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

Observability for Schrödinger equations with quadratic Hamiltonians

Alden Waters University of Groningen, Bernoulli Institute, Nijenborgh 9, 9747 AG Groningen, The Netherlands [email protected]
Abstract.

We consider time dependent harmonic oscillators and construct a parametrix to the corresponding Schrödinger equation using Gaussian wavepackets. This parametrix of Gaussian wavepackets is precise and tractable. Using this parametrix we prove L2L^{2} and L2LL^{2}\textendash L^{\infty} observability estimates on unbounded domains ω\omega for a restricted class of initial data. This data includes a class of compactly supported piecewise C1C^{1} functions which have been extended from characteristic functions. Initial data of this form which has the bulk of its mass away from ωc=Ω\omega^{c}=\Omega, a connected bounded domain, is observable, but data centered over Ω\Omega must be very nearly a single Gaussian to be observable. We also give counterexamples to established principles for the simple harmonic oscillator in the case of certain time dependent harmonic oscillators.

Keywords: control theory, Schrödinger equations, observability

MSC classes: 35R01, 35R30, 35L20, 58J45, 35A22

1. Introduction

We start by recalling some results for bounded domains. Let DD, D0D_{0} be bounded smooth subdomains of d\mathbb{R}^{d} with D0DD_{0}\subset D. We consider the controlled linear Schrödinger equation

itu(t,x)=Δu(t,x)+V(t,x)u(t,x)+f(t,x)𝟏D0(x)in(0,T)×D\displaystyle i\partial_{t}u(t,x)=-\Delta u(t,x)+V(t,x)u(t,x)+f(t,x)\mathbf{1}_{D_{0}}(x)\quad\mathrm{in}\quad(0,T)\times D (1.1)
u(t,x)=0on(0,T)×D\displaystyle u(t,x)=0\quad\mathrm{on}\quad(0,T)\times\partial D
u0(x)=u(0,x)inD\displaystyle u_{0}(x)=u(0,x)\quad\mathrm{in}\quad D

here u(t,x)u(t,x) is the state, f(t,x)f(t,x) is the control, and V(t,x)V(t,x) is the potential. All can be complex functions. It is known that the system is well-posed in L2(D)L^{2}(D) with controls in L2((0,T)×D0)L^{2}((0,T)\times D_{0}) for L((0,T)×D)L^{\infty}((0,T)\times D) potentials.

Starting with [29] it was shown that provided the Hamiltonian flow and the observation set satisfies the Geometric Control Condition, then the solution to the Schrödinger equation (1.1) with an LL^{\infty} potential is controllable to another state uT=u(T,x)u_{T}=u(T,x). Later using Carleman estimates, [26, 1] and [27, 28] reduced the assumptions on the domain for LL^{\infty} time independent and time dependent potentials, respectively, c.f. [44] for a survey of such results. If a solution is observable, then often by duality the solution to the adjoint equation is controllable [30]. Interior observability typically amounts to a solution uu satisfying an estimate of the form

u0L2(D)CTuL2((0,T)×D0)\displaystyle\|u_{0}\|_{L^{2}(D)}\leq C_{T}\|u\|_{L^{2}((0,T)\times D_{0})}

for some nonzero CTC_{T}, 0<T<0<T<\infty (f=0f=0). Most of the references mentioned here establish observability estimates.

While much is known about the observability problem for Schrödinger equations on a bounded domain with LL^{\infty} potentials, much less is known for general operators in the free space, that is when (1.1) has DD replaced with d\mathbb{R}^{d} and D0D_{0}, the support of the control, is replaced with ω=dΩ¯\omega=\mathbb{R}^{d}\setminus\overline{\Omega} where Ω\Omega is a bounded domain. Correspondingly, one wants to establish observability for

itu=(κ1(t)Δ+κ2(t)|x|2)u0<tT,xd\displaystyle i\partial_{t}u=\left(-\kappa_{1}(t)\Delta+\kappa_{2}(t)|x|^{2}\right)u\quad 0<t\leq T,\quad x\in\mathbb{R}^{d} (1.2)
u(0,x)=u0(x).\displaystyle u(0,x)=u_{0}(x).

where κ1C1([0,T]),κ2C([0,T])\kappa_{1}\in C^{1}([0,T]),\kappa_{2}\in C([0,T]), and u(0,x)u(0,x) is real valued and in a weighted L2(d)L^{2}(\mathbb{R}^{d}) space. This amounts to establishing a bound for the solution of (1.2) as

u0L2(d)CTuL2((0,T)×ω)\displaystyle\|u_{0}\|_{L^{2}(\mathbb{R}^{d})}\leq C_{T}\|u\|_{L^{2}((0,T)\times\omega)}

which is challenging as ω\omega is no longer bounded. While it seems that this would be an easier problem because the background space is d\mathbb{R}^{d}, the potential itself is also not compactly supported. The Schrödinger operator in the free space with a non-compactly supported potential behaves much differently. The only results for control and observability for this problem in arbitrary dimensions are in [35] for the case of the simple harmonic oscillator κ1(t)=κ2(t)=1/2\kappa_{1}(t)=\kappa_{2}(t)=1/2. Our goal is then to establish observability for a more general class of time-dependent harmonic oscillators. We end up establishing an approximate observability theorem for a limited class initial data u0u_{0}, and an observability estimate for some examples of ’Gaussian-like’ initial data. Approximate observability here means that u(t,x)u(t,x) satisfies

u0L2(d)ηCT(uL2((0,T)×ω)+Tη)\displaystyle\|u_{0}\|_{L^{2}(\mathbb{R}^{d})}-\eta\leq C_{T}(\|u\|_{L^{2}((0,T)\times\omega)}+T\eta) (1.3)

where η>0\eta>0 can be made very small. This error term has to do with the fact that translated real Gaussians are dense in L1(d)L^{1}(\mathbb{R}^{d}) and hence can approximate L2L(d)L^{2}\cap L^{\infty}(\mathbb{R}^{d}) functions, but are not a basis for either space. Whenever η\eta depending on the L(d)L^{\infty}(\mathbb{R}^{d}) norm of u0u_{0} is small enough that the η\eta ’errors’ in (1.3) are absorbed into the L2(d)L^{2}(\mathbb{R}^{d}) norm of u0u_{0} we call this an L2L(d)L^{2}\textendash L^{\infty}(\mathbb{R}^{d}) observability estimate.

The underlying idea here is the use of sophisticated Gaussian wave packets. We prove an explicit approximation theorem which decomposes fL2(d,ex2)f\in L^{2}(\mathbb{R}^{d},e^{x^{2}}) into a finite sum of Gaussians up to some quantifiable error, which is based on [5]. Thus there are two main goals here, one is to construct a nearly explicit parametrix using this decomposition and the other is to prove some localization properties of solutions to the Schrödinger equation in the form of observability estimates. While a least squares approximation may require fewer terms with a higher degree of accuracy, we give an explicit characterization of the decomposition needed to create a completely explicit parametrix with a high degree of accuracy. This parametrix in Theorem 2 was largely inspired by the Fourier integral operators developed in [38]. The derivation of the theorem constructing the parametrix is based on properties of Hermite functions and polynomials.

To see why this is important, we remark that any quantum mechanical system with a potential energy V(x)V(x) has local equilibrium points which can be analysed by the model for a quadratic Hamiltonian, (1.2) with κ1(t)=κ2(t)=1/2\kappa_{1}(t)=\kappa_{2}(t)=1/2. In other words, Taylor expanding V(x)V(x) around the point x0x_{0} gives:

V(x)V(x0)+V(x0)(xx0)+122V(x0)(xx0)2.\displaystyle V(x)\approx V(x_{0})+\nabla V(x_{0})(x-x_{0})+\frac{1}{2}\nabla^{2}V(x_{0})(x-x_{0})^{2}.

If x0x_{0} is a critical point, then the second term vanishes. Translating x0x_{0} to zero we obtain:

V(x)V(0)+122V(0)x2,\displaystyle V(x)\approx V(0)+\frac{1}{2}\nabla^{2}V(0)x^{2},

and we see that the model is reduced to the one of the harmonic oscillator. The Hamiltonian associated to this operator is 12(|p|2+|x|2)\frac{1}{2}(|p|^{2}+|x|^{2}). The Hamiltonian ray path for this operator is computed by solving the system of ODE’s

dx(t)dt=p(t)dp(t)dt=x(t).\displaystyle\frac{dx(t)}{dt}=p(t)\quad\frac{dp(t)}{dt}=-x(t).

The key point is that this operator has a cusp at zero, e.g. one point for which the Hamiltonian ray path is such that dx(t)dt=0\frac{dx(t)}{dt}=0 which can affect the validity of a Gaussian wavepacket construction. Making sense of this phenomenon and the regions where a solution to a much more general time-dependent classical problem (1.2) are concentrated is one of the goals of this paper. The examples presented here all represent toy models of time dependent metrics and potentials V(t,x)V(t,x) in the free space background case around spatial equilibrium. Establishing parametrices and observability for such potentials is key to understanding the behavior of free space quantum mechanical waves.

The use of Gaussian wavepackets has been around since the work of [16] and [11]. For the semi-classical Schrödinger equation in the free space

ihtu=(h2Δ+V)uin(0,T)×d\displaystyle ih\partial_{t}u=(-h^{2}\Delta+V)u\quad\mathrm{in}\quad(0,T)\times\mathbb{R}^{d} (1.4)
u(0,x)=u0(x)\displaystyle u(0,x)=u_{0}(x)

where u0C(d)u_{0}\in C^{\infty}(\mathbb{R}^{d}), the leading order term of uu up to a high degree of accuracy in hh has been shown to take the form of a Gaussian Fourier Integral Operator (FIO). This description holds for very general time-dependent Hamiltonians see [8, 9, 10, 31] and also [25], and finite times TT. However for the classical Schrödinger equation the situation is a bit different as we do not have the added advantage of calculating errors in terms of the scale of the semi-classical parameter hh. For control theory results in the semi-classical case, we direct the reader to [32, 33, 34].

We still have the property in the classical case presented here that the FIOs which are constructed for the non-autonomous problem are Gaussian distributions when the Hamiltonian is quadratic c.f. [38] for a full treatment. Hence the reduction to the model in (1.2). A Gaussian FIO applied to a Gaussian function is again, another Gaussian, which motivates our choice of approximation. The FIO construction now allows us to create a parametrix solution to the Schrödinger equation consisting of a finite collection of tractable propagated wavepackets whose properties, while technical, are not impossible to describe. Some same principles from the compact case related to the Geometric Control Condition still carry over. Moreover, because the representation used here is very close to the classical Hermite functions, the analysis of inner products on L2(ω)L^{2}(\omega) spaces has a direct correlation to spectral theory analysis. Using separation of variables it is usually desirable to prove ψn,ψmL2(ω)\langle\psi_{n},\psi_{m}\rangle_{L^{2}(\omega)} is small in order that the ψnL2(ω)2\|\psi_{n}\|^{2}_{L^{2}(\omega)} terms dominate for some ψn\psi_{n} a spectral basis for the underlying space (d\mathbb{R}^{d}, in this case and ω=dΩ¯\omega=\mathbb{R}^{d}\setminus\overline{\Omega} with Ω\Omega a bounded domain). We see the analogue of this idea directly for the dynamic wavepackets in Lemma 6 and the proof of Theorem 4 in the text. Unfortunately this means we have to restrict the class of initial data for the proofs.

What the theorems here say is that initial data which can be approximated by Gaussians (including some piecewise C1C^{1} compactly supported functions which have been extended from characteristic functions) are L2L(d)L^{2}\textendash L^{\infty}(\mathbb{R}^{d}) observable far away from the ’hole’, Ω\Omega and only initial data which is very nearly a Gaussian is observable when the support of the bulk of the ’mass’ (L2(d)L^{2}(\mathbb{R}^{d}) norm) of the initial data sits over the ’hole’. This makes sense intuitively as initial data which is Gaussian propagates as another Gaussian. Far away from the ’hole’ we see most of the mass of the Gaussian if it started off there, and near to the complement of Ω\Omega, the ’hole’, we recover almost nothing if the mass started over Ω\Omega. We give some explicit examples of such initial data and κ1(t)\kappa_{1}(t) and κ2(t)\kappa_{2}(t) in Section 7. The various values of κ1(t)\kappa_{1}(t) and κ2(t)\kappa_{2}(t) can make the behaviour of the position and the spread of the parametrix quite different even though the overall shape is still Gaussian in the spacial variable. It is curious that in the case of the 1d1d harmonic oscillator κ1=κ2=1/2\kappa_{1}=\kappa_{2}=1/2 that observability on these types of domains ω\omega is true for all fL2()f\in L^{2}(\mathbb{R}) if TπT\geq\pi because the solutions are periodic in time, c.f, Theorem 2.2 [3]. Therefore in this case our construction still has a gap to be filled because the parametrix here is only valid until T=π2T=\frac{\pi}{2} for κ1=κ2=1/2\kappa_{1}=\kappa_{2}=1/2. While this parametrix could be extended to TπT\geq\pi the problem is still that the class of data which is approximately observable depends on location of its support. This gap in our proof techniques cannot easily be closed because it would involve analyzing lower bounds on erfc(z)\mathrm{erfc}(z) for complex zz. However there are no known results on observability when κ1(t)\kappa_{1}(t) and κ2(t)\kappa_{2}(t) both depend on time.

For the free Schrödinger equation and classical harmonic oscillator, the basis of Hermite functions and principles of spectral decomposition have already resulted in our Theorem 3 with η=0\eta=0 in [36, 35, 21, 41, 37]. However one does not expect these techniques to extend themselves to parametrices or observability in the case of time dependent operators as these are purely time separable techniques. Indeed, to this end some counterexamples are shown to established principles for the time independent simple harmonic oscillator in Example 2. Therefore the theorems here to try new techniques which may be applicable to a more general class of operators which are not accessible with spectral theory directly.

