}^{\mathcal{R}_{2}}{\mathrm{e}}^{2\mathcal{R}_{3}}{\mathrm{e}}^{\mathcal{R}_{2}}{\mathrm{e}}^{\mathcal{R}_{1}}{\mathrm{e}}^{\mathcal{R}_{0}}u^{n}(x)+{\cal O}\!\left(h^{7}\varepsilon^{-1}\right), (2.17)

where

3\displaystyle\mathcal{R}_{3} =\displaystyle= 1120τ5ε1V(2)(V(1))2+124τ3εV(4)\displaystyle-\frac{1}{120}\tau^{5}\varepsilon^{-1}V^{(2)}(V^{(1)})^{2}+\frac{1}{24}\tau^{3}\varepsilon V^{(4)}
+1120τ5ε{x2[(7(V(2))2+V(3)V(1))]+(7(V(2))2+V(3)V(1))x2}\displaystyle+\frac{1}{120}\tau^{5}\varepsilon\left\{\partial_{x}^{2}\left[\left(7(V^{(2)})^{2}+V^{(3)}V^{(1)}\right)\,\cdot\right]+\left(7(V^{(2)})^{2}+V^{(3)}V^{(1)}\right)\partial_{x}^{2}\right\}
+160τ5ε3{x4[V(4)]+V(4)x4}.\displaystyle+\frac{1}{60}\tau^{5}\varepsilon^{-3}\left\{\partial_{x}^{4}\left[V^{(4)}\,\cdot\,\right]+V^{(4)}\partial_{x}^{4}\right\}.

Note that 3=𝒪(h5ε1)\mathcal{R}_{3}={\cal O}\!\left(h^{5}\varepsilon^{-1}\right). We refer the reader to (?) for discussion of deriving such higher order methods via a sequence of skew-Hermitian operators 0,1,\mathcal{R}_{0},\mathcal{R}_{1},\ldots. Our new analysis encapsulated in Theorem 3 shows that each term \mathcal{R}_{\ell} is actually of size 𝒪(h21ε1){\cal O}\!\left(h^{2\ell-1}\varepsilon^{-1}\right) for =1,2,\ell=1,2,\ldots. In Section 5, we will discuss how to go about computing e{\mathrm{e}}^{\mathcal{R}_{\ell}} for each \ell.

3 Orthonormal systems and free Schrödinger evolutions

3.1 Orthogonal systems with tridiagonal differentiation matrices

Solving (1.1) by spectral methods based upon symmetric Zassenhaus splittings (2.15) or (2.16) involves three ingredients: the splitting itself into 0,1,2,\mathcal{R}_{0},\mathcal{R}_{1},\mathcal{R}_{2},\ldots, the choice of spectral basis, and the means to compute the exponentials e{\mathrm{e}}^{\mathcal{R}_{\ell}}. The generalisation of each to the new setting requires new ideas and substantial effort. In this subsection we are concerned with the choice of the spectral basis.

We seek a set Φ={φn}n=0\Phi=\{\varphi_{n}\}_{n=0}^{\infty} which forms an orthonormal basis of L2()\mathrm{L}_{2}(\mbox{\Bbb R}) – this means that any fL2()f\in\mathrm{L}_{2}(\mbox{\Bbb R}) can be expanded in the form

f(x)=n=0f^nφn(x),wheref^n=f(x)φn(x)¯dx,n+.f(x)=\sum_{n=0}^{\infty}\hat{f}_{n}\varphi_{n}(x),\qquad\mbox{where}\qquad\hat{f}_{n}=\int_{-\infty}^{\infty}f(x)\overline{\varphi_{n}(x)}\,\mathrm{d}x,\quad n\in\mbox{\Bbb Z}_{+}.

For the time being we require the φn\varphi_{n}s to be real, although this will be lifted as necessary (with suitable changes). In addition we require that Φ\Phi has a tridiagonal differentiation matrix (which, it is easy to prove, must be skew-symmetric),

φn=bn1φn1+bnφn+1,n+,\varphi_{n}^{\prime}=-b_{n-1}\varphi_{n-1}+b_{n}\varphi_{n+1},\qquad n\in\mbox{\Bbb Z}_{+}, (3.1)

where b1=0b_{-1}=0 and bn>0b_{n}>0, n+n\in\mbox{\Bbb Z}_{+}. This makes both computation and analysis considerably easier.

A comprehensive theory of such orthogonal systems has been developed in (??). The main issue, making (3.1) compatible with orthonormality, can be explicated by considering Fourier transforms of the φn\varphi_{n}s. Specifically, let w(ξ)dξw(\xi)\,\mathrm{d}\xi be a Borel measure over and its Radon–Nikodym derivative w0w\geq 0 be absolutely continuous and even. Furthermore assume that all the moments of this measure are finite. Such measure generates a system of orthonormal polynomials {pn}n=0\{p_{n}\}_{n=0}^{\infty},

pn(ξ)pm(ξ)w(ξ)dξ=0,mn,pn2(ξ)w(ξ)dξ=1.\int_{-\infty}^{\infty}p_{n}(\xi)p_{m}(\xi)w(\xi)\,\mathrm{d}\xi=0,\quad m\neq n,\qquad\int_{-\infty}^{\infty}p_{n}^{2}(\xi)w(\xi)\,\mathrm{d}\xi=1.

Then the scaled inverse Fourier transform,

φn(x)=(i)n2πpn(ξ)g(ξ)eixξdξ,n+,\varphi_{n}(x)=\frac{(-{\mathrm{i}})^{n}}{\sqrt{2\pi}}\int_{-\infty}^{\infty}p_{n}(\xi)g(\xi){\mathrm{e}}^{{\mathrm{i}}x\xi}\,\mathrm{d}\xi,\qquad n\in\mbox{\Bbb Z}_{+}, (3.2)

where gg is any function satisfying |g(ξ)|2=w(ξ)|g(\xi)|^{2}=w(\xi), forms an orthonormal system on the real line which satisfies (3.1). Note that this system is real-valued if and only if gg has even real part and odd imaginary part, for example g(ξ)=w(ξ)g(\xi)=\sqrt{w(\xi)}. The constants bnb_{n} in (3.1) are inherited from the recurrence relation for orthonormal polynomials,

bnpn+1(ξ)=ξpn(ξ)bn1pn1(ξ),n+.b_{n}p_{n+1}(\xi)=\xi p_{n}(\xi)-b_{n-1}p_{n-1}(\xi),\qquad n\in\mbox{\Bbb Z}_{+}.

The orthonormal system given by (3.2) need not be dense in – as a matter of fact, it is dense in the Paley–Wiener space 𝒫𝒲supp(w)()L2()\mathcal{PW}_{\mbox{supp}(w)}(\mbox{\Bbb R})\subseteq\mathrm{L}_{2}(\mbox{\Bbb R}) which is the space of L2()L_{2}(\mbox{\Bbb R}) functions whose Fourier transforms vanish outside of the support of ww. Therefore, the system is a basis of L2()\mathrm{L}_{2}(\mbox{\Bbb R}) if and only if the weight function ww is positive on the whole real line.

Complete orthonormal bases can be formed also from polynomials P={pn}n=0P=\{p_{n}\}_{n=0}^{\infty} orthogonal on the half-line [0,)[0,\infty) (?), e.g. the Laguerre polynomials whose orthogonality measure is eξdξ{\mathrm{e}}^{-\xi}\,\mathrm{d}\xi, ξ0\xi\geq 0: The representation (3.2) survives intact but, to render the system dense in L2()\mathrm{L}_{2}(\mbox{\Bbb R}), we need to complement PP with orthogonal polynomials with respect to the mirror image of ww in the left half-line, w(ξ)dξw(-\xi)\,\mathrm{d}\xi for ξ0\xi\leq 0. The new system Φ\Phi is enumerated by nn\in\mbox{\Bbb Z} and in place of (3.1) we have

φn=bn1φn1+icnφn+bnφn+1,n,\varphi_{n}^{\prime}=-b_{n-1}\varphi_{n-1}+{\mathrm{i}}c_{n}\varphi_{n}+b_{n}\varphi_{n+1},\qquad n\in\mbox{\Bbb Z},

with bn>0b_{n}>0, n0n\neq 0, b0=0b_{0}=0 and real cnc_{n} – note that the new differentiation matrix is skew-Hermitian.

3.2 Free Schrödinger evolutions

Given an orthonormal system Φ\Phi on the real line, we denote by ψn\psi_{n}, n+n\in\mbox{\Bbb Z}_{+}, the solution of the free Schrödinger equation (1.3) with the initial condition φn\varphi_{n} – in other words,

ψnt=iε2ψnx2,ψn(x,0)=φn(x),x.\frac{\partial\psi_{n}}{\partial t}=-{\mathrm{i}}\varepsilon\frac{\partial^{2}\psi_{n}}{\partial x^{2}},\qquad\psi_{n}(x,0)=\varphi_{n}(x),\;x\in\mbox{\Bbb R}. (3.3)

We call Ψ(t)={ψn(,t)}n=0\Psi(t)=\{\psi_{n}(\cdot,t)\}_{n=0}^{\infty} the free Schrödinger evolution (FSE) of Φ\Phi.