2. Background and Main Theorems

2.1. Time dependent quadratic operators

We consider a class of time-dependent quadratic operators:

L=κ1(t)Δ+κ2(t)|x|2\displaystyle L=-\kappa_{1}(t)\Delta+\kappa_{2}(t)|x|^{2}

where 0κ2(t)0\leq\kappa_{2}(t) depends continuously on time tt for 0tT0\leq t\leq T, and for simplicity we also assume κ1(t)\kappa_{1}(t) is bounded below by a positive constant and κ1C1([0,T])\kappa_{1}\in C^{1}([0,T]). We are interested in observability for the non-autonomous Cauchy problem

tu(t,x)+iLu(t,x)=00<tT,xd\displaystyle\partial_{t}u(t,x)+iLu(t,x)=0\quad 0<t\leq T,\quad x\in\mathbb{R}^{d} (2.1)
u0=u(0,x).\displaystyle u_{0}=u(0,x).

Because of the non L2(d)L^{2}(\mathbb{R}^{d}) term |x|2|x|^{2} in the operator, we need a different definition of well-posedness. To this end we set

B={uL2(d):xαDxβuL2(d),α,β0d,|α+β|2}.\displaystyle B=\{u\in L^{2}(\mathbb{R}^{d}):x^{\alpha}D_{x}^{\beta}u\in L^{2}(\mathbb{R}^{d}),\,\,\alpha,\beta\in\mathbb{N}^{d}_{0},\,\,|\alpha+\beta|\leq 2\}.

The space BB is a Hilbert space equipped with the norm

uB2=α,β0d|α+β|=2xαDxβuL2(d)2.\displaystyle\|u\|_{B}^{2}=\sum\limits_{\begin{subarray}{c}\alpha,\beta\in\mathbb{N}_{0}^{d}\\ |\alpha+\beta|=2\end{subarray}}\|x^{\alpha}D_{x}^{\beta}u\|_{L^{2}(\mathbb{R}^{d})}^{2}.

Because there is no easy condition that guarantees the existence of classical solutions for non-autonomous Cauchy problems, it was shown in [38], that the equation (2.1) has a B-valued solution and that this solution exists and is unique. We recall the definition of a B-valued solution.

Definition 1 (B-valued solutions [38]).

A continous function uC0([0,T],B)u\in C^{0}([0,T],B) is a B-valued solution of the non-autonomous Cauchy problem (2.1) if uC1([0,T],L2(d))u\in C^{1}([0,T],L^{2}(\mathbb{R}^{d})), then equation (2.1) is satisfied in L2(d)L^{2}(\mathbb{R}^{d}).

A portion of Theorem 1.2, by Pravda-Starov in [38] gives that for every u0Bu_{0}\in B, the non-autonomous Cauchy problem (2.1) and its adjoint have a unique B-valued solution, and this solution is unitary for all t[0,T]t\in[0,T]. Theorem 1.2 of [38] is actually more general and includes complex-valued quadratic operators iLiL with a non-positive real part for their Weyl symbol. For time independent quadratic Hamiltonians, the theorem is due to Hörmander [20]. We could also analyze complex-valued quadratic operators with the same methodology presented here, but the computations for completely general operators become very difficult quickly, unless a specific example is specified.

2.2. Statement of the Main Theorems

We recall that the classical Wiener’s Tauberian theorem in [43] says that the span of translates of functions are dense in L1()L^{1}(\mathbb{R}) if they have Fourier transform which is non vanishing everywhere. This theorem implies that for fL1()f\in L^{1}(\mathbb{R}) and η(0,1)\eta\in(0,1) that there exists a finite number NN and real numbers αn\alpha_{n} and βn\beta_{n} such that

|f(x)|n|Nβne(xαn)2|𝑑x<η.\displaystyle\int\limits_{\mathbb{R}}\left|f(x)-\sum\limits_{|n|\leq N}\beta_{n}e^{-(x-\alpha_{n})^{2}}\right|\,dx<\eta. (2.2)

It does not tell what the coefficients βn\beta_{n} are nor what the αn\alpha_{n} might be. This statement can be generalized to higher dimensions. Furthermore since real Gaussians cannot be made into an orthonormal basis for L2(d)L^{2}(\mathbb{R}^{d}), there is also no least squares method to generate the βn\beta_{n} and αn\alpha_{n}. In order to build an explicit parametrix for L2(d)L^{2}(\mathbb{R}^{d}) functions, we start by approximating them in a different way in Theorem 1 below. Theorem 1 can be thought of as more explicit than the implication of Wiener’s Tauberian theorem applied to estimate L2(d)L(d)L^{2}(\mathbb{R}^{d})\cap L^{\infty}(\mathbb{R}^{d}) functions since it specifies the coefficients and errors needed for approximation of L2(d,ex2)L^{2}(\mathbb{R}^{d},e^{x^{2}}) functions.

Let hk(x)h_{k}(x) be the kthk^{th} Hermite function. We prove the following Theorem expanding on the work of [5].

Theorem 1.

Let ff be in L2(d,ex2)L^{2}(\mathbb{R}^{d},e^{x^{2}}), and NN a fixed natural number with N>2N>2 and ϵ0(0,1)\epsilon_{0}\in(0,1). Let n=(n1,,nd)n=(n_{1},...,n_{d}) and k=(k1,,kd)k=(k_{1},...,k_{d}) be such that ni,ki{0,,N}n_{i},k_{i}\in\{0,...,N\} for all ii and define

cn=\displaystyle c_{n}= (2.3)
1n1!πk1=n1N1(1)n1k1(k1n1)!(2ϵ0)k1...1nd!πkd=ndNd(1)ndkd(kdnd)!(2ϵ0)kddf(x)ex2kxkex2dx\displaystyle\frac{1}{n_{1}!\sqrt{\pi}}\sum\limits_{k_{1}=n_{1}}^{N_{1}}\frac{(-1)^{n_{1}-k_{1}}}{(k_{1}-n_{1})!(2\epsilon_{0})^{k_{1}}}\,.\,.\,.\frac{1}{n_{d}!\sqrt{\pi}}\sum\limits_{k_{d}=n_{d}}^{N_{d}}\frac{(-1)^{n_{d}-k_{d}}}{(k_{d}-n_{d})!(2\epsilon_{0})^{k_{d}}}\int\limits_{\mathbb{R}^{d}}f(x)e^{x^{2}}\frac{\partial^{k}}{\partial x^{k}}e^{-x^{2}}\,dx

for Ni{0,,N}N_{i}\in\{0,...,N\} such that |(N1,,Nd)|=N|(N_{1},...,N_{d})|=N. We then have that

(d|f(x)|n|Ncne|x+nϵ0|2|2𝑑x)12(eNNϵ0)dfL2(d)+EN.\displaystyle\left(\int\limits_{\mathbb{R}^{d}}\left|f(x)-\sum\limits_{|n|\leq N}c_{n}e^{-|x+n\epsilon_{0}|^{2}}\right|^{2}\,dx\right)^{\frac{1}{2}}\leq(e^{N}N\epsilon_{0})^{d}\|f\|_{L^{2}(\mathbb{R}^{d})}+E_{N}.

The term ENE_{N} is given by

EN=|n|=N+1|dn|2dn=df(x)ex22hn(x)𝑑x.\displaystyle E_{N}=\sum\limits_{|n|=N+1}^{\infty}|d_{n}|^{2}\quad\quad d_{n}=\int\limits_{\mathbb{R}^{d}}f(x)e^{\frac{x^{2}}{2}}h_{n}(x)\,dx.

In some cases EN=0E_{N}=0 such as f=|k|Nαkhkex22f=\sum\limits_{|k|\leq N}\alpha_{k}h_{k}e^{-\frac{x^{2}}{2}} with αk\alpha_{k}\in\mathbb{R}. Bounds on ENE_{N} for H3([M,M]d)H^{3}([-M,M]^{d}) functions can be found in Proposition 3.

We use this wavepacket approximation to generate accurate solutions to the Schrödinger equation. We must make some assumptions on the time span of the solutions in order for the solution not to have any singularities. As such we make the following definition.

Definition 2.

Let H=κ1(t)|p|2+κ2(t)|x|2H=\kappa_{1}(t)|p|^{2}+\kappa_{2}(t)|x|^{2} be the Hamiltonian associated to (2.1) where κ1(t),κ2(t)\kappa_{1}(t),\kappa_{2}(t) are continuous functions. The solution (x(t),p(t))(x(t),p(t)) to

dxi(t)dt=2κ1(t)pi(t)dpi(t)dt=2κ2(t)xi(t)i=1,,d0tT\displaystyle\frac{dx_{i}(t)}{dt}=2\kappa_{1}(t)p_{i}(t)\quad\frac{dp_{i}(t)}{dt}=-2\kappa_{2}(t)x_{i}(t)\quad i=1,...,d\quad 0\leq t\leq T

with (x(0),p(0))(x(0),p(0)) such that x(0)=(1,,1)x(0)=(1,...,1), p(0)=(0,,0)p(0)=(0,...,0) are the Hamiltonian trajectories. We say it is non-zero if for all times tt, 0tT0\leq t\leq T, if xi(t)0x_{i}(t)\neq 0 for all i=1,,di=1,...,d and the quantity

a(t)=e0tpi(s)κ1(s)xi(s)𝑑s\displaystyle a(t)=e^{-\int\limits_{0}^{t}\frac{p_{i}(s)\kappa_{1}(s)}{x_{i}(s)}\,ds}

is bounded. This quantity is the amplitude of a one dimensional wavepacket.

Proposition 1.

The potential pairs κ1(t)=κ2(t)=1/2\kappa_{1}(t)=\kappa_{2}(t)=1/2 (harmonic oscillator) and κ1(t)=e2at2,κ2(t)=e2at2\kappa_{1}(t)=\frac{e^{2at}}{2},\kappa_{2}(t)=\frac{e^{-2at}}{2}, a>1/2a>1/2 (Cadirola-Kanai oscillator) have non-zero Hamiltonian flows for T=π2,T=\frac{\pi}{2},\infty, respectively. Thus they satisfy the conditions of Theorem 3 with these particular TDT_{D}. For the Hamiltonian with σ\sigma a constant

H(t)=|p|22(t+d)a+σ2(t+d)b|x|22,a,b,d>0\displaystyle H(t)=\frac{|p|^{2}}{2(t+d)^{a}}+\frac{\sigma^{2}(t+d)^{b}|x|^{2}}{2},\quad a,b,d>0

it is possible to find a nonzero TDT_{D} depending on d,b,a since the solutions to the Hamiltonian ODEs are given in terms of Bessel functions.

Proposition 1 is shown in Section 7. We build an explicit parametrix for 0tT0\leq t\leq T with TT in the definition above for generic time dependent Schrödinger equations for initial data u0=f(x)e|x|22u_{0}=f(x)e^{-\frac{|x|^{2}}{2}} with fL2(d)f\in L^{2}(\mathbb{R}^{d}) in Theorem 2 stated below whose proof is given in Section 5.

Theorem 2.

Let u0=f(x)e|x|2/2u_{0}=f(x)e^{-|x|^{2}/2}, with ff in L2(d)L^{2}(\mathbb{R}^{d}). Then for some η(0,1)\eta\in(0,1) there is an NN sufficiently large such that

u0|n|Ndnhn(x)e|x|22L2(d)η\displaystyle||u_{0}-\sum\limits_{|n|\leq N}d_{n}h_{n}(x)e^{\frac{-|x|^{2}}{2}}||_{L^{2}(\mathbb{R}^{d})}\leq\eta (2.4)

where dn=f,hnL2(d)d_{n}=\langle f,h_{n}\rangle_{L^{2}(\mathbb{R}^{d})}. We let ϵ0(0,1)\epsilon_{0}\in(0,1) be small and an=nϵ0a_{n}=n\epsilon_{0} with n=(n1,,nd)n=(n_{1},...,n_{d}), k=(k1,,kd)k=(k_{1},...,k_{d}) such that ni,ki{0,,N}n_{i},k_{i}\in\{0,...,N\} for all ii and define

cn=\displaystyle c_{n}= (2.5)
1n1!πk1=n1N1(1)n1(k1n1)!(2ϵ0)k1...1nd!πkd=ndNd(1)nd(kdnd)!(2ϵ0)kddk\displaystyle\frac{1}{n_{1}!\sqrt{\pi}}\sum\limits_{k_{1}=n_{1}}^{N_{1}}\frac{(-1)^{n_{1}}}{(k_{1}-n_{1})!(2\epsilon_{0})^{k_{1}}}\,.\,.\,.\frac{1}{n_{d}!\sqrt{\pi}}\sum\limits_{k_{d}=n_{d}}^{N_{d}}\frac{(-1)^{n_{d}}}{(k_{d}-n_{d})!(2\epsilon_{0})^{k_{d}}}d_{k}

for Ni{0,,N}N_{i}\in\{0,...,N\} such that |(N1,,Nd)|=N|(N_{1},...,N_{d})|=N. Let uu be the solution to (2.1) for all times tt with 0tT0\leq t\leq T and TT in Definition 2 with this type of initial data, u0u_{0}, then

(0Td|u(t,x)|n|Ncnϕn(t,x)|2dxdt)12(η+(eNNϵ0)d))T\displaystyle\left(\int\limits_{0}^{T}\int\limits_{\mathbb{R}^{d}}\left|u(t,x)-\sum\limits_{|n|\leq N}c_{n}\phi_{n}(t,x)\right|^{2}\,dx\,dt\right)^{\frac{1}{2}}\leq\left(\eta+(e^{N}N\epsilon_{0})^{d})\right)T (2.6)

where

ϕn(t,x)=(a2(t)(14iy3(t)))d2eiy1(t)|x|2|y2(t)x+an|2(14iy3(t)).\displaystyle\phi_{n}(t,x)=\left(\frac{a^{2}(t)}{(1-4iy_{3}(t))}\right)^{\frac{d}{2}}e^{iy_{1}(t)|x|^{2}-\frac{|y_{2}(t)x+a_{n}|^{2}}{(1-4iy_{3}(t))}}.

The functions y1(t),y2(t),y3(t)y_{1}(t),y_{2}(t),y_{3}(t) are solutions to the Riccati equations (recalled in (3.2) Section 5) associated with the Hamiltonian in Definition 2.

This parametrix is very explicit as all the functions involved in the representation have an exact form. Indeed, the y1,y2,y3y_{1},y_{2},y_{3} can be computed in terms of rational functions for the Cadirola Kanai oscillator (Example 5 in Section 7 here) and for the harmonic oscillator

(tan(t)2,1cos(t),tan(t)2)\left(-\frac{\tan(t)}{2},\frac{1}{\cos(t)},-\frac{\tan(t)}{2}\right)

(c.f. Exercise 11.1, p.129 [15]) and standard Schrödinger equation

(y1(t)=0,y2(t)=1,y3(t)=t).(y_{1}(t)=0,y_{2}(t)=1,y_{3}(t)=-t).

We now make the following definition about the class of 𝒜\mathcal{A} which are candidates for being approximately observable.

Definition 3.

We let the class of admissible functions 𝒜\mathcal{A} be defined as follows. Let fL2(d)f\in L^{2}(\mathbb{R}^{d}) be such that there exists a natural number NN with N>2N>2, η(0,1)\eta\in(0,1) a~nd\tilde{a}_{n}\in\mathbb{R}^{d}, and c~n+\tilde{c}_{n}\in\mathbb{R}_{+} so we can approximate ff as

(d|f(x)|n|Nc~ne|x+a~n|2|2𝑑x)12η\displaystyle\left(\int\limits_{\mathbb{R}^{d}}\left|f(x)-\sum\limits_{|n|\leq N}\tilde{c}_{n}e^{-|x+\tilde{a}_{n}|^{2}}\right|^{2}\,dx\right)^{\frac{1}{2}}\leq\eta (2.7)

where n=(n1,,nd)n=(n_{1},...,n_{d}) ni{0,,N}n_{i}\in\{0,...,N\}, |n|N|n|\leq N. We also assume that ff has bounded BB norm.

The a~n\tilde{a}_{n} and c~n\tilde{c}_{n} in the above definition may or may not coincide with the ana_{n} and cnc_{n} given in Theorem 1 if we further assume fL2(d,ex2dx)f\in L^{2}(\mathbb{R}^{d},e^{x^{2}}\,dx). We give an elementary example (Lemma 3) in Section 4 that shows this class 𝒜\mathcal{A} includes some compactly supported piecewise C1C^{1} functions which have been extended from characteristic functions. We prove the following two approximate observability theorems with u0𝒜u_{0}\in\mathcal{A}. We can think of each theorem having conditions on the spread of the initial data,

maxn,m|m|,|n|N|a~na~m|,\max\limits_{\begin{subarray}{c}n,m\\ |m|,|n|\leq N\end{subarray}}|\tilde{a}_{n}-\tilde{a}_{m}|,

and placement of the support, but not on the coefficients c~n\tilde{c}_{n} in the definition of 𝒜\mathcal{A}. We say a bounded domain Ωd\Omega\subset\mathbb{R}^{d} is centered at the origin, if the ball containing Ω\Omega has center at the origin.

Theorem 3.

Let Ω\Omega be a bounded domain centered at the origin and assume ω=dΩ¯\omega=\mathbb{R}^{d}\setminus\overline{\Omega}. Assume the Hamiltonian associated to (2.1) is non-zero on the interval [0,TD][0,T_{D}]. Then, there exists a corresponding nonzero constant CTC_{T} for all T(0,TD)T\in(0,T_{D}) with T<T<\infty depending on Ω\Omega, the L([0,T])L^{\infty}([0,T]) norm of κ2(t)\kappa_{2}(t), and the C1([0,T])C^{1}([0,T]) norm of κ1(t)\kappa_{1}(t) such that solutions to the non-autonomous evolution equation (2.1) with initial data u0𝒜u_{0}\in\mathcal{A} satisfy the following inequality

u0L2(d)ηCT(uL2((0,T)×ω)+Tη)\displaystyle\|u_{0}\|_{L^{2}(\mathbb{R}^{d})}-\eta\leq C_{T}\left(\|u\|_{L^{2}((0,T)\times\omega)}+T\eta\right) (2.8)

provided a~n=εn\tilde{a}_{n}=\varepsilon n with ε\varepsilon sufficiently small. In particular, ε\varepsilon, for fixed NN satisfies inequality (6.10) essentially giving εeC1NC2N\varepsilon\leq\frac{e^{-C_{1}N}}{C_{2}N} for C1C_{1}, C2C_{2} constants depending on time and Ω\Omega when 2N<diam(Ω)2N<\mathrm{diam}(\Omega).

Note that we explicitly describe CTC_{T} in the text for all values of diam(Ω)\mathrm{diam}(\Omega) and give examples which show this condition (6.10) gives a nonempty set of ε\varepsilon for each NN and TT and Ω\Omega. We call this an approximate observability inequality since spectral representations such in [3] and [2] in the literature use exact representations of u0L2(d)u_{0}\in L^{2}(\mathbb{R}^{d}) to establish their observability inequalities. Here our η\eta, even though it is small, can depend on another norm of the initial data such as the LL^{\infty} norm, see Lemma 3 in Section 4.

Example 1.

Take u0=|k|Nbke|x+kε|2u_{0}=\sum\limits_{|k|\leq N}b_{k}e^{-|x+k\varepsilon|^{2}} where bk0b_{k}\geq 0 are (any) real positive numbers. Consider Ω\Omega to be the ball of radius R>1R>1. We have that η=0\eta=0, and since u0u_{0} is in 𝒜\mathcal{A}

u0L2(d)CTuL2((0,T)×ω)\displaystyle\|u_{0}\|_{L^{2}(\mathbb{R}^{d})}\leq C_{T}\|u\|_{L^{2}((0,T)\times\omega)}

holds provided (6.10) holds. This inequality (6.10) gives εeC1NRC2N\varepsilon\leq\frac{e^{-C_{1}NR}}{C_{2}N} where C1C_{1} and C2C_{2} are constants depending on time for N<RN<R. These constants are computed in Section 7 for various pairs of κ1(t)\kappa_{1}(t) and κ2(t)\kappa_{2}(t).

While it may seem that this class of functions is very small, and indeed it is, there is no known observability results for these types of oscillators in which both κ1(t)\kappa_{1}(t) and κ2(t)\kappa_{2}(t) depend on time. Furthermore this says that if the bulk of the L2(d)L^{2}(\mathbb{R}^{d}) initial data class sits over the origin, over the center of Ω\Omega– the ’hole” then they must be very nearly perturbations of a Gaussian to be observable. Intuitively, the second theorem says that if we stay far away from Ω\Omega then the initial data does not reach the ’hole’. There is only one similar result in the literature in the d=1d=1 case for κ1=κ2=1/2\kappa_{1}=\kappa_{2}=1/2 which is Theorem 1.5 of [40]. Their theorem states that the support of solution must intersect the observation region in a way which is bounded below by a Gaussian to be L2(d)L^{2}(\mathbb{R}^{d}) observable.

Theorem 4.

Let Ω\Omega be a bounded domain centered at the origin and assume ω=dΩ¯\omega=\mathbb{R}^{d}\setminus\overline{\Omega}. Assume the Hamiltonian associated to (2.1) is non-zero on the interval [0,TD][0,T_{D}]. Then, there exists a corresponding nonzero constant CTC_{T} for all T(0,TD)T\in(0,T_{D}), with T<T<\infty, depending on Ω\Omega, the L([0,T])L^{\infty}([0,T]) norm of κ2(t)\kappa_{2}(t), and the C1([0,T])C^{1}([0,T]) norm of κ1(t)\kappa_{1}(t) such that solutions to the non-autonomous evolution equation (2.1) with initial data u0𝒜u_{0}\in\mathcal{A} satisfy the following inequality

u0L2(d)ηCT(uL2((0,T)×ω)+Tη)\displaystyle\|u_{0}\|_{L^{2}(\mathbb{R}^{d})}-\eta\leq C_{T}\left(\|u\|_{L^{2}((0,T)\times\omega)}+T\eta\right) (2.9)

provided that for |αN|=maxn,m|m|,|n|N|a~na~m||\alpha_{N}|=\max\limits_{\begin{subarray}{c}n,m\\ |m|,|n|\leq N\end{subarray}}|\tilde{a}_{n}-\tilde{a}_{m}| we have

maxn|a~n|>|αN|2+22log(|y2(t)|(1+16(y3(t))2)1)4(y2(t))2(1+16(y3(t))2)1+|αN|y2(t)+diam(Ω)2\displaystyle\max\limits_{n}|\tilde{a}_{n}|>\sqrt{\frac{|\alpha_{N}|^{2}+2-2\log(\sqrt{|y_{2}(t)|(1+16(y_{3}(t))^{2})^{-1}})}{4(y_{2}(t))^{2}(1+16(y_{3}(t))^{2})^{-1}}}+\frac{|\alpha_{N}|}{y_{2}(t)}+\frac{\mathrm{diam}(\Omega)}{2} (2.10)

The functions y2(t),y3(t)y_{2}(t),y_{3}(t) are solutions to the Riccati equations (recalled in (3.2) Section 5) associated with the Hamiltonian in Definition 2

Functions which are L2L(d)L^{2}\textendash L^{\infty}(\mathbb{R}^{d}) observable are examined in Section 8 in Corollary 2. This theorem shows that Gaussian like data sufficiently far away from the ’hole” Ω\Omega is nearly observable. We also compute examples in Section 7 which show that the right hand side of (2.10) is bounded and therefore this condition holds for a nonempty set of u0𝒜u_{0}\in\mathcal{A}. The idea behind the above theorem also lends itself to counterexamples when Ω\Omega is no longer bounded. We recall the following definition which is a modification from the one in [35].

Definition 4.

[Geometric Filling] Let ω\omega be a measurable subset of d\mathbb{R}^{d}. We say that ω\omega satisfies the geometric filling condition if

lim infR1+lim infRd+|ω[R1,R1]×..×[Rd,Rd]||[R1,R1]××[Rd,Rd]|>0\displaystyle\liminf_{R_{1}\rightarrow+\infty}...\liminf_{R_{d}\rightarrow+\infty}\frac{|\omega\cap[-R_{1},R_{1}]\times..\times[-R_{d},R_{d}]|}{|[-R_{1},R_{1}]\times...\times[-R_{d},R_{d}]|}>0

holds.

Given this definition, we have the following counterexample(s).

Example 2.

If the Geometric Filling condition in Definition 4 fails, then ω\omega is not necessarily observable for all u0Bu_{0}\in B in any dimension d2d\geq 2. However, the condition is not sufficient to guarantee observability of (2.1) for all time dependent operators. In particular it is not enough to guarantee the observability inequality holds when the Hamiltonian flow does not change sign.

The demonstration of the above statement is based on two counterexamples. We now remark that Theorem 3 is present in the literature but only for the free space Schrödinger equation and the simple harmonic oscillator.

Remark 1.

In [35] the authors prove exact control for the time-independent fractional Schrödinger equation

(ti(Δ+|x|2)s)u=0\displaystyle(\partial_{t}-i(-\Delta+|x|^{2})^{s})u=0
u(0,x)=u0(x)\displaystyle u(0,x)=u_{0}(x)

for s1s\geq 1 by giving very precise spectral estimates for s1/2s\geq 1/2 and d1d\geq 1, but only general principles for the time of observability. We are not able to handle the case when s1s\neq 1 here, since our methods are based on FIO solutions which do not exist for fractional harmonic oscillators, but we are able to say something about general time dependent quadratic Schrödinger equations. In contrast, the authors of [40] analyze control for

(ti(Δ+|x|2m)u=0m\displaystyle(\partial_{t}-i(-\Delta+|x|^{2m})u=0\quad m\in\mathbb{N}
u(0,x)=u0(x).\displaystyle u(0,x)=u_{0}(x).

We would expect that in the case m>1m>1 FIO solutions also exist, so an extension to time dependent operators would also be possible.

Many of the lemmata and theorems presented in this text for building the solution are also applicable to the time-dependent heat equation under the Wick rotation titt\mapsto it. Observability for the corresponding time-independent heat equation was considered using a spectral decomposition [2] and for the free space heat equation in [42]. However at a certain critical points in the estimates we use the unitarity of the Schrödinger equation to establish observability, and it is unclear what the corresponding replacement analogues are to this property for the heat equation. In specific, we need unitarity of the Schrödinger equation for finding lower bounds for the parametrix solutions. However the parametrix estimates still hold under the Wick rotation.

3. Construction of the FIO for generic κ(t)\kappa(t)

In this section we construct explicit examples of FIO solutions to (2.1) which are only abstractly constructed in [38]. Our treatment is similar to [15], c.f. also [18].

Lemma 1.

Assume the Hamiltonian flow (Definition 2) is non-zero for all 0tT0\leq t\leq T. The associated FIO solution to (2.1) can be written as

u(t,x)=1(2π)d2deiϕ(t,x,η)(a(t))du^0(η)𝑑η.\displaystyle u(t,x)=\frac{1}{(2\pi)^{\frac{d}{2}}}\int\limits_{\mathbb{R}^{d}}e^{i\phi(t,x,\eta)}(a(t))^{d}\hat{u}_{0}(\eta)\,d\eta.

The equality is understood for non-L1L^{1} u0u_{0} functions as in the sense of distributions as above. The phase is of the form

iϕ(t,x,η)=iy1(t)|x|2+iy2(t)xη+iy3(t)|η|2\displaystyle i\phi(t,x,\eta)=iy_{1}(t)|x|^{2}+iy_{2}(t)x\cdot\eta+iy_{3}(t)|\eta|^{2}

where y1(t),y2(t),y3(t)y_{1}(t),y_{2}(t),y_{3}(t) are functions of tt as determined by the following system of ordinary differential equations

y1(t)=4κ1(t)(y1(t))2κ2(t)y1(0)=0\displaystyle y^{\prime}_{1}(t)=-4\kappa_{1}(t)(y_{1}(t))^{2}-\kappa_{2}(t)\quad y_{1}(0)=0 (3.1)
y2(t)=4κ1(t)y1(t)y2(t)y2(0)=1\displaystyle y_{2}^{\prime}(t)=-4\kappa_{1}(t)y_{1}(t)y_{2}(t)\quad y_{2}(0)=1
y3(t)=κ1(t)y22(t)y3(0)=0\displaystyle y_{3}^{\prime}(t)=-\kappa_{1}(t)y_{2}^{2}(t)\quad y_{3}(0)=0

and the amplitude satisfies

a(t)=e20ty1(s)κ1(s)𝑑s.\displaystyle a(t)=e^{-2\int\limits_{0}^{t}y_{1}(s)\kappa_{1}(s)\,ds}.
Proof.