The exact solution of (3.3) via the Fourier transform is well known and can be easily verified by direct differentiation:

ψn(x,t)=12πφ^n(η)eiη2εt+iηxdη,\psi_{n}(x,t)=\frac{1}{\sqrt{2\pi}}\int_{-\infty}^{\infty}\hat{\varphi}_{n}(\eta){\mathrm{e}}^{{\mathrm{i}}\eta^{2}\varepsilon t+{\mathrm{i}}\eta x}\,\mathrm{d}\eta, (3.4)

where

φ^n(η)=12πφn(ξ)eiηξdξ\hat{\varphi}_{n}(\eta)=\frac{1}{\sqrt{2\pi}}\int_{-\infty}^{\infty}\varphi_{n}(\xi){\mathrm{e}}^{-{\mathrm{i}}\eta\xi}\,\mathrm{d}\xi

is the familiar Fourier transform of φn\varphi_{n}.

On the face of it, our job is done: any mention of the phrase “Fourier transform” elicits from a numerical analyst the instinctive response “Fast Fourier Transform!”. This, however, is somewhat rash. An FFT computes rapidly the discrete Fourier transform which, in turn, is a very precise (at any rate, for very smooth functions) approximation of the Fourier transform of a periodic function in a compact interval, while our setting is the entire real line. One possibility is to clip the real line, approximating it by a sufficiently large interval and disregarding the Gibbs effect at the endpoints. This immediately begs the question “how large” which, while not beyond the ken of numerical reasoning, presents its own challenges. In this paper we adopt an alternative – and arguably more effective – point of view, seeking the exact solution of (3.4) for specific orthonormal systems Φ\Phi. While this approach cannot be expected to apply to each and every Φ\Phi consistent with the setting of Subsection 2.1, it does so with the two most interesting orthonormal systems: Hermite functions and Malmquist–Takenaka functions.

Once FSEs Ψ(t)\Psi(t) are known, the solution of the free Schrödinger equation (1.3) with the initial condition u(x,kh)u(x,kh) proceeds as follows: The function u(x,kh)u(x,kh) is expanded in the orthonormal basis Φ\Phi,

u(x,kh)n=0Nu^nφn(x)u(x,kh)\approx\sum_{n=0}^{N}\hat{u}_{n}\varphi_{n}(x) (3.5)

for a sufficiently large truncation parameter NN. Having done so, linearity of (1.3) implies that

u(x,(k+1)h)n=0Nu^nψn(x,h).u(x,(k+1)h)\approx\sum_{n=0}^{N}\hat{u}_{n}\psi_{n}(x,h). (3.6)

We get the coefficients for free because they do not change — it is the basis which changes. The choice of NN is governed by approximation properties of the spectral basis, and its ability to approximate spatial oscillations of frequency 𝒪(ε1){\cal O}\!\left(\varepsilon^{-1}\right) as discussed in the introduction.

Indeed, orthonormal systems are not all of equal value: more specifically, they can approximate functions at different speeds. While standard spectral methods on a torus are known to converge (for analytic functions) at an exponential speed, equivalent theory does not exist yet on the real line. Recalling from Section 1 that solutions of (1.1) are typically composed of wave packets, it is instructive to enquire how well different orthonormal systems approximate wave packets. This is investigated in (?) for the two families Φ\Phi described in the sequel: in both cases we can prove exponential convergence to any set error tolerance.

We note for further reference that the computation of (3.6) (once NN and hh have been appropriately chosen) requires both the knowledge of Ψ(h)\Psi(h) and the means to evaluate an expansion as in (3.5).

Theorem 4

Let Φ\Phi be as in (3.2). Then the functions,

ψn(x,t)=(i)n2πpn(ξ)g(ξ)eixξ+iεtξ2dξ,n+,\psi_{n}(x,t)=\frac{(-{\mathrm{i}})^{n}}{\sqrt{2\pi}}\int_{-\infty}^{\infty}p_{n}(\xi)g(\xi){\mathrm{e}}^{{\mathrm{i}}x\xi+{\mathrm{i}}\varepsilon t\xi^{2}}\,\mathrm{d}\xi,\qquad n\in\mbox{\Bbb Z}_{+}, (3.7)

where {pn}n=0\{p_{n}\}_{n=0}^{\infty} is the system of orthonormal polynomials with respect to the measure |g(ξ)|2dξ|g(\xi)|^{2}\,\mathrm{d}\xi, satisfies (3.3) (in particular ψn(x,0)=φn(x)\psi_{n}(x,0)=\varphi_{n}(x)) and for all tt is itself an orthonormal basis of L2()\mathrm{L}_{2}(\mbox{\Bbb R}) satisfying,

ψn(x,t)x=bn1ψn1(x,t)+bnψn+1(x,t),n+,\frac{\partial\psi_{n}(x,t)}{\partial x}=-b_{n-1}\psi_{n-1}(x,t)+b_{n}\psi_{n+1}(x,t),\qquad n\in\mbox{\Bbb Z}_{+}, (3.8)

where {bn}n+\{b_{n}\}_{n\in\mbox{\sBbb Z}_{+}} are the same constants as in (3.1).

Proof Differentiating under the integral sign with respect to xx twice and tt once demonstrates that ψn(x,t)\psi_{n}(x,t) satisfies the free Schrödinger equation (3.3), and it is clear that setting t=0t=0 in this formula yields φn(x)\varphi_{n}(x).

To show that ψn\psi_{n} is an orthonormal system satisfying (3.8), note that

|g(ξ)eiεtξ2|2=|g(ξ)|2=w(ξ),\left|g(\xi){\mathrm{e}}^{{\mathrm{i}}\varepsilon t\xi^{2}}\right|^{2}=|g(\xi)|^{2}=w(\xi), (3.9)

so these functions still come under the framework of (3.2), with exactly the same polynomials {pn}n+\{p_{n}\}_{n\in\mbox{\sBbb Z}_{+}}, but with the function g(ξ)eiεtξ2g(\xi){\mathrm{e}}^{{\mathrm{i}}\varepsilon t\xi^{2}} in place of g(ξ)g(\xi).

\Box

3.3 Re-expanding an FSE expansion in the original basis

Suppose that have an expansion the FSE basis Ψ(t)={ψn}n=0\Psi(t)=\{\psi_{n}\}_{n=0}^{\infty},

u(x,t)=n=0anψn(x,t),u(x,t)=\sum_{n=0}^{\infty}a_{n}\psi_{n}(x,t), (3.10)

and we wish to re-expand this basis in terms of the original basis Φ(=Ψ(0))\Phi(=\Psi(0)) for each tt. Let us consider time-dependent coefficients αn(t)\alpha_{n}(t) satisfying

u(x,t)=n=0αn(t)φn(x).u(x,t)=\sum_{n=0}^{\infty}\alpha_{n}(t)\varphi_{n}(x). (3.11)

The relationship between 𝜶(t)\boldsymbol{\alpha}(t) and 𝒂\boldsymbol{a} is simple when considered in terms of the polynomial basis PP. Indeed, the relationship is given by

n=0(i)nαn(t)pn(ξ)=n=0(i)nanpn(ξ)eitξ2,\sum_{n=0}^{\infty}(-{\mathrm{i}})^{n}\alpha_{n}(t)p_{n}(\xi)=\sum_{n=0}^{\infty}(-{\mathrm{i}})^{n}a_{n}p_{n}(\xi){\mathrm{e}}^{{\mathrm{i}}t\xi^{2}}, (3.12)

where the expansions are convergent in the space L2(,w(ξ)dξ)\mathrm{L}_{2}(\mbox{\Bbb R},w(\xi)\mathrm{d}\xi). Writing this in terms of operators acting on the vectors 𝒂,𝜶(t)2\boldsymbol{a},\boldsymbol{\alpha}(t)\in\ell_{2}, we have

TS𝜶(t)=M(t)TS𝒂,TS\boldsymbol{\alpha}(t)=M(t)TS\boldsymbol{a}, (3.13)

where S:22S:\ell_{2}\to\ell_{2} simply multiplies nnth component of a sequence by (i)n(-{\mathrm{i}})^{n}, T:2L2(,w(ξ)dξ)T:\ell_{2}\to\mathrm{L}_{2}(\mbox{\Bbb R},w(\xi)\mathrm{d}\xi) is the synthesis operator for the basis PP (AKA coefficients-to-values operator), and M(t):L2(,w(ξ)dξ)L2(,w(ξ)dξ)M(t):\mathrm{L}_{2}(\mbox{\Bbb R},w(\xi)\mathrm{d}\xi)\to\mathrm{L}_{2}(\mbox{\Bbb R},w(\xi)\mathrm{d}\xi) multiplies functions by exp(itξ2)\exp\left({\mathrm{i}}t\xi^{2}\right). Note that since PP is an orthonormal basis for L2(,w(ξ)dξ)\mathrm{L}_{2}(\mbox{\Bbb R},w(\xi)\mathrm{d}\xi), we have that TT is a unitary operator (the inverse, TT^{*} is usually called the analysis operator or values-to-coefficients operator). SS is also clearly unitary, so we can invert operators to find,

𝜶(t)=STM(t)TS𝒂.\boldsymbol{\alpha}(t)=S^{*}T^{*}M(t)TS\boldsymbol{a}. (3.14)

Since M(t)M(t) is unitary, we see that as expected, the operation sending 𝒂\boldsymbol{a} to 𝜶(t)\boldsymbol{\alpha}(t) is unitary overall.