By Theorem 1.3 in [38], as discussed in the previous section, the phase function ϕ(t,x,η)\phi(t,x,\eta) can be written in the form

y1(t)|x|2+y2(t)xη+y3(t)|η|2\displaystyle y_{1}(t)|x|^{2}+y_{2}(t)x\cdot\eta+y_{3}(t)|\eta|^{2}

where y1(t),y2(t),y3(t)y_{1}(t),y_{2}(t),y_{3}(t) are continuous functions of time for 0tT0\leq t\leq T. The phase function will necessarily solve the eikonal equation

tϕ+κ1(t)|xϕ|2+κ2(t)|x|2=0.\displaystyle\partial_{t}\phi+\kappa_{1}(t)|\nabla_{x}\phi|^{2}+\kappa_{2}(t)|x|^{2}=0.

This leads to the following system of ordinary differential equations

y1(t)=4κ1(t)(y1(t))2+κ2(t)\displaystyle-y^{\prime}_{1}(t)=4\kappa_{1}(t)(y_{1}(t))^{2}+\kappa_{2}(t) (3.2)
y2(t)=4κ1(t)y1(t)y2(t)\displaystyle-y_{2}^{\prime}(t)=4\kappa_{1}(t)y_{1}(t)y_{2}(t)
y3(t)=κ1(t)y22(t).\displaystyle-y_{3}^{\prime}(t)=\kappa_{1}(t)y_{2}^{2}(t).

Using the condition xϕ(0,x)=η\nabla_{x}\phi(0,x)=\eta and assuming y1(t)y_{1}(t) is purely real, we obtain the initial conditions for the system. It remains to solve the the transport equation to find a(t)a(t). We then have that

a(t)=2y1(t)κ1(t)a(t)a(0)=1\displaystyle a^{\prime}(t)=-2y_{1}(t)\kappa_{1}(t)a(t)\quad a(0)=1

and the proof is finished. ∎

Lemma 2.

The solution y1(t)y_{1}(t) to the Riccati equation in (3.1) (first equation) is well-posed when the Hamiltonian flow is non-zero.

Proof.

There are examples of κ1(t),κ2(t)\kappa_{1}(t),\kappa_{2}(t) for which the Riccati equation is not well-posed (either at all, or after a certain interval) and a corresponding FIO solution cannot be constructed. While this is implied by the abstract constructions in [20], [19] and [38], it is important to see explicitly in our section on observability estimates. Let S(t)=4κ1(t)κ2(t)S(t)=4\kappa_{1}(t)\kappa_{2}(t), and R(t)=ddtln(κ1(t))R(t)=\frac{d}{dt}\mathrm{ln}(\kappa_{1}(t)). Then a solution v(t)v(t) to

v′′Rv+Sv=0v(0)=1,v(0)=0\displaystyle v^{\prime\prime}-Rv^{\prime}+Sv=0\quad v(0)=1,\quad v^{\prime}(0)=0

is related to the Riccati solution (first equation in (3.1)) by y1=v4κ1vy_{1}=\frac{v^{\prime}}{4\kappa_{1}v}. Therefore in order for the Riccati equation to be well-posed, κ1(t)\kappa_{1}(t) must also be differentiable and nonzero, and vv must not be zero. (In addition to κ2(t)\kappa_{2}(t) being continuous). Notice that in our case x(t)=vx(t)=v, which is why we assume the Hamiltonian flow is non-zero. ∎

Theorem 1.3 in [38] then gives that the FIO is unitary when the flow is non-zero. This statement also follows by direct calculation in our case. Notice that the second assumption in the definition is equivalent to a(t)2<a(t)^{2}<\infty.

4. Proof of Theorem 1, Approximation with Gaussians

We would like to use the following Hermite inspired expansion for all f(x)f(x) in L2(,ex2dx):L^{2}(\mathbb{R},e^{x^{2}}\,dx):

f(x)=n=0bndndxnex2\displaystyle f(x)=\sum\limits_{n=0}^{\infty}b_{n}\frac{d^{n}}{dx^{n}}e^{-x^{2}}

where

bn=12nn!πf(x)ex2dndxnex2𝑑x.\displaystyle b_{n}=\frac{1}{2^{n}n!\sqrt{\pi}}\int\limits_{\mathbb{R}}f(x)e^{x^{2}}\frac{d^{n}}{dx^{n}}e^{-x^{2}}\,dx.

Indeed, recall that the Hermite polynomials are defined as

Hn(x)=(1)nex2dndxnex2.\displaystyle H_{n}(x)=(-1)^{n}e^{x^{2}}\frac{d^{n}}{dx^{n}}e^{-x^{2}}. (4.1)

Then the set of Hermite functions

{hn(x)=12nn!πHn(x)ex2/2:n}\displaystyle\left\{h_{n}(x)=\frac{1}{\sqrt{2^{n}n!\sqrt{\pi}}}H_{n}(x)e^{-x^{2}/2}:n\in\mathbb{N}\right\} (4.2)

is a well known orthonormal basis of L2()L^{2}(\mathbb{R}). For any gL2()g\in L^{2}(\mathbb{R}) it is possible to write

g(x)=n=0dnhn(x)\displaystyle g(x)=\sum\limits_{n=0}^{\infty}d_{n}h_{n}(x)

where

dn=g(x)hn(x)¯𝑑x.\displaystyle d_{n}=\int\limits_{-\infty}^{\infty}g(x)\overline{h_{n}(x)}\,dx.

Applying the representation with g(x)=f(x)ex2/2g(x)=f(x)e^{x^{2}/2} gives the desired formula for the bnb_{n}, and moreover we have then the following relationship:

bn=dn(1)n2nn!π.\displaystyle b_{n}=d_{n}\frac{(-1)^{n}}{\sqrt{2^{n}n!\sqrt{\pi}}}. (4.3)

Now we want to approximate the derivatives of ex2e^{-x^{2}} in L2()L^{2}(\mathbb{R}). This proposition is loosely based on Proposition 2 in [5], which in which the errors are done in a different norm only for compactly supported functions.

Proposition 2.

For all fL2(d,ex2dx)f\in L^{2}(\mathbb{R}^{d},e^{x^{2}}\,dx), NN a fixed natural number and ϵ0(0,1)\epsilon_{0}\in(0,1) we have that

(|f(x)n=0Ncne(x+nϵ0)2|2𝑑x)12NeNϵ0fL2()+EN\displaystyle\left(\int\limits_{\mathbb{R}}\left|f(x)-\sum\limits_{n=0}^{N}c_{n}e^{-(x+n\epsilon_{0})^{2}}\right|^{2}\,dx\right)^{\frac{1}{2}}\leq Ne^{N}\epsilon_{0}\|f\|_{L^{2}(\mathbb{R})}+E_{N}

where

cn=1n!πk=nN(1)nk(kn)!(2ϵ0)kf(x)ex2dkdxkex2𝑑x\displaystyle c_{n}=\frac{1}{n!\sqrt{\pi}}\sum\limits_{k=n}^{N}\frac{(-1)^{n-k}}{(k-n)!(2\epsilon_{0})^{k}}\int\limits_{\mathbb{R}}f(x)e^{x^{2}}\frac{d^{k}}{dx^{k}}e^{-x^{2}}\,dx (4.4)

and

EN=(n=N+1|dn|2)12dn=(1)nf(x)ex2dndxnex2𝑑x.\displaystyle E_{N}=\left(\sum\limits_{n=N+1}^{\infty}|d_{n}|^{2}\right)^{\frac{1}{2}}\qquad d_{n}=(-1)^{n}\int\limits_{\mathbb{R}}f(x)e^{x^{2}}\frac{d^{n}}{dx^{n}}e^{-x^{2}}\,dx.
Proof of Proposition 2.

We start with setting

En(x)=(dndxnex21ϵ0nk=0n(1)nk(nk)e(x+kϵ0)2).\displaystyle E_{n}(x)=\left(\frac{d^{n}}{dx^{n}}e^{-x^{2}}-\frac{1}{\epsilon_{0}^{n}}\sum\limits_{k=0}^{n}(-1)^{n-k}\binom{n}{k}e^{-(x+k\epsilon_{0})^{2}}\right).

This is in L2()L^{2}(\mathbb{R}) since each of the terms is in L2()L^{2}(\mathbb{R}). By Taylor’s theorem and the method of forward finite differences for gCn+2()g\in C^{n+2}(\mathbb{R}), ϵ0(0,1)\epsilon_{0}\in(0,1) we have

dndxng(x)=k=0n(1)nk((nk)1ϵ0ng(x+kϵ0)ϵ0nk+1(n+1)!(nk)kn+1gn+1(ξk))\displaystyle\frac{d^{n}}{dx^{n}}g(x)=\sum\limits_{k=0}^{n}(-1)^{n-k}\left(\binom{n}{k}\frac{1}{\epsilon_{0}^{n}}g(x+k\epsilon_{0})-\frac{\epsilon_{0}^{n-k+1}}{(n+1)!}\binom{n}{k}k^{n+1}g^{n+1}(\xi_{k})\right)

where ξk\xi_{k} is between xx and x+nϵ0x+n\epsilon_{0}. This is the Lagrange form of the remainder. If we set g(x)=ex2g(x)=e^{-x^{2}} then this formula gives the difference En(x)E_{n}(x). We let

g~n+1(x)=(1)n2nn!πdn+1dxn+1ex2.\tilde{g}^{n+1}(x)=\frac{(-1)^{n}}{\sqrt{2^{n}n!\sqrt{\pi}}}\frac{d^{n+1}}{dx^{n+1}}e^{-x^{2}}.

The point is that

supξk(x,x+nϵ0)|gn+1(ξk)|2=supl(0,nϵ0)|gn+1(x+l)|2=|gn+1(x+l)|2\displaystyle\sup_{\xi_{k}\in(x,x+n\epsilon_{0})}|g^{n+1}(\xi_{k})|^{2}=\sup_{l\in(0,n\epsilon_{0})}|g^{n+1}(x+l)|^{2}=|g^{n+1}(x+l_{*})|^{2}

for some l[0,nϵ0]l_{*}\in[0,n\epsilon_{0}]. Let rr be a positive real number. Then by differentiability of gn+1(x)g^{n+1}(x), we have that there exists a sequence ljl_{j} such that |gn+1(x+lj)|2|gn+1(x+l)|2|g^{n+1}(x+l_{j})|^{2}\rightarrow|g^{n+1}(x+l_{*})|^{2} uniformly on {x:|x|r}\{x:|x|\leq r\}. It follows that

|x|rsupξk|g~n+1(ξk)|2dx=supl[0,ϵ0n]|x|r|g~n+1(x+l)|2𝑑x.\displaystyle\int\limits_{|x|\leq r}\sup\limits_{\xi_{k}}|\tilde{g}^{n+1}(\xi_{k})|^{2}\,dx=\sup\limits_{l\in[0,\epsilon_{0}n]}\int\limits_{|x|\leq r}|\tilde{g}^{n+1}(x+l)|^{2}\,dx. (4.5)

We then notice that

|Hn(x)|2ex2𝑑x=ex2|dndxnex2|2𝑑x=2nn!π\displaystyle\int\limits_{\mathbb{R}}|H_{n}(x)|^{2}e^{-x^{2}}\,dx=\int\limits_{\mathbb{R}}e^{x^{2}}\left|\frac{d^{n}}{dx^{n}}e^{-x^{2}}\right|^{2}\,dx=2^{n}n!\sqrt{\pi}

which implies since ϵ0n>0\epsilon_{0}n>0 and ex21e^{x^{2}}\geq 1 for all xx\in\mathbb{R}

|x|rsupξk|g~n+1(ξk)|2dx\displaystyle\int\limits_{|x|\leq r}\sup\limits_{\xi_{k}}|\tilde{g}^{n+1}(\xi_{k})|^{2}\,dx\leq (4.6)
supl[0,ϵ0n]e(x+l)2|g~n+1(x+l)|2𝑑xex2|g~n+1(x)|2𝑑x2(n+1).\displaystyle\sup\limits_{l\in[0,\epsilon_{0}n]}\int\limits_{\mathbb{R}}e^{(x+l)^{2}}|\tilde{g}^{n+1}(x+l)|^{2}\,dx\leq\int\limits_{\mathbb{R}}e^{x^{2}}|\tilde{g}^{n+1}(x)|^{2}\,dx\leq 2(n+1).

Using Minkowski’s inequality followed by Cauchy-Schwarz we have that

n=0NbnEnL2()=n=0NdnE~nL2()n=0N|dn|E~nL2()\displaystyle\|\sum\limits_{n=0}^{N}b_{n}E_{n}\|_{L^{2}(\mathbb{R})}=\|\sum\limits_{n=0}^{N}d_{n}\tilde{E}_{n}\|_{L^{2}(\mathbb{R})}\leq\sum\limits_{n=0}^{N}|d_{n}|\|\tilde{E}_{n}\|_{L^{2}(\mathbb{R})}\leq (4.7)
(n=0N|dn|2)12(n=0NE~nL2()2)12.\displaystyle\left(\sum\limits_{n=0}^{N}|d_{n}|^{2}\right)^{\frac{1}{2}}\left(\sum\limits_{n=0}^{N}\|\tilde{E}_{n}\|^{2}_{L^{2}(\mathbb{R})}\right)^{\frac{1}{2}}.

We then have

ϵ02E~nL2(|x|r)2|x|r|k=0nkk+1(nk)(n+1)!supξk|g~n+1(ξk)||2𝑑x\displaystyle\epsilon_{0}^{-2}\|\tilde{E}_{n}\|_{L^{2}(|x|\leq r)}^{2}\leq\int\limits_{|x|\leq r}\left|\sum\limits_{k=0}^{n}\frac{k^{k+1}\binom{n}{k}}{(n+1)!}\sup\limits_{\xi_{k}}|\tilde{g}^{n+1}(\xi_{k})|\right|^{2}\,dx
k=0n|kk+1(nk)(n+1)!|2m=0n(|x|rsupξm|g~n+1(ξm)|2dx)\displaystyle\leq\sum\limits_{k=0}^{n}\left|\frac{k^{k+1}\binom{n}{k}}{(n+1)!}\right|^{2}\sum\limits_{m=0}^{n}\left(\int\limits_{|x|\leq r}\sup\limits_{\xi_{m}}|\tilde{g}^{n+1}(\xi_{m})|^{2}\,dx\right) (4.8)
k=0n|kk+1k!(nk)!|2e2nn\displaystyle\leq\sum\limits_{k=0}^{n}\left|\frac{k^{k+1}}{k!(n-k)!}\right|^{2}\leq e^{2n}n

where in the last line we have used estimate (4.6). Combining (4.7) and (4) we have by the monotone convergence theorem

n=0NbnEnL2()ϵ0NeNfL2().\displaystyle\|\sum\limits_{n=0}^{N}b_{n}E_{n}\|_{L^{2}(\mathbb{R})}\leq\epsilon_{0}Ne^{N}\|f\|_{L^{2}(\mathbb{R})}.

It remains to bound

n=N+1bndndxnex2L2().\displaystyle\|\sum\limits_{n=N+1}^{\infty}b_{n}\frac{d^{n}}{dx^{n}}e^{-x^{2}}\|_{L^{2}(\mathbb{R})}.

By orthogonality of the hnsh_{n}^{\prime}s,

n=N+1bndndxnex2L2(,ex2dx)(n=N+1|dn|2)12.\displaystyle\|\sum\limits_{n=N+1}^{\infty}b_{n}\frac{d^{n}}{dx^{n}}e^{-x^{2}}\|_{L^{2}(\mathbb{R},e^{x^{2}}\,dx)}\leq\left(\sum\limits_{n=N+1}^{\infty}|d_{n}|^{2}\right)^{\frac{1}{2}}.

In order to obtain the coefficients cnc_{n} we note that

n=0Nbnϵ0nk=0n(1)nk(nk)e(x+kϵ0)2=k=0N(n=kN(1)nk(nk)bnϵ0n)e(x+kϵ0)2\displaystyle\sum\limits_{n=0}^{N}b_{n}\epsilon_{0}^{-n}\sum\limits_{k=0}^{n}(-1)^{n-k}\binom{n}{k}e^{-(x+k\epsilon_{0})^{2}}=\sum\limits_{k=0}^{N}\left(\sum\limits_{n=k}^{N}(-1)^{n-k}\binom{n}{k}b_{n}\epsilon_{0}^{-n}\right)e^{-(x+k\epsilon_{0})^{2}}

which is Iverson’s summation technique (c.f. equation 2.32 in Section 2.4 of [14] p.36). ∎

Proof of Theorem 1.

The proof follows using the product topology in the previous definition. We define a generalized Hermite function as

hm(x)=i=1dhmi(xi)\displaystyle h_{m}(x)=\prod\limits_{i=1}^{d}h_{m_{i}}(x_{i})

where m=(m1,,md)m=(m_{1},...,m_{d}) is now a multi-index of degree dd and hmi(xi)h_{m_{i}}(x_{i}) is the one dimensional Hermite function. The proof follows using the product topology from the previous Proposition. Indeed if

f(x)e|x|2/2=m=0αmi=1dhmi(xi)\displaystyle f(x)e^{|x|^{2}/2}=\sum\limits_{m=0}^{\infty}\alpha_{m}\prod\limits_{i=1}^{d}h_{m_{i}}(x_{i})

with

αm=df(x)e|x|22i=1dhmi(xi)dx\displaystyle\alpha_{m}=\int\limits_{\mathbb{R}^{d}}f(x)e^{\frac{|x|^{2}}{2}}\prod\limits_{i=1}^{d}h_{m_{i}}(x_{i})\,dx

Then we can define a βm\beta_{m} (d-dimensional version of bmb_{m}) corresponding to αm\alpha_{m} (dmd_{m} as in the 1-dimensional case), as in (4.3)

βm=i=1d((1)mimi!2miπdf(x)e|x|22hmi(xi)𝑑x).\displaystyle\beta_{m}=\prod_{i=1}^{d}\left(\frac{(-1)^{m_{i}}}{\sqrt{m_{i}!2^{m_{i}}\sqrt{\pi}}}\int\limits_{\mathbb{R}^{d}}f(x)e^{\frac{|x|^{2}}{2}}h_{m_{i}}(x_{i})\,dx\right).

We can write almost as before:

dndxnex2=i=1d(ki=0ni(1)niki(niki)1ϵ0ni)e|x+nϵ0|2+𝒪(ϵ0d(e2NdNd)12)\displaystyle\frac{d^{n}}{dx^{n}}e^{x^{2}}=\prod_{i=1}^{d}\left(\sum\limits_{k_{i}=0}^{n_{i}}(-1)^{n_{i}-k_{i}}\binom{n_{i}}{k_{i}}\frac{1}{\epsilon_{0}^{n_{i}}}\right)e^{-|x+n\epsilon_{0}|^{2}}+\mathcal{O}(\epsilon_{0}^{d}\left(e^{2Nd}N^{d}\right)^{\frac{1}{2}})

with the equality in the sense of L2(d)L^{2}(\mathbb{R}^{d}). If f(x)=i=1dfi(xi)f(x)=\prod\limits_{i=1}^{d}f_{i}(x_{i}) then βn=i=1dβni\beta_{n}=\prod\limits_{i=1}^{d}\beta_{n_{i}}. This gives

|n|Nβni=1d(ki=0ni(1)niki(niki)1ϵ0ni)e|x+nϵ0|2=\displaystyle\sum\limits_{|n|\leq N}\beta_{n}\prod\limits_{i=1}^{d}\left(\sum\limits_{k_{i}=0}^{n_{i}}(-1)^{n_{i}-k_{i}}\binom{n_{i}}{k_{i}}\frac{1}{\epsilon_{0}^{n_{i}}}\right)e^{-|x+n\epsilon_{0}|^{2}}=
|n|Ni=1d(βniki=0ni(1)niki(niki)1ϵ0ni)e|x+nϵ0|2=\displaystyle\sum\limits_{|n|\leq N}\prod\limits_{i=1}^{d}\left(\beta_{n_{i}}\sum\limits_{k_{i}=0}^{n_{i}}(-1)^{n_{i}-k_{i}}\binom{n_{i}}{k_{i}}\frac{1}{\epsilon_{0}^{n_{i}}}\right)e^{-|x+n\epsilon_{0}|^{2}}=
i=1dni=0Ni(βniki=0ni(1)niki(niki)1ϵ0ni)e|x+nϵ0|2.\displaystyle\prod\limits_{i=1}^{d}\sum\limits_{n_{i}=0}^{N_{i}}\left(\beta_{n_{i}}\sum\limits_{k_{i}=0}^{n_{i}}(-1)^{n_{i}-k_{i}}\binom{n_{i}}{k_{i}}\frac{1}{\epsilon_{0}^{n_{i}}}\right)e^{-|x+n\epsilon_{0}|^{2}}.

Again using Iverson’s summation technique on each of the coordinates/sums i=1,,di=1,...,d seperately we have that

ni=0Ni(βniki=0ni(1)niki(niki)1ϵ0ni)e|x+niϵ0|2=ni=0Ni(ki=niNi(1)kini(kini)βkiϵ0ki)e|x+niϵ0|2.\displaystyle\sum\limits_{n_{i}=0}^{N_{i}}\left(\beta_{n_{i}}\sum\limits_{k_{i}=0}^{n_{i}}(-1)^{n_{i}-k_{i}}\binom{n_{i}}{k_{i}}\frac{1}{\epsilon_{0}^{n_{i}}}\right)e^{-|x+n_{i}\epsilon_{0}|^{2}}=\sum\limits_{n_{i}=0}^{N_{i}}\left(\sum\limits_{k_{i}=n_{i}}^{N_{i}}(-1)^{k_{i}-n_{i}}\binom{k_{i}}{n_{i}}\frac{\beta_{k_{i}}}{\epsilon_{0}^{k_{i}}}\right)e^{-|x+n_{i}\epsilon_{0}|^{2}}.

We then obtain

i=1dni=0Ni(ki=niNi(1)kini(kini)βkiϵ0ki)e|x+nϵ0|2=\displaystyle\prod\limits_{i=1}^{d}\sum\limits_{n_{i}=0}^{N_{i}}\left(\sum\limits_{k_{i}=n_{i}}^{N_{i}}(-1)^{k_{i}-n_{i}}\binom{k_{i}}{n_{i}}\frac{\beta_{k_{i}}}{\epsilon_{0}^{k_{i}}}\right)e^{-|x+n\epsilon_{0}|^{2}}=
|n|N(k1=n1N1(1)k1n1(k1n1)βk1ϵ0k1.kd=ndNd(1)kdnd(kdnd)βkdϵ0kd)e|x+nϵ0|2=\displaystyle\sum\limits_{|n|\leq N}\left(\sum\limits_{k_{1}=n_{1}}^{N_{1}}(-1)^{k_{1}-n_{1}}\binom{k_{1}}{n_{1}}\frac{\beta_{k_{1}}}{\epsilon_{0}^{k_{1}}}....\sum\limits_{k_{d}=n_{d}}^{N_{d}}(-1)^{k_{d}-n_{d}}\binom{k_{d}}{n_{d}}\frac{\beta_{k_{d}}}{\epsilon_{0}^{k_{d}}}\right)e^{-|x+n\epsilon_{0}|^{2}}=
|n|N((k1=n1N1(1)k1n1(k1n1)1ϵ0k1.kd=ndNd(1)kdnd(kdnd)1ϵ0kd)βk)e|x+nϵ0|2\displaystyle\sum\limits_{|n|\leq N}\left(\left(\sum\limits_{k_{1}=n_{1}}^{N_{1}}(-1)^{k_{1}-n_{1}}\binom{k_{1}}{n_{1}}\frac{1}{\epsilon_{0}^{k_{1}}}....\sum\limits_{k_{d}=n_{d}}^{N_{d}}(-1)^{k_{d}-n_{d}}\binom{k_{d}}{n_{d}}\frac{1}{\epsilon_{0}^{k_{d}}}\right)\beta_{k}\right)e^{-|x+n\epsilon_{0}|^{2}}

which defines the coefficients cnc_{n}, (4.4), in the dd dimensional case. Here we have used the fact that f(x)f(x) can be approximated by products of one dimensional functions and taken the limit. ∎

We have the bound on the errors for compactly supported functions with sufficient regularity. We define Hj([M,M])H^{j}([-M,M]) as the space of Hj(d)H^{j}(\mathbb{R}^{d}) functions having compact support in [M,M]d[-M,M]^{d}.

Proposition 3.

We have for fH3([M,M]d)f\in H^{3}([-M,M]^{d}), N>2N>2

ENp3(M)fH3(d,ex2dx)Nd4\displaystyle E_{N}\leq\frac{p_{3}(M)||f||_{H^{3}(\mathbb{R}^{d},e^{x^{2}}\,dx)}}{N^{\frac{d}{4}}}

with p3(M)p_{3}(M) denoting a polynomial of degree 33 in MdM^{d}.

Proof.

We show the result in dimension 1 with the obvious generalisation to dimension dd. Loosely inspired by [4], if generically f1Hj([M,M])f_{1}\in H^{j}([-M,M]), mm\in\mathbb{N} then the bound on the corresponding dnd_{n} is given to us by

|dn|(x+ddx)jf1L1()(2(n+1))j2.\displaystyle|d_{n}|\leq\frac{\|\left(x+\frac{d}{dx}\right)^{j}f_{1}\|_{L^{1}(\mathbb{R})}}{(2(n+1))^{\frac{j}{2}}}.

This is a consequence of the fact

2(n+1)hn+1(x)=xhn(x)ddxhn(x)\displaystyle\sqrt{2(n+1)}h_{n+1}(x)=xh_{n}(x)-\frac{d}{dx}h_{n}(x)

and integration by parts. This bound implies for f1=fex22f_{1}=fe^{\frac{x^{2}}{2}}, fHj([M,M])f\in H^{j}([-M,M])

(N+1|dn|2)12=EN(x+ddx)j(fex22)L1()(n=N+11(2(n+1))j2)1/2.\displaystyle\left(\sum\limits_{N+1}^{\infty}|d_{n}|^{2}\right)^{\frac{1}{2}}=E_{N}\leq\|\left(x+\frac{d}{dx}\right)^{j}(fe^{\frac{x^{2}}{2}})\|_{L^{1}(\mathbb{R})}\left(\sum\limits_{n=N+1}^{\infty}\frac{1}{(2(n+1))^{\frac{j}{2}}}\right)^{1/2}.

We set j=3j=3 and want to estimate the sum on the right hand side using an integral. We recall the Euler summation formula which is eq. 9.67, Section 9.5 p 455 of [14]. Given an integer valued function h(α)h(\alpha), h(α)h(\alpha) may be estimated by the Euler summation formula

aαbh(α)=abh(x)𝑑x+m=1kBmm!hm1(x)|x=ax=b+Rk\sum\limits_{a\leq\alpha\leq b}h(\alpha)=\int\limits_{a}^{b}h(x)\,dx+\sum\limits_{m=1}^{k}\frac{B_{m}}{m!}h^{m-1}(x)|_{x=a}^{x=b}+R_{k} (4.9)

where BmB_{m} is the mthm^{th} Bernoulli number and hm(x)h^{m}(x) denotes the mthm^{th} derivative of h(x)h(x). The remainder RkR_{k} is defined as

Rk=(1)k+1Bk({x})k!hk(x)𝑑x.R_{k}=(-1)^{k+1}\int\limits_{\mathbb{R}}\frac{B_{k}(\{x\})}{k!}h^{k}(x)\,dx.

The notation {x}\{x\} denotes the fractional part of xx, and Bk({x})B_{k}(\{x\}) denotes the kthk^{th} Bernoulli polynomial. Applying this gives for N>2N>2

(n=N+11(2(n+1))32)1/210N14\displaystyle\left(\sum\limits_{n=N+1}^{\infty}\frac{1}{(2(n+1))^{\frac{3}{2}}}\right)^{1/2}\leq\frac{10}{N^{\frac{1}{4}}}

were we have used the fact that |B2({x})|B2=1/2|B_{2}(\{x\})|\leq B_{2}=1/2 for all xx\in\mathbb{R}. ∎

Corollary 1.