Now, let us project these equations onto the first N+1N+1 terms of Φ\Phi. We obtain,

𝜶[N](t)=SNTNMN(t)TNSN𝒂[N],\boldsymbol{\alpha}^{[N]}(t)=S_{N}^{*}T_{N}^{*}M_{N}(t)T_{N}S_{N}\boldsymbol{a}^{[N]}, (3.15)

where 𝜶[N](t),𝒂[N]N+1\boldsymbol{\alpha}^{[N]}(t),\boldsymbol{a}^{[N]}\in\mbox{\Bbb C}^{N+1}, and SN,TN,MN(t):N+1N+1S_{N},T_{N},M_{N}(t):\mbox{\Bbb C}^{N+1}\to\mbox{\Bbb C}^{N+1}. These discretised operators are, SN=diag((i)n)n=0NS_{N}=\mathrm{diag}((-{\mathrm{i}})^{n})_{n=0}^{N}, MN(t)=diag(exp(itξk2))k=0NM_{N}(t)=\mathrm{diag}(\exp({\mathrm{i}}t\xi_{k}^{2}))_{k=0}^{N}, and

TN=(w0p0(ξ0)w0p1(ξ0)w0pN(ξ0)w1p0(ξ1)w1p1(ξ1)w1pN(ξ1)wNp0(ξN)wNp1(ξN)wNpN(ξN)),T_{N}=\begin{pmatrix}\sqrt{w_{0}}p_{0}(\xi_{0})&\sqrt{w_{0}}p_{1}(\xi_{0})&\cdots&\sqrt{w_{0}}p_{N}(\xi_{0})\\ \sqrt{w_{1}}p_{0}(\xi_{1})&\sqrt{w_{1}}p_{1}(\xi_{1})&\cdots&\sqrt{w_{1}}p_{N}(\xi_{1})\\ \vdots&\vdots&&\vdots\\ \sqrt{w_{N}}p_{0}(\xi_{N})&\sqrt{w_{N}}p_{1}(\xi_{N})&\cdots&\sqrt{w_{N}}p_{N}(\xi_{N})\\ \end{pmatrix}, (3.16)

where w0,,wN,ξ0,,ξNw_{0},\ldots,w_{N},\xi_{0},\ldots,\xi_{N} are Gauss quadrature weights and nodes (respectively) for the measure w(ξ)dξw(\xi)\mathrm{d}\xi. First, note that the unitarity of the operators has been preserved by this discretisation. Second, note that the unitary matrix TNT_{N} and the nodes {ξk}k=0N\{\xi_{k}\}_{k=0}^{N} can be computed rapidly and stably by computing the eigendecomposition of the Jacobi matrix for the orthonormal polynomials PP, as in the Golub–Welsch algorithm (?). However, if PP is an orthonormal polynomial basis which enjoys fast transforms from coefficients to values and back, and fast computation of Gaussian quadrature nodes (Jacobi polynomials, for example (?)) then such algorithms can be used in place of the generic Golub–Welsch approach.

4 Examples of orthonormal systems

In this section we describe two systems Φ\Phi and their free Schrödinger evolutions Ψ(t)\Psi(t).

4.1 Hermite functions

Hermite functions

φn(x)=1(2nn!π1/2)1/2Hn(x)ex2/2,n+,\varphi_{n}(x)=\frac{1}{{(2^{n}n!\pi^{1/2})}^{1/2}}\mathrm{H}_{n}(x){\mathrm{e}}^{-x^{2}/2},\qquad n\in\mbox{\Bbb Z}_{+}, (4.1)

where Hn\mathrm{H}_{n} is the nnth Hermite polynomial, are eigenfunctions of the Fourier transform,

12πφn(ξ)eixξdξ=inφn(x),x,n+.\frac{1}{\sqrt{2\pi}}\int_{-\infty}^{\infty}\varphi_{n}(\xi){\mathrm{e}}^{{\mathrm{i}}x\xi}\,\mathrm{d}\xi={\mathrm{i}}^{n}\varphi_{n}(x),\qquad x\in\mbox{\Bbb R},\quad n\in\mbox{\Bbb Z}_{+}. (4.2)

Their orthonormality in L2()\mathrm{L}_{2}(\mbox{\Bbb R}) follows from that of the familiar Hermite polynomials (?, 18.3) in L2(;eξ2)\mathrm{L}_{2}(\mbox{\Bbb R};{\mathrm{e}}^{-\xi^{2}}), they obey the differential recurrence relation (3.2) with bn=n/2b_{n}=\sqrt{n/2} and the Cramér inequality |φn(x)|π1/4|\varphi_{n}(x)|\leq\pi^{-1/4}, xx\in\mbox{\Bbb R}.111They should not be confused with Hermite functions from (?, p. 84).

To derive the FSE Ψ={ψn}n=0\Psi=\{\psi_{n}\}_{n=0}^{\infty} we assume the atomistic setting ε=1\varepsilon=1: to translate to semiclassical setting, we will replace tt by εt\varepsilon t in the final formula. Our starting point is the standard generating function for Hermite polynomials,

n=0Hn(x)n!zn=e2xzz2\sum_{n=0}^{\infty}\frac{\mathrm{H}_{n}(x)}{n!}z^{n}={\mathrm{e}}^{2xz-z^{2}}

(?, 18.12.15). It now follows from (4.1) that

π1/4ex2/2n=0φn(x)n!(21/2z)n=e2xzz2\pi^{1/4}{\mathrm{e}}^{x^{2}/2}\sum_{n=0}^{\infty}\frac{\varphi_{n}(x)}{\sqrt{n!}}(2^{1/2}z)^{n}={\mathrm{e}}^{2xz-z^{2}}

or, replacing z21/2zz\rightarrow 2^{-1/2}z,

n=0φn(x)n!zn=π1/4exp(x22+21/2xzz22).\sum_{n=0}^{\infty}\frac{\varphi_{n}(x)}{\sqrt{n!}}z^{n}=\pi^{-1/4}\exp\!\left(-\frac{x^{2}}{2}+2^{1/2}xz-\frac{z^{2}}{2}\right)\!.

It now follows from (3.4) and (4.2) that

n=0ψn(x,t)n!(iz)n\displaystyle\sum_{n=0}^{\infty}\frac{\psi_{n}(x,t)}{\sqrt{n!}}({\mathrm{i}}z)^{n} =\displaystyle= 12πn=0znn!φn(ξ)ei(ξ2t+ξx)dξ\displaystyle\frac{1}{\sqrt{2\pi}}\sum_{n=0}^{\infty}\frac{z^{n}}{\sqrt{n!}}\int_{-\infty}^{\infty}\varphi_{n}(\xi){\mathrm{e}}^{{\mathrm{i}}(\xi^{2}t+\xi x)}\,\mathrm{d}\xi
=\displaystyle= 12π[n=0φn(ξ)n!zn]ei(ξ2t+ξx)dξ\displaystyle\frac{1}{\sqrt{2\pi}}\int_{-\infty}^{\infty}\left[\sum_{n=0}^{\infty}\frac{\varphi_{n}(\xi)}{\sqrt{n!}}z^{n}\right]\!{\mathrm{e}}^{{\mathrm{i}}(\xi^{2}t+\xi x)}\,\mathrm{d}\xi
=\displaystyle= 121/2π3/4exp(12ξ2+21/2ξz12z2+iξ2t+iξx)dξ\displaystyle\frac{1}{2^{1/2}\pi^{3/4}}\int_{-\infty}^{\infty}\exp\!\left(-\frac{1}{2}\xi^{2}+2^{1/2}\xi z-\frac{1}{2}z^{2}+{\mathrm{i}}\xi^{2}t+{\mathrm{i}}\xi x\right)\!\,\mathrm{d}\xi
=\displaystyle= 1π1/4(12it)1/2exp(z2+23/2ixzx2+2itz22(2it1)).\displaystyle\frac{1}{\pi^{1/4}(1-2{\mathrm{i}}t)^{1/2}}\exp\!\left(-\frac{z^{2}+2^{3/2}{\mathrm{i}}xz-x^{2}+2{\mathrm{i}}tz^{2}}{2(2{\mathrm{i}}t-1)}\right)\!.

We conclude that

n=0ψn(x,t)n!(iz)n=1π1/4(12it)1/2exp(x22(2it1))exp(21/2ixz2it1122it+12it1z2).\sum_{n=0}^{\infty}\frac{\psi_{n}(x,t)}{\sqrt{n!}}({\mathrm{i}}z)^{n}=\frac{1}{\pi^{1/4}(1-2{\mathrm{i}}t)^{1/2}}\exp\!\left(\frac{x^{2}}{2(2{\mathrm{i}}t-1)}\right)\exp\!\left(-\frac{2^{1/2}{\mathrm{i}}xz}{2{\mathrm{i}}t-1}-\frac{1}{2}\frac{2{\mathrm{i}}t+1}{2{\mathrm{i}}t-1}z^{2}\right)\!.