Define for bkb_{k} real numbers indexed by k=(k1,,kd)k=(k_{1},...,k_{d}), ki{1,,N}k_{i}\in\{1,...,N\}

ψ(x)=|k|Nbk(i=1dhki(xi))e|x|22\displaystyle\psi(x)=\sum\limits_{|k|\leq N}b_{k}\left(\prod\limits_{i=1}^{d}h_{k_{i}}(x_{i})\right)e^{-\frac{|x|^{2}}{2}}

Then ψ\psi is such that η\eta in (2.7) has EN=0E_{N}=0.

Proof.

The coefficients dn=0d_{n}=0 for all |n|N+1|n|\geq N+1 in the case d=1d=1 with this choice of ψ\psi. The result then follows immediately since the same statement holds for each coordinate xix_{i}. ∎

Unfortunately because the approximation method is based on the method of forward finite differences the coefficients of the Gaussians cnc_{n} alternate and only very specific sums of the form

|k|Nϵ0kbkhk(x)e|x|22\displaystyle\sum\limits_{|k|\leq N}\epsilon_{0}^{k}b_{k}h_{k}(x)e^{\frac{-|x|^{2}}{2}}

have positive coefficients c~n\tilde{c}_{n} in their approximations and are therefore in the class 𝒜\mathcal{A}. For functions which have been extended from characteristic functions to piecewise C1C^{1} compactly supported functions are better off decomposed into Gaussians in a natural way based on Riemann integration in Lemma 3 below.

Lemma 3 (Example of functions in 𝒜\mathcal{A}).

There exist piecewise C1C^{1} compactly supported functions which agree with a characteristic function χB\chi_{B} on BM/2B_{M/2} a ball of radius M/2M/2, for all M2M\geq 2 which are in 𝒜\mathcal{A} for some η(0,1)\eta\in(0,1).

Proof.

We show this in 1d1d with the obvious generalisation holding. Let MM be a real positive number, and Δx(0,1)\Delta x\in(0,1). Loosely, following [17] in 1d we create a Riemann sum

n=0NΔxnπe(xxn+M)2|xn|2Mxn=0+nΔx\displaystyle\sum\limits_{n=0}^{N}\frac{\Delta x_{n}}{\sqrt{\pi}}e^{-(x-x_{n}+M)^{2}}\quad|x_{n}|\leq 2M\quad x_{n}=0+n\Delta x

whose coefficients Δxn\Delta x_{n} are all positive. Using the fact that erf(x)\mathrm{erf}(x) is monotone decreasing in xx, by the standard error estimates in Riemann integration

supx|nΔxnπe(xxn+M)21π02Me|xy+M|2𝑑y|<2MΔx.\displaystyle\sup\limits_{x\in\mathbb{R}}\left|\sum\limits_{n}\frac{\Delta x_{n}}{\sqrt{\pi}}e^{-(x-x_{n}+M)^{2}}-\frac{1}{\sqrt{\pi}}\int\limits_{0}^{2M}e^{-|x-y+M|^{2}}\,dy\right|<2M\Delta x. (4.10)

We take this further. Let χBM/2\chi_{B_{M/2}} be the characteristic function on [M/2,M/2][-M/2,M/2]. Then changing variables gives

supx[M/2,M/2]|χBM/21πMMe|xy|2𝑑y|2erfc(M/2).\displaystyle\sup\limits_{x\in[-M/2,M/2]}\left|\chi_{B_{M/2}}-\frac{1}{\sqrt{\pi}}\int\limits_{-M}^{M}e^{-|x-y|^{2}}\,dy\right|\leq 2\mathrm{erfc}(M/2).

We extend the characteristic function

ϕ1(x)=\displaystyle\phi_{1}(x)=
χBM/2(x)+χ[M/2,)(x)1πMMe|xy|2𝑑y1πMMe|M/2y|2𝑑y+χ(,M/2](x)1πMMe|xy|2𝑑y1πMMe|M/2+y|2𝑑y.\displaystyle\chi_{B_{M/2}}(x)+\chi_{[M/2,\infty)}(x)\frac{\frac{1}{\sqrt{\pi}}\int\limits_{-M}^{M}e^{-|x-y|^{2}}\,dy}{\frac{1}{\sqrt{\pi}}\int\limits_{-M}^{M}e^{-|M/2-y|^{2}}\,dy}+\chi_{(-\infty,-M/2]}(x)\frac{\frac{1}{\sqrt{\pi}}\int\limits_{-M}^{M}e^{-|x-y|^{2}}\,dy}{\frac{1}{\sqrt{\pi}}\int\limits_{-M}^{M}e^{-|-M/2+y|^{2}}\,dy}.

It follows that

supx|1πMMe|xy|2ϕ1|erfc(M/2)+erfc(3M/2)\displaystyle\sup\limits_{x\in\mathbb{R}}\left|\frac{1}{\sqrt{\pi}}\int\limits_{-M}^{M}e^{-|x-y|^{2}}-\phi_{1}\right|\leq\mathrm{erfc}(M/2)+\mathrm{erfc}(3M/2)

because for x[M/2,)x\in[M/2,\infty)

2|1πMMe|xy|2𝑑y1πMMe|M/2y|2𝑑y1πMMe|xy|2𝑑y|erfc(M/2)+erfc(3M/2)\displaystyle 2\left|\frac{\frac{1}{\sqrt{\pi}}\int\limits_{-M}^{M}e^{-|x-y|^{2}}\,dy}{\frac{1}{\sqrt{\pi}}\int\limits_{-M}^{M}e^{-|M/2-y|^{2}}\,dy}-\frac{1}{\sqrt{\pi}}\int\limits_{-M}^{M}e^{-|x-y|^{2}}\,dy\right|\leq\mathrm{erfc}(M/2)+\mathrm{erfc}(3M/2)

and similarly for the other integral. Here we have used the fact

erfc(1)0.16erfc(3)0.01\displaystyle\mathrm{erfc}(1)\leq 0.16\quad\mathrm{erfc}(3)\leq 0.01

and for M2M\geq 2

|11πMMe|M/2y|2𝑑y1|erfc(1)+erfc(3)2erfc(1)erfc(3).\displaystyle\left|\frac{1}{\frac{1}{\sqrt{\pi}}\int\limits_{-M}^{M}e^{-|M/2-y|^{2}}\,dy}-1\right|\leq\frac{\mathrm{erfc}(1)+\mathrm{erfc}(3)}{2-\mathrm{erfc}(1)-\mathrm{erfc}(3)}.

If we want the extension to be compactly supported we define

ϕ(x)=\displaystyle\phi(x)=
χBM/2(x)+φ[M/2,10M)(x)1πMMe|xy|2𝑑y1πMMe|M/2y|2𝑑y+φ(10M,M/2](x)1πMMe|xy|2𝑑y1πMMe|M/2+y|2𝑑y\displaystyle\chi_{B_{M/2}}(x)+\varphi_{[M/2,10M)}(x)\frac{\frac{1}{\sqrt{\pi}}\int\limits_{-M}^{M}e^{-|x-y|^{2}}\,dy}{\frac{1}{\sqrt{\pi}}\int\limits_{-M}^{M}e^{-|M/2-y|^{2}}\,dy}+\varphi_{(-10M,-M/2]}(x)\frac{\frac{1}{\sqrt{\pi}}\int\limits_{-M}^{M}e^{-|x-y|^{2}}\,dy}{\frac{1}{\sqrt{\pi}}\int\limits_{-M}^{M}e^{-|-M/2+y|^{2}}\,dy}

where φ\varphi is 11 on [M/2,9M)[M/2,9M), smooth on [9M,10M][9M,10M] and 0 elsewhere. We see that for M2M\geq 2

supx|ϕϕ1|erfc(9M).\displaystyle\sup\limits_{x\in\mathbb{R}}\left|\phi-\phi_{1}\right|\leq\mathrm{erfc}(9M).

It follows that

supx|ϕnΔxnπe(xxnM)2|\displaystyle\sup\limits_{x\in\mathbb{R}}\left|\phi-\sum\limits_{n}\frac{\Delta x_{n}}{\sqrt{\pi}}e^{-(x-x_{n}-M)^{2}}\right|
erfc(9M)+erfc(M/2)+2ΔxM2eM2/4+2ΔxM.\displaystyle\leq\mathrm{erfc}(9M)+\mathrm{erfc}(M/2)+2\Delta xM\leq 2e^{-M^{2}/4}+2\Delta xM.

Thus ϕ(x)\phi(x) is the desired extension of χB\chi_{B} if MM is sufficiently large and Δx<4M\Delta x<4M. We can also shift the function further away from the origin if we want all the xnx_{n} positive or negative. ∎

If we want to rescale the ball to MϵM\epsilon then we use χBM/2(xϵ)\chi_{B_{M/2}}(\frac{x}{\epsilon}) with ϵ(0,1)\epsilon\in(0,1) and the Gaussians are also rescaled. The proof of Theorem 4 still holds with minor modifications under this rescaling. Notice that for the dd-dimensional extension ϕL2(d)=𝒪(Md)\|\phi\|_{L^{2}(\mathbb{R}^{d})}=\mathcal{O}(M^{d}), so this error η\eta in (2.9) is relatively quite small compared to the L2L^{2} norm of ϕ\phi. We show in Section 8 that this ϕ\phi is L2LL^{2}\textendash L^{\infty} observable.

5. Representations of Solutions in terms of Gaussian Wavepackets, Proof of Theorem 2

We would like to represent solutions to (2.1) in terms of Gaussians. To do this, we decompose functions in L2(d)L^{2}(\mathbb{R}^{d}) as the sum of a finite number of real gaussians. Then we will apply the FIO from the previous section.

Lemma 4.

The solution to (2.1) with ana_{n} in d\mathbb{R}^{d}

u0=ϕn(x)=ϕn(0,x)=e|x+an|2,\displaystyle u_{0}=\phi_{n}(x)=\phi_{n}(0,x)=e^{-|x+a_{n}|^{2}},

can be written as

u(t,x)=ϕn(t,x)=(a2(t)(14iy3(t)))d2eiy1(t)|x|2|y2(t)x+an|2(14iy3(t)).\displaystyle u(t,x)=\phi_{n}(t,x)=\left(\frac{a^{2}(t)}{(1-4iy_{3}(t))}\right)^{\frac{d}{2}}e^{iy_{1}(t)|x|^{2}-\frac{|y_{2}(t)x+a_{n}|^{2}}{(1-4iy_{3}(t))}}.
Proof.

Recall that

deiyηec|y|2𝑑y=(πc)d2e|η|24c.\displaystyle\int\limits_{\mathbb{R}^{d}}e^{-iy\cdot\eta}e^{-c|y|^{2}}\,dy=\left(\frac{\pi}{c}\right)^{\frac{d}{2}}e^{-\frac{|\eta|^{2}}{4c}}.

The Fourier transform of ϕn(x)\phi_{n}(x) is

2d2eianη|η|24.\displaystyle 2^{-\frac{d}{2}}e^{ia_{n}\cdot\eta-\frac{|\eta|^{2}}{4}}.

The FIO applied to ϕn(x)\phi_{n}(x) is then

(a2(t)4π)d2deiy1(t)|x|2+iy2(t)xη+iy3(t)|η|2+ianη|η|24𝑑η\displaystyle\left(\frac{a^{2}(t)}{4\pi}\right)^{\frac{d}{2}}\int\limits_{\mathbb{R}^{d}}e^{iy_{1}(t)|x|^{2}+iy_{2}(t)x\cdot\eta+iy_{3}(t)|\eta|^{2}+ia_{n}\cdot\eta-\frac{|\eta|^{2}}{4}}\,d\eta
=(a2(t)(14iy3(t)))d2eiy1(t)|x|2|y2(t)x+an|2(14iy3(t))\displaystyle=\left(\frac{a^{2}(t)}{(1-4iy_{3}(t))}\right)^{\frac{d}{2}}e^{iy_{1}(t)|x|^{2}-\frac{|y_{2}(t)x+a_{n}|^{2}}{(1-4iy_{3}(t))}}

which gives the desired result. ∎

We can then construct a nearly explicit parametrix to all u0=f(x)e|x|2/2u_{0}=f(x)e^{-|x|^{2}/2} with f(x)f(x) an L2(d)L^{2}(\mathbb{R}^{d}) function which is interesting in its own right

Proof of Theorem 2.

The approximation in (2.4) holds for any fL2(d)f\in L^{2}(\mathbb{R}^{d}) because hk(x)h_{k}(x) is a well known (complete) orthonormal basis for L2(d)L^{2}(\mathbb{R}^{d}). More explicitly we see, (2.4) holds for a fixed η\eta and fH3([M,M]d)f\in H^{3}([-M,M]^{d}) since we can select NN larger than MM based on the threshold η\eta, using Proposition 3. We then use Corollary 1 to give the two term bound on the error approximating u0u_{0}. The principle of superposition and the construction of the individual ϕn(t,x)\phi_{n}(t,x) in Lemma 4 gives the desired result. The error over time can be bounded since the non-autonomous propagation operator is still an isometry on L2(d)L^{2}(\mathbb{R}^{d}) as a result of [38]. ∎

6. Proof of Theorems 3 and 4 for Observability

From Theorem 1 and Lemma 4, we are able to construct accurate solutions to (2.1) using sophisticated Gaussian wavepackets. Gaussian wavepackets have localization features which will allow us to analyze the observable sets, ω\omega. Therefore, we start by proving some elementary bounds on the solutions to (3.1), provided it is well-posed, which will then allow us to analyze the observability of the solution. A future goal is the development of complex valued initial data into wavepackets which would pose the added challenge that there could be a wavefront set, [39].

Lemma 5.

Assume the Hamiltonian flow is non-zero for all 0<tTD0<t\leq T_{D}. Let K0,KK_{0},K be positive constants such that for all t[0,TD]t\in[0,T_{D}]

|y1(t)|K00<κ1(t)K.\displaystyle|y_{1}(t)|\leq K_{0}\quad\quad 0<\kappa_{1}(t)\leq K.

Then the solutions y2(t)y_{2}(t) and y3(t)y_{3}(t) to (3.1) satisfy the following bounds

e4K0Kty2(t)e4K0Kty3(t)0.\displaystyle e^{-4K_{0}Kt}\leq y_{2}(t)\leq e^{4K_{0}Kt}\quad\quad y_{3}(t)\leq 0. (6.1)
Proof.

The proof follows immediately from Gronwall’s inequality and (3.1). ∎

In practice we will compute the bounds explicitly in Section 7 when we know that the Hamiltonian flow is non-zero. Before proving a lower bound, we recall the following result of [7]. The error function is defined by

erfc(x)=2πxex2𝑑x.\displaystyle\mathrm{erfc}(x)=\frac{2}{\sqrt{\pi}}\int\limits_{x}^{\infty}e^{-x^{2}}\,dx.

We have the following single term lower bound for the complementary error function

erfc(x)2eπβ1βeβx2x0,β>1.\displaystyle\mathrm{erfc}(x)\geq\sqrt{\frac{2e}{\pi}}\sqrt{\frac{\beta-1}{\beta}}e^{-\beta x^{2}}\quad x\geq 0,\,\,\beta>1. (6.2)

For the rest of this paper we make the definitions

A=A(t)=2|y2(t)|2(1+16y32(t))γ(t)=(a2(t)(14iy3(t)))d2\displaystyle A=A(t)=\frac{2|y_{2}(t)|^{2}}{(1+16y_{3}^{2}(t))}\quad\gamma(t)=\left(\frac{a^{2}(t)}{(1-4iy_{3}(t))}\right)^{\frac{d}{2}} (6.3)

which will control the spread of the wave packets and ultimately the constant CTC_{T}. The above results will be useful in proving the lower bound in the next lemma.

Lemma 6.

Let ϕn(t,x)\phi_{n}(t,x) be the same as in Lemma 4 for arbitrary an+a_{n}\in\mathbb{R}^{+}. Let cnc_{n} be a finite collection of real numbers indexed by ndn\in\mathbb{R}^{d}, |n|N|n|\leq N, then have the following lower bound for all 0tT0\leq t\leq T

ϵ(t,R)|n|N|cn||n|N([R,R]d|cnϕn(t,x)|2𝑑x)12\displaystyle\epsilon(t,R)\sum\limits_{|n|\leq N}|c_{n}|\leq\sum\limits_{|n|\leq N}\left(\int\limits_{[-R,R]^{d}}|c_{n}\phi_{n}(t,x)|^{2}\,dx\right)^{\frac{1}{2}}

where ϵ(t,R)\epsilon(t,R) is a positive constant.

Proof.

Define

Bn=any2(t)\displaystyle\quad\quad B_{n}=\frac{a_{n}}{y_{2}(t)}

then we have that

([R,R]d)c|ϕn(t,x)|2𝑑x|γ(t)|2([R,R]d)ce2|y2(t)x+an|2(1+16y32(t))𝑑x\displaystyle\int\limits_{([-R,R]^{d})^{c}}|\phi_{n}(t,x)|^{2}\,dx\geq|\gamma(t)|^{2}\int\limits_{([-R,R]^{d})^{c}}e^{-\frac{2|y_{2}(t)x+a_{n}|^{2}}{(1+16y_{3}^{2}(t))}}\,dx
=|γ(t)|2(π4A)d2ni(h(|Bni|,A,R))\displaystyle=|\gamma(t)|^{2}\left(\frac{\pi}{4A}\right)^{\frac{d}{2}}\prod\limits_{n_{i}}(h(|B_{n_{i}}|,A,R))

where

h(|Bni|,A,R)\displaystyle h(|B_{n_{i}}|,A,R)
={erfc(A(R+|Bni|))+erfc(A(R|Bni|))R>|Bni|erfc(A(R+|Bni|))+(2erfc(A(|Bni|R)))R|Bni|.\displaystyle=\begin{cases}\mathrm{erfc}\left(\sqrt{A}\left(R+\left|B_{n_{i}}\right|\right)\right)+\mathrm{erfc}\left(\sqrt{A}\left(R-\left|B_{n_{i}}\right|\right)\right)&R>\left|B_{n_{i}}\right|\\ \mathrm{erfc}\left(\sqrt{A}\left(R+\left|B_{n_{i}}\right|\right)\right)+\left(2-\mathrm{erfc}\left(\sqrt{A}\left(\left|B_{n_{i}}\right|-R\right)\right)\right)&R\leq\left|B_{n_{i}}\right|.\end{cases}

If R>|Bni|=m0R>\left|B_{n_{i}}\right|=m\geq 0, then from equation (6.2) we have with β=2\beta=2

erfc(A(Rm))eπe2A(Rm)2eπe2AR2.\displaystyle\mathrm{erfc}\left(\sqrt{A}(R-m)\right)\geq\sqrt{\frac{e}{\pi}}e^{-2A(R-m)^{2}}\geq\sqrt{\frac{e}{\pi}}e^{-2AR^{2}}.

If |Bni|R0\left|B_{n_{i}}\right|\geq R\geq 0 then

(2erfc(A(mR)))>1.\displaystyle\left(2-\mathrm{erfc}(\sqrt{A}(m-R))\right)>1.

As a result we obtain

ϵ(t,R)=(π4A)d4|γ(t)|min{1,eπ4eAR2}.\displaystyle\epsilon(t,R)=\left(\frac{\pi}{4A}\right)^{\frac{d}{4}}|\gamma(t)|\min\left\{1,\sqrt[4]{\frac{e}{\pi}}e^{-AR^{2}}\right\}. (6.4)

Noticing that |γ(t)|=(A2)d4|\gamma(t)|=\left(\frac{A}{2}\right)^{\frac{d}{4}} we can simplify further

ϵ(t,R)=min{πd14e1423d4eAR2,(π8)d4}.\displaystyle\epsilon(t,R)=\min\left\{\frac{\pi^{\frac{d-1}{4}}e^{\frac{1}{4}}}{2^{\frac{3d}{4}}}e^{-AR^{2}},\left(\frac{\pi}{8}\right)^{\frac{d}{4}}\right\}. (6.5)

By combining Lemmata 2 and 5, AA is bounded above and below for all 0tT0\leq t\leq T depending on the L([0,T])L^{\infty}([0,T]) norms of κ1(t),κ2(t)\kappa_{1}(t),\kappa_{2}(t), implying ϵ(t,R)\epsilon(t,R) is positive. An explicit lower bound for (6.5) is computed for various examples in Section 7. ∎

Recall the Diaz-Metcalf inequality

Lemma 7.

[Diaz-Metcalf [13]] Let ww be a unit vector in the inner product space (H;,)(H;\langle\cdot,\cdot\rangle) over the real or complex number field. Suppose that the vectors vnH0v_{n}\in H\setminus{0}, i{1,..,N}i\in\{1,..,N\} satisfy

0rRevn,wvn\displaystyle 0\leq r\leq\frac{\mathrm{Re}\langle v_{n},w\rangle}{\|v_{n}\|} (6.6)

then

ri=1Nvni=1Nvn.\displaystyle r\sum\limits_{i=1}^{N}\|v_{n}\|\leq\|\sum\limits_{i=1}^{N}v_{n}\|. (6.7)

Using the above lemma we have can find a lower bound on the L2((0,T)×ω)L^{2}((0,T)\times\omega) norm of u(t,x)u(t,x) in terms of the Gaussian wave packets which suits our needs. The L2((0,T)×ω)L^{2}((0,T)\times\omega) norm is difficult to work with directly when the initial data is propagated into complex wavepackets, because their inner products take the form of Frensel integrals. This is the reason for the intermediate Lemmata. We make the following choice of ana_{n}.

Lemma 8.

There is a choice of ε\varepsilon in terms of a fixed natural number NN such that for an=εna_{n}=\varepsilon n with n=(n1,,nd)n=(n_{1},...,n_{d}), |n|N|n|\leq N ni{1,2,N}n_{i}\in\{1,2,...N\} and a ψ1(x)\psi_{1}(x) independent of nn such that

Re|x|>R0(γ(t))1ϕn(t,x)ψ1(x)¯𝑑xδ~(t,R0)>0\displaystyle\mathrm{Re}\int\limits_{|x|>R_{0}}(\gamma(t))^{-1}\phi_{n}(t,x)\overline{\psi_{1}(x)}\,dx\geq\tilde{\delta}(t,R_{0})>0

for all of these nn, where BR0B_{R_{0}} is the ball of radius R0R_{0} with ψ1L2(BR0c)=1||\psi_{1}||_{L^{2}(B_{R_{0}}^{c})}=1.

Proof.

Select

ψ0(x)=eA2|x|22iy3(t)A|x|2+iy1(t)|x|2|x|d1\displaystyle\psi_{0}(x)=\frac{e^{-\frac{A}{2}|x|^{2}-2iy_{3}(t)A|x|^{2}+iy_{1}(t)|x|^{2}}}{|x|^{d-1}}

then we have that

|x|>R0|ψ0(x)|2𝑑x=|x|>R0eA|x|2|x|2(d1)𝑑x=(C(R0,A))2\displaystyle\int\limits_{|x|>R_{0}}|\psi_{0}(x)|^{2}\,dx=\int\limits_{|x|>R_{0}}\frac{e^{-A|x|^{2}}}{|x|^{2(d-1)}}\,dx=(C(R_{0},A))^{2}

where C(R0,A)=𝒪(eR02A/2)C(R_{0},A)=\mathcal{O}(e^{-R_{0}^{2}A/2}). We set ψ0=(C(R0,A))ψ1\psi_{0}=(C(R_{0},A))\psi_{1}. It follows that

|x|>R0ϕn(t,x)ψ0(x)¯𝑑x=|x|>R0eA|x|2+A~y2(t)anxA~2|an|2|x|d1𝑑x\displaystyle\int\limits_{|x|>R_{0}}\phi_{n}(t,x)\overline{\psi_{0}(x)}\,dx=\int\limits_{|x|>R_{0}}\frac{e^{-A|x|^{2}+\tilde{A}y_{2}(t)a_{n}\cdot x-\frac{\tilde{A}}{2}|a_{n}|^{2}}}{|x|^{d-1}}\,dx
=θ𝕊d1r>R0eAr2+A~y2(t)anb(θ)rA~2an2𝑑r𝑑θ.\displaystyle=\int\limits_{\theta\in\mathbb{S}^{d-1}}\int\limits_{r>R_{0}}e^{-Ar^{2}+\tilde{A}y_{2}(t)a_{n}\cdot b(\theta)r-\frac{\tilde{A}}{2}a_{n}^{2}}\,dr\,d\theta.

Here A~=2(14iy3(t))1\tilde{A}=2(1-4iy_{3}(t))^{-1} and b(θ)b(\theta) is a vector depending on θ𝕊d1\theta\in\mathbb{S}^{d-1} with norm 1. It follows that if an=εna_{n}=\varepsilon n, where we have control over the parameter ε\varepsilon which is small, then we can Taylor expand this integral in terms of ε\varepsilon. Set

|x|>R0eA|x|2+A~y2(t)anxA~2|an|2|x|d1𝑑x=f(ε,A,R0,n).\displaystyle\int\limits_{|x|>R_{0}}\frac{e^{-A|x|^{2}+\tilde{A}y_{2}(t)a_{n}\cdot x-\frac{\tilde{A}}{2}|a_{n}|^{2}}}{|x|^{d-1}}\,dx=f(\varepsilon,A,R_{0},n).

Let C(d)C(d) be the volume of the unit ball in dd dimensions. The leading order term is

f(0,A,R0,n)=C(d)r>R0eAr2𝑑r,\displaystyle f(0,A,R_{0},n)=C(d)\int\limits_{r>R_{0}}e^{-Ar^{2}}\,dr,

which is independent of nn. The error in this approximation is bounded by

|supε(0,1)ddεf(ε,A,R0,n)|\displaystyle\left|\sup\limits_{\varepsilon\in(0,1)}\frac{d}{d\varepsilon}f(\varepsilon,A,R_{0},n)\right|\leq
supε(0,1)r>R0C(d)|A~n|(1+r|y2(t)|)eAr2+A1|n|r𝑑r=C(d)nC2(A1,R0,n).\displaystyle\sup\limits_{\varepsilon\in(0,1)}\int\limits_{r>R_{0}}C(d)|\tilde{A}n|(1+r|y_{2}(t)|)e^{-Ar^{2}+A_{1}|n|r}\,dr=C(d)nC_{2}(A_{1},R_{0},n).

with A1=2|y2(t)(1+16y32(t))1|A_{1}=2|y_{2}(t)(1+16y_{3}^{2}(t))^{-1}|. For positive constants a,b,c,ma,b,c,m we have that

m(a+bx)ecxx2𝑑x=ec244(π(2a+bc)(erf(12(c2m))+1)+2be14(c2m)2).\displaystyle\int\limits_{m}^{\infty}(a+bx)e^{cx-x^{2}}\,dx=\frac{e^{\frac{c^{2}}{4}}}{4}\left(\sqrt{\pi}(2a+bc)\left(\mathrm{erf}\left(\frac{1}{2}(c-2m)\right)+1\right)+2be^{-\frac{1}{4}(c-2m)^{2}}\right). (6.8)

Using the substitution

a=1b=|y2(t)|Ac=A1nAAR0=m,\displaystyle a=1\quad b=\frac{|y_{2}(t)|}{\sqrt{A}}\quad c=\frac{A_{1}n}{\sqrt{A}}\quad\sqrt{A}R_{0}=m,

we find

supnNC2(A1,R0,n)\displaystyle\sup\limits_{n\leq N}C_{2}(A_{1},R_{0},n)\leq
C(d)|A~|2A(|y2(t)|AeAR02+A1NR0+2πeA12N2A(1+A1NA|y2(t)|)).\displaystyle\frac{C(d)|\tilde{A}|}{2\sqrt{A}}\left(\frac{|y_{2}(t)|}{\sqrt{A}}e^{-AR_{0}^{2}+A_{1}NR_{0}}+2\sqrt{\pi}e^{\frac{A_{1}^{2}N^{2}}{A}}\left(1+\frac{A_{1}N}{A}|y_{2}(t)|\right)\right).