Set

X=x(1+4t2)1/2,Z=121/2(2it+12it1)1/2z,X=-\frac{x}{(1+4t^{2})^{1/2}},\qquad Z=\frac{1}{2^{1/2}}\left(\frac{2{\mathrm{i}}t+1}{2{\mathrm{i}}t-1}\right)^{\!1/2}z,

which satisfy,

2XZZ2=21/2ixz2it1122it+12it1z2,2XZ-Z^{2}=-\frac{2^{1/2}{\mathrm{i}}xz}{2{\mathrm{i}}t-1}-\frac{1}{2}\frac{2{\mathrm{i}}t+1}{2{\mathrm{i}}t-1}z^{2},

and we deduce, using again the generating function for Hermite polynomials, that

exp(21/2ixz2it1122it+12it1z2)=n=0Hn(X)n!Zn.\exp\!\left(-\frac{2^{1/2}{\mathrm{i}}xz}{2{\mathrm{i}}t-1}-\frac{1}{2}\frac{2{\mathrm{i}}t+1}{2{\mathrm{i}}t-1}z^{2}\right)=\sum_{n=0}^{\infty}\frac{\mathrm{H}_{n}(X)}{n!}Z^{n}.

All we thus need is to compare the powers of zz in

n=0ψn(x,t)n!(iz)n\displaystyle\sum_{n=0}^{\infty}\frac{\psi_{n}(x,t)}{\sqrt{n!}}({\mathrm{i}}z)^{n}
=\displaystyle= 1π1/4(12it)1/2exp(x22(2it1))n=0Hn(X)n!Zn\displaystyle\frac{1}{\pi^{1/4}(1-2{\mathrm{i}}t)^{1/2}}\exp\!\left(\frac{x^{2}}{2(2{\mathrm{i}}t-1)}\right)\sum_{n=0}^{\infty}\frac{\mathrm{H}_{n}(X)}{n!}Z^{n}
=\displaystyle= 1π1/4(12it)1/2exp(x22(2it1))n=01n!Hn(x(1+4t2)1/2)(121/2(2it+12it1)1/2z)n.\displaystyle\frac{1}{\pi^{1/4}(1-2{\mathrm{i}}t)^{1/2}}\exp\!\left(\frac{x^{2}}{2(2{\mathrm{i}}t-1)}\right)\sum_{n=0}^{\infty}\frac{1}{n!}\mathrm{H}_{n}\!\left(\!-\frac{x}{(1+4t^{2})^{1/2}}\right)\!\!\left(\frac{1}{2^{1/2}}\left(\frac{2{\mathrm{i}}t\!+\!1}{2{\mathrm{i}}t\!-\!1}\right)^{\!1/2}\!z\!\right)^{\!n}\!\!\!.

The outcome is

ψn(x,t)=in(2nn!π1/2)1/2(12it)1/2exp(x22(2it1))(2it+12it1)n/2Hn(x(1+4t2)1/2).\psi_{n}(x,t)=\frac{{\mathrm{i}}^{n}}{(2^{n}n!\pi^{1/2})^{1/2}(1-2{\mathrm{i}}t)^{1/2}}\exp\!\left(\frac{x^{2}}{2(2{\mathrm{i}}t-1)}\right)\!\left(\frac{2{\mathrm{i}}t+1}{2{\mathrm{i}}t-1}\right)^{\!n/2}\!\mathrm{H}_{n}\!\left(\frac{x}{(1+4t^{2})^{1/2}}\right)\!.

Finally, since

Hn(x(1+4t2)1/2)=(2nn!)1/2π1/4exp(x22(1+4t2))φn(x(1+4t2)1/2),\mathrm{H}_{n}\!\left(\frac{x}{(1+4t^{2})^{1/2}}\right)=(2^{n}n!)^{1/2}\pi^{1/4}\exp\!\left(\frac{x^{2}}{2(1+4t^{2})}\right)\!\varphi_{n}\!\left(\frac{x}{(1+4t^{2})^{1/2}}\right)\!,

we deduce, restoring the semiclassical setting, that

Lemma 5

The explicit form of the Hermite FSE is

ψn(x,t)=(1+2iεt)n/2(12iεt)(n+1)/2exp(itεx21+4ε2t2)φn(x(1+4ε2t2)1/2).\psi_{n}(x,t)=\frac{(1+2{\mathrm{i}}\varepsilon t)^{n/2}}{(1-2{\mathrm{i}}\varepsilon t)^{(n+1)/2}}\exp\!\left(-\frac{{\mathrm{i}}t\varepsilon x^{2}}{1+4\varepsilon^{2}t^{2}}\right)\!\varphi_{n}\!\left(\frac{x}{(1+4\varepsilon^{2}t^{2})^{1/2}}\right)\!. (4.3)

Moreover, the functions ψn\psi_{n} are subject to the bound

|ψn(x,t)|1[π(1+4ε2t2)]1/4,t0,x.|\psi_{n}(x,t)|\leq\frac{1}{[\pi(1+4\varepsilon^{2}t^{2})]^{1/4}},\qquad t\geq 0,\;\;x\in\mbox{\Bbb R}. (4.4)

Proof The expression (4.3) follows from the preceding analysis, while (4.4) is an immediate consequence of the Cramér inequality.        \Box

Refer to caption
Figure 4.1: The Hermite FSE: the functions |ψn(x,t)||\psi_{n}(x,t)| for n=0,,3n=0,\ldots,3, x[12,12]x\in[-12,12] and t[0,4]t\in[0,4].

Fig. 4.1 displays the magnitude of the first six ψn\psi_{n}s. It is evident that they are consistent with the inequality (4.4). There are two facts to bear in mind. Firstly, examining the modulus hides the oscillations in (4.3): in reality, the ψn\psi_{n}s are considerably more violent. Secondly, while the functions ψn\psi_{n} appear to spread energy and |ψn||\psi_{n}| seems to approach a steady steady, in reality we are interested only is small values of tt, a single time step, so that t=h=𝒪(ε1/2)t=h={\cal O}\!\left(\varepsilon^{1/2}\right).

An implementation of FSEs based on Hermite functions necessitates in each time step the expansion of the initial value in Hermite functions. There exist powerful algorithms to this end, many based upon the fast multipole algorithm and generalisable to higher spatial dimensions (?).

Lemma 6

The Hermite FSE in Lemma 5 satisfy the three term recurrence,

xψn(x,t)=n2(1+2iεt12iεt)12ψn1(x,t)+n+12(12iεt1+2iεt)12ψn+1(x,t).x\psi_{n}(x,t)=\sqrt{\frac{n}{2}}\left(\frac{1+2{\mathrm{i}}\varepsilon t}{1-2{\mathrm{i}}\varepsilon t}\right)^{\tfrac{1}{2}}\psi_{n-1}(x,t)+\sqrt{\frac{n+1}{2}}\left(\frac{1-2{\mathrm{i}}\varepsilon t}{1+2{\mathrm{i}}\varepsilon t}\right)^{\tfrac{1}{2}}\psi_{n+1}(x,t). (4.5)

This three term recurrence allows us to evaluate finite expansions in this basis in a stable manner using Clenshaw’s algorithm (?).

4.2 Malmquist–Takenaka functions

The Malmquist–Takenaka system is a complex-valued rational basis of L2()\mathrm{L}_{2}(\mbox{\Bbb R}), introduced independently by Malmquist and Takenaka and repeatedly rediscovered: we refer to (?) for its brief history. It is instructive to introduce them within the narrative of Subsection 2.1, while extending it to complex-valued bases. The starting point is the Laguerre measure eξdξ{\mathrm{e}}^{-\xi}\,\mathrm{d}\xi, ξ0\xi\geq 0. We can use (3.2) to generate an orthonormal system on the real line but this system is not dense in L2()\mathrm{L}_{2}(\mbox{\Bbb R}). It is a basis of 𝒫𝒲[0,)()\mathcal{PW}_{[0,\infty)}(\mbox{\Bbb R}), of fL2()f\in\mathrm{L}_{2}(\mbox{\Bbb R}) whose Fourier transform is supported inside [0,)[0,\infty). To recover the orthogonal complement of 𝒫𝒲[0,)()\mathcal{PW}_{[0,\infty)}(\mbox{\Bbb R}) in L2()\mathrm{L}_{2}(\mbox{\Bbb R}), namely 𝒫𝒲(,0]()\mathcal{PW}_{(-\infty,0]}(\mbox{\Bbb R}), thereby ensuring that the system is dense in L2()\mathrm{L}_{2}(\mbox{\Bbb R}), we need to complement it by the orthonormal system generated by the measure eξdξ{\mathrm{e}}^{\xi}\,\mathrm{d}\xi for ξ(,0]\xi\in(-\infty,0] which, conveniently, we label by φn\varphi_{n}, n1n\leq-1. The outcome, the MT system, is

φn(x)=2πin(1+2ix)n(12ix)n+1,n,\varphi_{n}(x)=\sqrt{\frac{2}{\pi}}{\mathrm{i}}^{n}\frac{(1+2{\mathrm{i}}x)^{n}}{(1-2{\mathrm{i}}x)^{n+1}},\qquad n\in\mbox{\Bbb Z}, (4.6)

(?). The MT system has a number of elegant features:

φn\displaystyle\varphi_{n}^{\prime} =\displaystyle= nφn1+i(2n+1)φn+(n+1)φn+1,n,\displaystyle-n\varphi_{n-1}+{\mathrm{i}}(2n+1)\varphi_{n}+(n+1)\varphi_{n+1},\qquad n\in\mbox{\Bbb Z},
|φn(x)|\displaystyle|\varphi_{n}(x)| \displaystyle\leq 2π1(1+4x2)1/2,x,\displaystyle\sqrt{\frac{2}{\pi}}\frac{1}{(1+4x^{2})^{1/2}},\qquad x\in\mbox{\Bbb R},
φmφn\displaystyle\varphi_{m}\varphi_{n} =\displaystyle= 12π(φm+niφm+n+1),m,n,\displaystyle\frac{1}{\sqrt{2\pi}}(\varphi_{m+n}-{\mathrm{i}}\varphi_{m+n+1}),\qquad m,n\in\mbox{\Bbb Z},
2xφn\displaystyle 2x\varphi_{n}^{\prime} =\displaystyle= inφn1φni(n+1)φn+1,\displaystyle-{\mathrm{i}}n\varphi_{n-1}-\varphi_{n}-{\mathrm{i}}(n+1)\varphi_{n+1},
φn+1(x)\displaystyle\varphi_{n+1}(x) =\displaystyle= i(1+2ix12ix)φn(x),\displaystyle{\mathrm{i}}\left(\frac{1+2{\mathrm{i}}x}{1-2{\mathrm{i}}x}\right)\varphi_{n}(x),
φ1n(x)\displaystyle\varphi_{-1-n}(x) =\displaystyle= i2n1φn(x).\displaystyle{\mathrm{i}}^{2n-1}\varphi_{n}(-x).

– which make its implementation as a spectral basis considerably easier. However, the most valuable feature of the MT system is that, subject to the change of variables x=12tan(θ/2)x=\frac{1}{2}\tan(\theta/2), we have

f^n=f(x)φn(x)¯dx=(i)n2πππ(1itanθ2)f(12tanθ2)einθdθ,n.\hat{f}_{n}=\int_{-\infty}^{\infty}f(x)\overline{\varphi_{n}(x)}\,\mathrm{d}x=\frac{(-{\mathrm{i}})^{n}}{\sqrt{2\pi}}\int_{-\pi}^{\pi}\!\left(1-{\mathrm{i}}\tan\frac{\theta}{2}\right)\!f\!\left(\frac{1}{2}\tan\frac{\theta}{2}\right)\!{\mathrm{e}}^{-{\mathrm{i}}n\theta}\,\mathrm{d}\theta,\qquad n\in\mbox{\Bbb Z}. (4.7)

In other words, the computation of expansion coefficients is equivalent to the evaluation of standard Fourier coefficients of a modified function, a task that can be accomplished (for sufficiently smooth functions) to very high accuracy using the Fast Fourier Transform.

We note in passing that this feature – the computation of the first NN expansion coefficients in 𝒪(NlogN){\cal O}\!\left(N\log N\right) operations – is highly unusual in the setting of Section 3.1: it can be accomplished only for the MT basis (or its minor generalisation) using FFT and for four other ‘tanh-Chebyshev’ bases using Fast Cosine (or Sine) Transform (?).

Let us now investigate the FSEs Ψ(t)\Psi(t). For simplicity we consider this only for n+n\in\mbox{\Bbb Z}_{+}, noting that an extension to n1n\leq-1 is straightforward by the symmetry: ψ1n(x,t)=i2n1ψn(x,t)\psi_{-1-n}(x,t)={\mathrm{i}}^{2n-1}\psi_{n}(-x,t). As before, we assume for the time being that ε=1\varepsilon=1. Using (3.2) we have

ψn(x,t)=(i)n2π0Ln(ξ)exp(ξ2+itξ2+ixξ)dξ,n+,\psi_{n}(x,t)=\frac{(-{\mathrm{i}})^{n}}{\sqrt{2\pi}}\int_{0}^{\infty}\mathrm{L}_{n}(\xi)\exp\!\left(-\tfrac{\xi}{2}+{\mathrm{i}}t\xi^{2}+{\mathrm{i}}x\xi\right)\!\,\mathrm{d}\xi,\qquad n\in\mbox{\Bbb Z}_{+}, (4.8)

where Ln\mathrm{L}_{n} is the nnth Laguerre polynomial. We can remove the oscillation from exp(ixξ)\exp({\mathrm{i}}x\xi) by deforming the contour of integration to the line {z:Im(z)=2xRe(z)}\{z\in\mbox{\Bbb C}:\mathrm{Im}(z)=2x\mathrm{Re}(z)\} (technically by integrating over a the boundary of a sector of radius RR, where the contribution from the arc decays exponentially in RR), which yields the formula

ψn(x,t)=2π(i)n12ix0Ln(2s12ix)exp(4its2(12ix)2)esds.\psi_{n}(x,t)=\sqrt{\frac{2}{\pi}}\frac{(-{\mathrm{i}})^{n}}{1-2{\mathrm{i}}x}\int_{0}^{\infty}\mathrm{L}_{n}\left(\frac{2s}{1-2{\mathrm{i}}x}\right)\exp\left(\frac{4{\mathrm{i}}ts^{2}}{(1-2{\mathrm{i}}x)^{2}}\right){\mathrm{e}}^{-s}\mathrm{d}s. (4.9)

For small values of tt, this integral is not particularly oscillatory, since Re()\mathrm{Re}\left(\right) It is possible to produce for any specific value of nn, e.g.

ψ0(x,t)\displaystyle\psi_{0}(x,t) =\displaystyle= i8texp((2x+i)216it)erfc((2x+i)16it),\displaystyle\sqrt{\frac{{\mathrm{i}}}{8t}}\exp\!\left(\frac{(2x+{\mathrm{i}})^{2}}{16{\mathrm{i}}t}\right)\mathrm{erfc}\!\left(\frac{(2x+{\mathrm{i}})}{\sqrt{16{\mathrm{i}}t}}\right),
ψ1(x,t)\displaystyle\psi_{1}(x,t) =\displaystyle= iψ0(x,t)+(12ix)ψ0(x,t)ψ0(x,0)4t.\displaystyle-{\mathrm{i}}\psi_{0}(x,t)+\left(1-2{\mathrm{i}}x\right)\frac{\psi_{0}(x,t)-\psi_{0}(x,0)}{4t}.

There is no need to fear the power of tt in the denominator, which cancels as t0t\rightarrow 0. We discuss handling this removable singularity in the next subsection.

Fig. 4.2 displays |ψn||\psi_{n}|, n=0,,5n=0,\ldots,5, for the MT functions in a setting identical to Fig. 4.1. Note that the magnitude for small t>0t>0 varies much more violently for x>0x>0 – obviously, this is reversed for n1n\leq-1 – and that, like for the Hermite FSE, the magnitude tends to an increasingly regular profile once tt grows.

4.3 A four-term recurrence for the Malmquist–Takenaka FSE basis

While a closed form expression of the ψn\psi_{n}s is complicated and not clearly even possible, we can derive a useful recurrence formula. Begin from the following differential difference equation for the Laguerre polynomials (which follows by differentiating (?, 18.17.1)),

Ln(ξ)=Ln(ξ)Ln+1(ξ).\mathrm{L}_{n}(\xi)=\mathrm{L}_{n}^{\prime}(\xi)-\mathrm{L}_{n+1}^{\prime}(\xi). (4.10)

From this it follows immediately that,

ψn(x,t)=(i)n2π0(Ln(ξ)Ln+1(ξ))exp(ξ2+itξ2+ixξ)dξ.\psi_{n}(x,t)=\frac{(-{\mathrm{i}})^{n}}{\sqrt{2\pi}}\int_{0}^{\infty}\left(\mathrm{L}_{n}^{\prime}(\xi)-\mathrm{L}_{n+1}^{\prime}(\xi)\right)\exp\!\left(-\tfrac{\xi}{2}+{\mathrm{i}}t\xi^{2}+{\mathrm{i}}x\xi\right)\!\,\mathrm{d}\xi. (4.11)

Integrating by parts, noting that Ln(0)=1=Ln+1(0)\mathrm{L}_{n}(0)=1=\mathrm{L}_{n+1}(0) so the boundary terms vanish,

ψn(x,t)=(i)n2π0(Ln+1(ξ)Ln(ξ))(2itξ+ix12)exp(ξ2+itξ2+ixξ)dξ.\psi_{n}(x,t)=\frac{(-{\mathrm{i}})^{n}}{\sqrt{2\pi}}\int_{0}^{\infty}\left(\mathrm{L}_{n+1}(\xi)-\mathrm{L}_{n}(\xi)\right)(2{\mathrm{i}}t\xi+{\mathrm{i}}x-\tfrac{1}{2})\exp\!\left(-\tfrac{\xi}{2}+{\mathrm{i}}t\xi^{2}+{\mathrm{i}}x\xi\right)\!\,\mathrm{d}\xi. (4.12)

We can then use the three-term recurrence,

(n+1)Ln+1(ξ)=(2n+1ξ)Ln(ξ)nLn1(ξ),(n+1)\mathrm{L}_{n+1}(\xi)=(2n+1-\xi)\mathrm{L}_{n}(\xi)-n\mathrm{L}_{n-1}(\xi), (4.13)

to obtain,

ψn(x,t)\displaystyle\psi_{n}(x,t) =\displaystyle= (i)n2π0[2it((2n+3)Ln+1(n+1)Ln(n+2)Ln+2)\displaystyle\frac{(-{\mathrm{i}})^{n}}{\sqrt{2\pi}}\int_{0}^{\infty}\bigg{[}2{\mathrm{i}}t\left((2n+3)\mathrm{L}_{n+1}-(n+1)\mathrm{L}_{n}-(n+2)\mathrm{L}_{n+2}\right)
2it((2n+1)LnnLn1(n+1)Ln+1)\displaystyle\qquad\qquad-2{\mathrm{i}}t\left((2n+1)\mathrm{L}_{n}-n\mathrm{L}_{n-1}-(n+1)\mathrm{L}_{n+1}\right)
+(ix12)(Ln+1Ln)]exp(ξ2+itξ2+ixξ)dξ\displaystyle\qquad\qquad+({\mathrm{i}}x-\tfrac{1}{2})(\mathrm{L}_{n+1}-\mathrm{L}_{n})\bigg{]}\exp\!\left(-\tfrac{\xi}{2}+{\mathrm{i}}t\xi^{2}+{\mathrm{i}}x\xi\right)\!\,\mathrm{d}\xi
=\displaystyle= (i)n2π0[2it(n+2)Ln+2+(2it(3n+4)+ix12)Ln+1\displaystyle\frac{(-{\mathrm{i}})^{n}}{\sqrt{2\pi}}\int_{0}^{\infty}\bigg{[}-2{\mathrm{i}}t(n+2)\mathrm{L}_{n+2}+(2{\mathrm{i}}t(3n+4)+{\mathrm{i}}x-\tfrac{1}{2})\mathrm{L}_{n+1}
(2it(3n+2)+ix12)Ln+2itnLn1)]exp(ξ2+itξ2+ixξ)dξ\displaystyle-(2{\mathrm{i}}t(3n+2)+{\mathrm{i}}x-\tfrac{1}{2})\mathrm{L}_{n}+2{\mathrm{i}}tn\mathrm{L}_{n-1})\bigg{]}\exp\!\left(-\tfrac{\xi}{2}+{\mathrm{i}}t\xi^{2}+{\mathrm{i}}x\xi\right)\!\,\mathrm{d}\xi
=\displaystyle= (i)n2π0[(i)22it(n+2)Ln+2(i)(2t(3n+4)+x+12i)Ln+1\displaystyle\frac{(-{\mathrm{i}})^{n}}{\sqrt{2\pi}}\int_{0}^{\infty}\bigg{[}(-{\mathrm{i}})^{2}2{\mathrm{i}}t(n+2)\mathrm{L}_{n+2}-(-{\mathrm{i}})(2t(3n+4)+x+\tfrac{1}{2}{\mathrm{i}})\mathrm{L}_{n+1}
(2it(3n+2)+ix12)Ln+(i)12tnLn1)]exp(ξ2+itξ2+ixξ)dξ\displaystyle-(2{\mathrm{i}}t(3n+2)+{\mathrm{i}}x-\tfrac{1}{2})\mathrm{L}_{n}+(-{\mathrm{i}})^{-1}2tn\mathrm{L}_{n-1})\bigg{]}\exp\!\left(-\tfrac{\xi}{2}+{\mathrm{i}}t\xi^{2}+{\mathrm{i}}x\xi\right)\!\,\mathrm{d}\xi
=\displaystyle= 2it(n+2)ψn+2(x,t)(2t(3n+4)+x+12i)ψn+1(x,t)\displaystyle 2{\mathrm{i}}t(n+2)\psi_{n+2}(x,t)-(2t(3n+4)+x+\tfrac{1}{2}{\mathrm{i}})\psi_{n+1}(x,t)
(2it(3n+2)+ix12)ψn(x,t)+2tnψn1(x,t).\displaystyle-(2{\mathrm{i}}t(3n+2)+{\mathrm{i}}x-\tfrac{1}{2})\psi_{n}(x,t)+2tn\psi_{n-1}(x,t).

Collecting terms yields,

2it(n+2)ψn+2=(2t(3n+4)+x+12i)ψn+1+(2it(3n+2)+ix+12)ψn2tnψn1.\displaystyle 2{\mathrm{i}}t(n+2)\psi_{n+2}=(2t(3n+4)+x+\tfrac{1}{2}{\mathrm{i}})\psi_{n+1}+(2{\mathrm{i}}t(3n+2)+{\mathrm{i}}x+\tfrac{1}{2})\psi_{n}-2tn\psi_{n-1}.

We now undo the assignment ε=1\varepsilon=1 to obtain the following lemma.

Lemma 7

The FSE corresponding to the MT system obeys the recurrence for n1n\geq 1,

ψ0(x,t)\displaystyle\psi_{0}(x,t) =\displaystyle= i8εtexp((2x+i)216iεt)erfc((2x+i)16iεt),\displaystyle\sqrt{\frac{{\mathrm{i}}}{8\varepsilon t}}\exp\!\left(\frac{(2x+{\mathrm{i}})^{2}}{16{\mathrm{i}}\varepsilon t}\right)\mathrm{erfc}\!\left(\frac{(2x+{\mathrm{i}})}{\sqrt{16{\mathrm{i}}\varepsilon t}}\right),
ψ1(x,t)\displaystyle\psi_{1}(x,t) =\displaystyle= iψ0+(12ix)ψ0(x,t)ψ0(x,0)4εt\displaystyle-{\mathrm{i}}\psi_{0}+\left(1-2{\mathrm{i}}x\right)\frac{\psi_{0}(x,t)-\psi_{0}(x,0)}{4\varepsilon t}
i(n+1)ψn+1\displaystyle{\mathrm{i}}(n+1)\psi_{n+1} =\displaystyle= (3n+1+2x+i4εt)ψn+i(3n1+2xi4εt)ψn1(n1)ψn2.\displaystyle\left(3n+1+\frac{2x+{\mathrm{i}}}{4\varepsilon t}\right)\psi_{n}+{\mathrm{i}}\left(3n-1+\frac{2x-{\mathrm{i}}}{4\varepsilon t}\right)\psi_{n-1}-(n-1)\psi_{n-2}.
Refer to caption
Figure 4.2: The Malmquist–Takenaka FSE: the functions |ψn(x,t)||\psi_{n}(x,t)| for n=0,,3n=0,\ldots,3, x[20,20]x\in[-20,20] and t[0,4]t\in[0,4]

Lemma 7 indicates the possibility of computing an expansion in the Malmquist–Takenaka FSE basis using the (generalized) Clenshaw algorithm (?). The functions ψn\psi_{n} for n1n\leq-1 can be addressed using the symmetry ψ1n(x,t)=i2n1ψn(x,t)\psi_{-1-n}(x,t)={\mathrm{i}}^{2n-1}\psi_{n}(-x,t), which we omit. Clenshaw’s algorithm is best known to apply to bases satisfying three-term recurrences, and in the case of a two-term recurrence reduces to Horner’s algorithm. The following lemma spells out the Clenshaw algorithm for a basis with a four-term recurrence (such as the Malmquist–Takenaka FSE).

Lemma 8

Let Φ={φn}n=0\Phi=\{\varphi_{n}\}_{n=0}^{\infty} be a basis which satisfies the four-term recurrence,

φn+1(x)=An(x)φn(x)+Bn(x)φn1(x)+Cn(x)φn2(x),\varphi_{n+1}(x)=A_{n}(x)\varphi_{n}(x)+B_{n}(x)\varphi_{n-1}(x)+C_{n}(x)\varphi_{n-2}(x), (4.14)

for n1n\geq 1, where C1(x)=0C_{1}(x)=0, then the finite expansion,

f(x)=n=0Nanφn(x),f(x)=\sum_{n=0}^{N}a_{n}\varphi_{n}(x),

is equal to

v0(x)φ0(x)+v1(x)φ1(x),v_{0}(x)\varphi_{0}(x)+v_{1}(x)\varphi_{1}(x),

where 𝐯(x)=(v0(x),v1(x),,vN(x))\mbox{\boldmath$v$\unboldmath}(x)=(v_{0}(x),v_{1}(x),\ldots,v_{N}(x))^{\top} satisfies the backwards recurrence,

vN(x)\displaystyle v_{N}(x) =\displaystyle\!\!\!=\!\!\! aN\displaystyle a_{N}
vN1(x)\displaystyle v_{N-1}(x) =\displaystyle\!\!\!=\!\!\! aN1+AN1(x)vN(x)\displaystyle a_{N-1}+A_{N-1}(x)v_{N}(x)
vN2(x)\displaystyle v_{N-2}(x) =\displaystyle\!\!\!=\!\!\! aN2+AN2(x)vN1(x)+BN1(x)vN(x)\displaystyle a_{N-2}+A_{N-2}(x)v_{N-1}(x)+B_{N-1}(x)v_{N}(x)
vn(x)\displaystyle v_{n}(x) =\displaystyle\!\!\!=\!\!\! an+An(x)vn+1(x)+Bn+1(x)vn+2(x)+Cn+2(x)vn+3(x),\displaystyle a_{n}+A_{n}(x)v_{n+1}(x)+B_{n+1}(x)v_{n+2}(x)+C_{n+2}(x)v_{n+3}(x),

for n=N3,N4,0n=N-3,N-4,\ldots 0. Where we set A0=0A_{0}=0.

Proof We follow the derivation of Clenshaw’s algorithm in (?), but with an extra band below the diagonal in the associated linear system. Indeed, the vector 𝝋(x)=(φ0(x),,φN(x))T\boldsymbol{\varphi}(x)=(\varphi_{0}(x),\ldots,\varphi_{N}(x))^{T} satisfies

(1A01B1A11C2B2A210C3B3A310CN1BN1AN11)(φ0(x)φ1(x)φ2(x)φ3(x)φ4(x)φN(x))=(φ0(x)φ1(x)0000),\begin{pmatrix}1&&&&&&\\ -A_{0}&1&&&&&\\ -B_{1}&-A_{1}&1&&&&\\ -C_{2}&-B_{2}&-A_{2}&1&&&\\ 0&-C_{3}&-B_{3}&-A_{3}&1&&\\ &\ddots&\ddots&\ddots&\ddots&\ddots&\\ &&0&-C_{N-1}&-B_{N-1}&-A_{N-1}&1\end{pmatrix}\begin{pmatrix}\varphi_{0}(x)\\ \varphi_{1}(x)\\ \varphi_{2}(x)\\ \varphi_{3}(x)\\ \varphi_{4}(x)\\ \vdots\\ \varphi_{N}(x)\end{pmatrix}=\begin{pmatrix}\varphi_{0}(x)\\ \varphi_{1}(x)\\ 0\\ 0\\ 0\\ \vdots\\ 0\end{pmatrix},

since A0=0A_{0}=0. Let us write this as L(x)𝝋(x)=𝝆(x)L(x)\boldsymbol{\varphi}(x)=\boldsymbol{\rho}(x). Clearly L(x)L(x) is invertible, so

f(x)=𝒂T𝝋(x)=𝒂TL(x)1𝝆(x)=(L(x)T𝒂)T𝝆(x).f(x)=\boldsymbol{a}^{T}\boldsymbol{\varphi}(x)=\boldsymbol{a}^{T}L(x)^{-1}\boldsymbol{\rho}(x)=\left(L(x)^{-T}\boldsymbol{a}\right)^{T}\boldsymbol{\rho}(x).

The result is proved by noting that the backward recurrence for 𝒗(x)\boldsymbol{v}(x) merely computes 𝒗(x)=L(x)T𝒂\boldsymbol{v}(x)=L(x)^{-T}\boldsymbol{a} by back substitution.        \Box

In order to evaluate ψ0\psi_{0} without trouble from the removable singularity, we rewrite equation (4.2) in the form

ψ0(x,t)=φ0(x)G0(2ix116iεt),\psi_{0}(x,t)=\varphi_{0}(x)G_{0}\left(\frac{2{\mathrm{i}}x-1}{\sqrt{16{\mathrm{i}}\varepsilon t}}\right), (4.15)

where G0(z)=iπzez2erfc(iz)G_{0}(z)=-{\mathrm{i}}\sqrt{\pi}z{\mathrm{e}}^{-z^{2}}\mathrm{erfc}(-{\mathrm{i}}z). This function is related to w(z)=ez2erfc(iz)w(z)={\mathrm{e}}^{-z^{2}}\mathrm{erfc(-{\mathrm{i}}z)}, known as the Faddeeva function or plasma dispersion function (??). Note that x,tx,t\in\mbox{\Bbb R} corresponds to evaluating G0G_{0} in the complex plane in the sector {z:arg(z)(π/4,5π/4)}\{z\in\mbox{\Bbb C}:\mathrm{arg}(z)\in(\pi/4,5\pi/4)\} and we are particularly interested in small positive tt, which corresponds to large zz in this sector. The fact that G0(z)1G_{0}(z)\to 1 as |z||z|\to\infty within this sector shows the recovery of φ0(x)\varphi_{0}(x) as t0t\to 0.

Following (??), the following continued fraction for G0G_{0} at z=z=\infty is convergent in the upper half-plane (?, 7.9.3),

G0(z)=1112z21z2132z212z21.G_{0}(z)=\cfrac{1}{1-\cfrac{\tfrac{1}{2}z^{-2}}{1-\cfrac{z^{-2}}{1-\cfrac{\tfrac{3}{2}z^{-2}}{1-\cfrac{2z^{-2}}{1-\cdots}}}}}. (4.16)

Truncating this continued fraction yields an extremely good approximation for large zz in the upper half-plane, and for the lower half-plane we can use the relation (?, 7.4.3)

G0(z)=G0(z)2iπzez2,G_{0}(z)=G_{0}(-z)-2{\mathrm{i}}\sqrt{\pi}z{\mathrm{e}}^{-z^{2}}, (4.17)

but note that accuracy can be lost near the complex roots of erfc(iz)\mathrm{erfc}(-{\mathrm{i}}z) since it relies on heavy cancellation (??).

In order to evaluate ψ1\psi_{1} without trouble from the removable singularity, we rewrite the formula in Lemma 7 in the form

ψ1(x,t)=iψ0(x,t)+2π2i(12ix)2G1(2ix116iεt),\psi_{1}(x,t)=-{\mathrm{i}}\psi_{0}(x,t)+\sqrt{\frac{2}{\pi}}\frac{2{\mathrm{i}}}{(1-2{\mathrm{i}}x)^{2}}G_{1}\left(\frac{2{\mathrm{i}}x-1}{\sqrt{16{\mathrm{i}}\varepsilon t}}\right), (4.18)

where G1(z)=2z2(G0(z)1)G_{1}(z)=2z^{2}(G_{0}(z)-1). While this covers the evaluation of ψ0(x,t)\psi_{0}(x,t) and ψ1(x,t)\psi_{1}(x,t) for small tt, the full implementation of Clenshaw’s algorithm may still experience loss of numerical accuracy due to the 1/t1/t terms in the recurrence relation. However, numerical issues like this are beyond the scope of this paper.

5 Bringing the elements together

We bring together the different results of the paper into a cohesive whole. In Section 2, we reduced the problem of solving the semiclassical Schrödinger equation to combining time-steps of the form,

uk+1(x)=euk(x),u^{k+1}(x)={\mathrm{e}}^{\mathcal{R}_{\ell}}u^{k}(x),

where

0\displaystyle\mathcal{R}_{0} =\displaystyle= 12τε1V,\displaystyle-\frac{1}{2}\tau\varepsilon^{-1}V,
1\displaystyle\mathcal{R}_{1} =\displaystyle= 12τεx2,\displaystyle\frac{1}{2}\tau\varepsilon\partial_{x}^{2},
2\displaystyle\mathcal{R}_{2} =\displaystyle= 112τ3ε{x2[V(2)]+V(2)x2}+124τ3ε1(V(1))2,\displaystyle\color[rgb]{0,0,0}\frac{1}{12}\tau^{3}\varepsilon\left\{\partial_{x}^{2}[V^{(2)}\,\cdot\,]+V^{(2)}\partial_{x}^{2}\right\}+\frac{1}{24}\tau^{3}\varepsilon^{-1}(V^{(1)})^{2},\ldots

where τ=ih\tau={\mathrm{i}}h and =𝒪(h21ε1)\mathcal{R}_{\ell}={\cal O}\!\left(h^{2\ell-1}\varepsilon^{-1}\right) for =1,2,\ell=1,2,\ldots. We propose that the numerical solution be represented implicitly by

uk(x)=n=0Nu^nφn(x),u^{k}(x)=\sum_{n=0}^{N}\hat{u}_{n}\varphi_{n}(x),

where φn\varphi_{n} is either the Hermite function basis or the Malmquist–Takenaka basis (in the latter case the indices should extend from n=N1n=-N-1 to n=Nn=N). However, explicitly, we propose that the numerical solution be represented by its values on a grid appropriate to the basis. When this basis is Hermite functions, those points are Hermite quadrature points, and for Malmquist–Takenaka functions, those points are mapped equi-spaced points (?),

xj[N]=12tan(θj[N]/2),j=N1,N,x_{j}^{[N]}=\tfrac{1}{2}\tan\left(\theta_{j}^{[N]}/2\right),\qquad j=-N-1,\ldots N, (5.19)
θj[N]=jπN+1,j=N1,N.\theta_{j}^{[N]}=\frac{j\pi}{N+1},\qquad j=-N-1,\ldots N. (5.20)

We call these Malmquist–Takenaka points or MT points.

The reason for these choices of grid points are three-fold. First, the mapping from the values of a finite expansion in the basis at these specific grid points, weighted appropriately, to the coefficients in the finite expansion is unitary, so is invertible and perfectly stable. Second, there are known algorithms to compute this mapping and its inverse, which in the case of Malmquist–Takenaka, can be performed rapidly by the Fast Fourier Transform (FFT) and its inverse. Thirdly, at the end of a full time step, we have the solution given by its values on this grid. The computation of the values of the solution at arbitrary points on the real line can be performed stably by barycentric interpolation formula. The barycentric weights for Hermite quadrature points and for equispaced points on the unit circle (which map to MT points) are known explicitly (??).

When our solution is represented by values at the grid points, the case =0\ell=0 is straightforward — we simply multiply the function value at gridpoint xk[N]x_{k}^{[N]} by exp(12τε1V(xk[N]))\exp(-\tfrac{1}{2}\tau\varepsilon^{-1}V(x_{k}^{[N]})).

The case =1\ell=1 is more subtle, and we propose using the free Schrödinger evolutions developed in Section 3. We first compute the coefficients in the Φ\Phi basis, and then evaluate linear combination of those coefficients with the free Schrödinger evolution Ψ(12hε)\Psi(\tfrac{1}{2}h\varepsilon) at the grid points. This is a two-step process, as follows.

  • Compute the coefficients, a0,a1,,aNa_{0},a_{1},\ldots,a_{N} (in the Φ\Phi basis, indexed from N1-N-1 to NN in the case of the MT basis) from the values on the grid (using the FFT in the case of the MT basis)

  • Evaluate the sum k=0Nakψk(x,12hε)\sum_{k=0}^{N}a_{k}\psi_{k}(x,\tfrac{1}{2}h\varepsilon) at the grid points using Clenshaw’s algorithm (in the case of the MT basis, using the 4 term version in Lemma 8).

In the case 2\ell\geq 2 we propose the use of Krylov subspace methods. This was first proposed in (?), later generalised to time-dependent potentials (??) as well as the method of quasi-Magnus exponential integrators of (?). There are two facts which make this approach work well. First, =𝒪(h21ε1)\mathcal{R}_{\ell}={\cal O}\!\left(h^{2\ell-1}\varepsilon^{-1}\right) for >1\ell>1, so we are computing the exponential of a matrix which is small in spectral norm. As a result, a Krylov subspace with a miniscule dimension can be used (?). Second, the sparse differentiation matrix (see (3.1)) implies that the matrices which must be applied to a vector in the Krylov subspace method are a sum of compositions of: diagonal matrices coming from derivatives of the potential function VV, pentadiagonal matrices coming from the discretisation of x2\partial_{x}^{2} in coefficient space, and transforms between function values on the grid and coefficients (which can be performed using the FFT in the case of the MT basis).

Acknowledgments

This work is partially supported by the Simons Foundation Award No 663281 granted to the Institute of Mathematics of the Polish Academy of Sciences for the years 2021-2023

The authors thank the Isaac Newton Institute for Mathematical Sciences for support and hospitality during the programme “Geometry, compatibility and structure preservation in computational differential equations”, supported by EPSRC grant EP/R014604/1, where this work has been initiated.

Katharina Schratz has received funding from the European Research Council (ERC) under the European Unions Horizon 2020 research and innovation programme (grant agreement No. 850941)

The work of Karolina Kropielnicka and of Marcus Webb in this project was financed by The National Center for Science (NCN), based on Grant No. 2019/34/E/ST1/00390

References

  • [1]
  • [2] [] Bader, P., Iserles, A., Kropielnicka, K. & Singh, P. (2014), ‘Effective approximation for the semiclassical Schrödinger equation’, Found. Comput. Math. 14(4), 689–720.
  • [3]
  • [4] [] Berrut, J.-P. & Trefethen, L. N. (2004), ‘Barycentric lagrange interpolation’, SIAM review 46(3), 501–517.
  • [5]
  • [6] [] Blanes, S., Casas, F. & Thalhammer, M. (2017), ‘High-order commutator-free quasi-Magnus exponential integrators for non-autonomous linear evolution equations’, Comput. Phys. Commun. 220, 243–262.
  • [7]
  • [8] [] Clenshaw, C. W. (1955), ‘A note on the summation of Chebyshev series’, Math. Tables Aids Comput. 9, 118–120.
  • [9]
  • [10] [] Dutt, A., Gu, M. & Rokhlin, V. (1996), ‘Fast algorithms for polynomial interpolation, integration, and differentiation’, SIAM J. Numer. Anal. 33(5), 1689–1711.
  • [11]
  • [12] [] Gautschi, W. (1970), ‘Efficient computation of the complex error function’, SIAM Journal on Numerical Analysis 7(1), 187–198.
  • [13]
  • [14] [] Gautschi, W. (2004), Orthogonal polynomials: computation and approximation, Oxford University Press.
  • [15]
  • [16] [] Golub, G. H. & Welsch, J. H. (1969), ‘Calculation of gauss quadrature rules’, Mathematics of computation 23(106), 221–230.
  • [17]
  • [18] [] Hairer, E., Lubich, C. & Wanner, G. (2006), Geometric numerical integration: structure-preserving algorithms for ordinary differential equations, Vol. 31, Springer Science & Business Media.
  • [19]
  • [20] [] Hall, M. (1950), ‘A basis for free lie rings and higher commutators in free groups’, Proceedings of the American Mathematical Society 1(5), 575–581.
  • [21]
  • [22] [] Hochbruck, M. & Lubich, C. (1997), ‘On krylov subspace approximations to the matrix exponential operator’, SIAM Journal on Numerical Analysis 34(5), 1911–1925.
  • [23]
  • [24] [] Iserles, A. & Webb, M. (2019), ‘Orthogonal systems with a skew-symmetric differentiation matrix’, Found. Comput. Math. 19(6), 1191–1221.
  • [25]
  • [26] [] Iserles, A. & Webb, M. (2020a), ‘A differential analogue of Favard’s theorem’, arXiv preprint arXiv:2012.07400.
  • [27]
  • [28] [] Iserles, A. & Webb, M. (2020b), ‘A family of orthogonal rational functions and other orthogonal systems with a skew-Hermitian differentiation matrix’, J. Fourier Anal. Appl. 26(1), Paper No. 19.
  • [29]
  • [30] [] Iserles, A. & Webb, M. (2021), ‘Fast computation of orthogonal systems with a skew-symmetric differentiation matrix’, Communications on Pure and Applied Mathematics 74(3), 478–506.
  • [31]
  • [32] [] Iserles, A., Kropielnicka, K. & Singh, P. (2018), ‘Magnus-Lanczos methods with simplified commutators for the Schrödinger equation with a time-dependent potential’, SIAM J. Numer. Anal. 56(3), 1547–1569.
  • [33]
  • [34] [] Iserles, A., Kropielnicka, K. & Singh, P. (2019), ‘Solving Schrödinger equation in semiclassical regime with highly oscillatory time-dependent potentials’, J. Comput. Phys. 376, 564–584.
  • [35]
  • [36] [] Iserles, A., Luong, K. & Webb, M. (2021), ‘Approximation of wave packets on the real line’, arXiv preprint arXiv:2101.02566.
  • [37]
  • [38] [] Ismail, M. E. H., ed. (2020), Univariate Orthogonal Polynomials, Encyclopedia of Special Functions: The Askey–Bateman Project, Cambridge University Press, Cambridge.
  • [39]
  • [40] [] Jin, S., Markowich, P. & Sparber, C. (2011), ‘Mathematical and computational methods for semiclassical Schrödinger equations’, Acta Numer. 20, 121–209.
  • [41]
  • [42] [] Lasser, C. & Lubich, C. (2020), ‘Computing quantum dynamics in the semiclassical regime’, Acta Numerica 29, 229–401.
  • [43]
  • [44] [] Olver, F. W. J., Lozier, D. W., Boisvert, R. F. & Clark, C. W., eds (2010), NIST Handbook of Mathematical Functions, U.S. Department of Commerce, National Institute of Standards and Technology, Washington, DC; Cambridge University Press, Cambridge. With 1 CD-ROM (Windows, Macintosh and UNIX).
  • [45]
  • [46] [] Poppe, G. P. & Wijers, C. M. (1990), ‘More efficient computation of the complex error function’, ACM Transactions on Mathematical Software (TOMS) 16(1), 38–46.
  • [47]
  • [48] [] Reutenauer, C. (1993), Free Lie Algebras, London Maths Soc. Monographs 7, Oxford University Press, Oxford.
  • [49]
  • [50] [] Singh, P. (2016), ‘High accuracy computational methods for the semiclassical Schrödinger equation’.
  • [51]
  • [52] [] Townsend, A., Webb, M. & Olver, S. (2018), ‘Fast polynomial transforms based on toeplitz and hankel matrices’, Mathematics of Computation 87(312), 1913–1934.
  • [53]
  • [54] [] Wang, H., Huybrechs, D. & Vandewalle, S. (2014), ‘Explicit barycentric weights for polynomial interpolation in the roots or extrema of classical orthogonal polynomials’, Mathematics of Computation 83(290), 2893–2914.
  • [55]
  • [56] [] Weideman, J. A. C. (1994), ‘Computation of the complex error function’, SIAM Journal on Numerical Analysis 31(5), 1497–1518.
  • [57]