Plugging in the values of AA, A~\tilde{A} and A1A_{1} in terms of the Riccati equations we obtain

C(d)1supnNC2(A1,R0,n)\displaystyle C(d)^{-1}\sup\limits_{n\leq N}C_{2}(A_{1},R_{0},n)\leq (6.9)
1+16y3(t)22|y2(t)|e2|y2(t)|2R02+2|y2(t)|NR01+16y32(t)+2π(N+1|y2(t)|)e2N21+16y32(t).\displaystyle\frac{\sqrt{1+16y_{3}(t)^{2}}}{2|y_{2}(t)|}e^{\frac{2|y_{2}(t)|^{2}R_{0}^{2}+2|y_{2}(t)|NR_{0}}{1+16y_{3}^{2}(t)}}+\sqrt{2\pi}\left(\frac{N+1}{|y_{2}(t)|}\right)e^{\frac{2N^{2}}{1+16y_{3}^{2}(t)}}.

This gives

1C(R0,A)supnNRe|x|>R0ϕn(t,x)ψ0(x)¯𝑑x=C(d)r>R0eAr2𝑑rC(R0,A)+𝒪(ε)=δ~(t,R0)>0,\displaystyle\frac{1}{C(R_{0},A)}\sup\limits_{n\leq N}\operatorname{Re}\int\limits_{|x|>R_{0}}\phi_{n}(t,x)\overline{\psi_{0}(x)}\,dx=C(d)\frac{\int\limits_{r>R_{0}}e^{-Ar^{2}}\,dr}{C(R_{0},A)}+\mathcal{O}(\varepsilon)=\tilde{\delta}(t,R_{0})>0,

by our choice of ε\varepsilon sufficiently small. Notice that

C(d)r>R0eAr2𝑑rC(R0,A,n)R0d12(erfc(AR0)2A)12\displaystyle C(d)\frac{\int\limits_{r>R_{0}}e^{-Ar^{2}}\,dr}{C(R_{0},A,n)}\geq R_{0}^{\frac{d-1}{2}}\left(\frac{\mathrm{erfc}(\sqrt{A}R_{0})}{2\sqrt{A}}\right)^{\frac{1}{2}}

for R0>1R_{0}>1. We need to show that the 𝒪(ε)\mathcal{O}(\varepsilon) terms are not so large so that δ~(t,R0)\tilde{\delta}(t,R_{0}) is positive. Using Taylor’s theorem we see that this can be accomplished by using (6.9) to enforce the condition

1+16y3(t)22|y2(t)|e2|y2(t)|2R02+2|y2(t)|NR01+16y32(t)+2π(N+1|y2(t)|)e2N21+16y32(t)\displaystyle\frac{\sqrt{1+16y_{3}(t)^{2}}}{2|y_{2}(t)|}e^{\frac{2|y_{2}(t)|^{2}R_{0}^{2}+2|y_{2}(t)|NR_{0}}{1+16y_{3}^{2}(t)}}+\sqrt{2\pi}\left(\frac{N+1}{|y_{2}(t)|}\right)e^{\frac{2N^{2}}{1+16y_{3}^{2}(t)}}\leq (6.10)
12r>R0eAr2𝑑r.\displaystyle\frac{1}{2}\int\limits_{r>R_{0}}e^{-Ar^{2}}\,dr.

We have for ε\varepsilon sufficiently small that

2δ~(t,R0)>R0d12(erfc(AR0)2A)12>R0d12(e4Aπ)14eAR02.\displaystyle 2\tilde{\delta}(t,R_{0})>R_{0}^{\frac{d-1}{2}}\left(\frac{\mathrm{erfc}(\sqrt{A}R_{0})}{2\sqrt{A}}\right)^{\frac{1}{2}}>R_{0}^{\frac{d-1}{2}}\left(\frac{e}{4A\pi}\right)^{\frac{1}{4}}e^{-AR_{0}^{2}}. (6.11)

This finishes the proof. In practice the constants in time in (6.11) and (6.10) are computable and this is done in Section 7 for various examples. ∎

Using these lemmata we can show that in essence it is possible to ignore some of the problems caused by the off diagonal terms for certain f(x)f(x). Let R0R_{0} and RR be large enough such that ΩBR0[R,R]d\Omega\subset B_{R_{0}}\subset[-R,R]^{d}.

Lemma 9.

Let

δ(t,R0)=e14Ad14R0d122d4+1(4π)14eAR02\displaystyle\delta(t,R_{0})=\frac{e^{\frac{1}{4}}A^{\frac{d-1}{4}}R_{0}^{\frac{d-1}{2}}}{2^{\frac{d}{4}+1}(4\pi)^{\frac{1}{4}}}e^{-AR_{0}^{2}} (6.12)

Assume that ana_{n} satisfies Lemma 8, and cnc_{n} are finite real positive numbers then we have the following inequalities

(π2)d40Tδ(t,R0)|n|Ncnϕn(t,)L2(BR0c)dt\displaystyle\left(\frac{\pi}{2}\right)^{-\frac{d}{4}}\int\limits_{0}^{T}\delta(t,R_{0})\sum\limits_{|n|\leq N}||c_{n}\phi_{n}(t,\cdot)||_{L^{2}(B_{R_{0}}^{c})}\,dt\leq (6.13)
0T(ω||n|Ncnϕn(t,x)|2)12𝑑tT12(0Tω||n|Ncnϕn(t,x)|2𝑑x𝑑t)12.\displaystyle\int\limits_{0}^{T}\left(\int\limits_{\omega}\left|\sum\limits_{|n|\leq N}c_{n}\phi_{n}(t,x)\right|^{2}\right)^{\frac{1}{2}}\,dt\leq T^{\frac{1}{2}}\left(\int\limits_{0}^{T}\int\limits_{\omega}\left|\sum\limits_{|n|\leq N}c_{n}\phi_{n}(t,x)\right|^{2}\,dx\,dt\right)^{\frac{1}{2}}.
Proof.

For the first inequality, we note that cnc_{n} is real valued and γ(t)\gamma(t) is the same factor for all the propagated wavepackets then applying Lemma 8 gives the first inequality, provided ΩBR0\Omega\subset B_{R_{0}} and vn=cnϕn(t,x)v_{n}=c_{n}\phi_{n}(t,x) and δ~|γ(t)|=δ\tilde{\delta}|\gamma(t)|=\delta. The last inequality in the Lemma follows by Cauchy Schwartz. ∎

The 1\ell^{1} norm of the cnsc_{n}^{\prime}s can be related to the L2(d)L^{2}(\mathbb{R}^{d}) norm of ff as follows.

Lemma 10.

We have that for ff in 𝒜\mathcal{A} that there are cnc_{n} real positive numbers and ϕn(t,x)\phi_{n}(t,x) as in Lemma 4 so that

fL2(ddx)(π2)d4|n|N|cn|+η.\displaystyle\|f\|_{L^{2}(\mathbb{R}^{d}\,dx)}\leq\left(\frac{\pi}{2}\right)^{\frac{d}{4}}\sum\limits_{|n|\leq N}|c_{n}|+\eta.
Proof.

Using Minkowski’s inequality and the normalization of the ϕn(0)s\phi_{n}(0)^{\prime}s we see

(d||n|N|cnϕn(0,x)|2dx)12|n|N(d|cnϕn(0,x)|2dx)12=(π2)d4|n|N|cn|.\displaystyle\left(\int\limits_{\mathbb{R}^{d}}|\sum\limits_{|n|\leq N}|c_{n}\phi_{n}(0,x)|^{2}\,dx\right)^{\frac{1}{2}}\leq\sum\limits_{|n|\leq N}\left(\int\limits_{\mathbb{R}^{d}}|c_{n}\phi_{n}(0,x)|^{2}\,dx\right)^{\frac{1}{2}}=\left(\frac{\pi}{2}\right)^{\frac{d}{4}}\sum\limits_{|n|\leq N}|c_{n}|.

The last inequality comes from assuming that f𝒜f\in\mathcal{A}. ∎

Proof of Theorem 3.

Assume that ff satisfies Assumption 3 so that for some positive real numbers cnc_{n} we can write

f(x)=|n|Ncnϕn(x)+η(x)\displaystyle f(x)=\sum\limits_{|n|\leq N}c_{n}\phi_{n}(x)+\eta(x) (6.14)

with η(x)\eta(x) having L2(d)L^{2}(\mathbb{R}^{d}) norm less than η\eta. (The equality is understood in the sense of L2(d)L^{2}(\mathbb{R}^{d}) norms). Moreover from Lemma 4 and linearity we have

u(t,x)=|n|Ncnϕn(t,x)+η(t,x)\displaystyle u(t,x)=\sum\limits_{|n|\leq N}c_{n}\phi_{n}(t,x)+\eta(t,x) (6.15)

with η(t,x)\eta(t,x) having L2(d)L^{2}(\mathbb{R}^{d}) norm less than η\eta for all 0tT0\leq t\leq T. Now we proceed by a bit of a bootstrap argument. We additionally assume that an=εna_{n}=\varepsilon n are chosen such that Lemma 6 holds by the hypothesis on ε\varepsilon. By application of equalities (6.14) and (6.15) followed by Lemmata 6, 9, and 10 in direct succession we have that

T12(uL2((0,T)×ω)+ηT)\displaystyle T^{\frac{1}{2}}\left(\|u\|_{L^{2}((0,T)\times\omega)}+\eta T\right)
T12(0Tω||n|Ncnϕn(t,x)|2𝑑x𝑑t)12\displaystyle\geq T^{\frac{1}{2}}\left(\int\limits_{0}^{T}\int\limits_{\omega}|\sum\limits_{|n|\leq N}c_{n}\phi_{n}(t,x)|^{2}\,dx\,dt\right)^{\frac{1}{2}}
(π2)d4(0Tϵ(t,R)δ(t,R0)𝑑t)(|n|N|cn|)\displaystyle\geq\left(\frac{\pi}{2}\right)^{-\frac{d}{4}}\left(\int\limits_{0}^{T}\epsilon(t,R)\delta(t,R_{0})\,dt\right)\left(\sum\limits_{|n|\leq N}|c_{n}|\right)
(π2)d2(0Tϵ(t,R)δ(t,R0)𝑑t)(fL2(d)η).\displaystyle\geq\left(\frac{\pi}{2}\right)^{-\frac{d}{2}}\left(\int\limits_{0}^{T}\epsilon(t,R)\delta(t,R_{0})\,dt\right)(\|f\|_{L^{2}(\mathbb{R}^{d})}-\eta).

This concludes the proof of the theorem. The constant in the inequality is defined as

CT=(π2)d2T12(0Tϵ(t,R)δ(t,R0)𝑑t)1\displaystyle C_{T}=\left(\frac{\pi}{2}\right)^{\frac{d}{2}}T^{\frac{1}{2}}\left(\int\limits_{0}^{T}\epsilon(t,R)\delta(t,R_{0})\,dt\right)^{-1} (6.16)
=T12(0Te12Ad14R0d12πd+242d+32eA(R02+R2)𝑑t)1.\displaystyle=T^{\frac{1}{2}}\left(\int\limits_{0}^{T}\frac{e^{\frac{1}{2}}A^{\frac{d-1}{4}}R_{0}^{\frac{d-1}{2}}}{\pi^{\frac{d+2}{4}}2^{\frac{d+3}{2}}}e^{-A(R_{0}^{2}+R^{2})}\,dt\right)^{-1}.

Proof of Theorem 4.

Let ana_{n} be in +d\mathbb{R}^{d}_{+} and cn+c_{n}\in\mathbb{R}_{+}. By the hypothesis on the initial data u0(x)u_{0}(x) that it is in 𝒜\mathcal{A} we have that

u0(x)=|n|Ncne|x+an|2=|n|Ncnϕn(0,x)+η(x)\displaystyle u_{0}(x)=\sum\limits_{|n|\leq N}c_{n}e^{-|x+a_{n}|^{2}}=\sum\limits_{|n|\leq N}c_{n}\phi_{n}(0,x)+\eta(x) (6.17)

in the sense of L2(d)L^{2}(\mathbb{R}^{d}) norms for some η(x)\eta(x) with norm less than η\eta. This allows us to write

u(t,)L2(ω)2=n,mcncmϕn(t,),ϕm(t,)L2(ω)+η\displaystyle||u(t,\cdot)||^{2}_{L^{2}(\omega)}=\sum\limits_{n,m}c_{n}c_{m}\langle\phi_{n}(t,\cdot),\phi_{m}(t,\cdot)\rangle_{L^{2}(\omega)}+\eta^{\prime} (6.18)

where ηη2\eta^{\prime}\leq\eta^{2}. Let

|αN|=maxn,m|anam|.|\alpha_{N}|=\max\limits_{n,m}|a_{n}-a_{m}|.

Let ana_{n^{\prime}} be such that

R1=|an|=maxm|am|.R_{1}=|a_{n^{\prime}}|=\max\limits_{m}|a_{m}|.

We consider ana_{n^{\prime}} as the new origin in the coordinate plane. Under this coordinate change we have that ϕn\phi_{n} becomes ϕ~n\tilde{\phi}_{n} where ϕ~n(t,x)\tilde{\phi}_{n}(t,x) have a~n\tilde{a}_{n} as their centers. The number R2R_{2} we define as R2=R1diam(Ω)2R_{2}=R_{1}-\frac{\mathrm{diam}(\Omega)}{2} which is the distance from this new center ana_{n^{\prime}} to the side of the ball enclosing Ω\Omega.

Refer to caption
Figure 1. The Gaussians at time t=0t=0 are centered at ana_{n} n=1,2,3,4n=1,2,3,4 and have their L2(2)L^{2}(\mathbb{R}^{2}) norm largely within a ball of radius 12\frac{1}{\sqrt{2}}. For Ω=BR3\Omega=B_{R_{3}}, then we see if R1R3=R2R_{1}-R_{3}=R_{2} and 2R2R22R_{2}^{\prime}\leq R_{2} these smaller balls do not intersect BR3B_{R_{3}}, or the plane at a distance R2R_{2} from a4a_{4} containing it. In this case, R212maxn,m|anam|+12R_{2}^{\prime}\geq\frac{1}{2}\max\limits_{n,m}|a_{n}-a_{m}|+\frac{1}{\sqrt{2}}, and we should be able to compare the L2L^{2} norms of the sum of the Gaussians over the area with the plane removed, to their L2(d)L^{2}(\mathbb{R}^{d}) norms. When the coordinates and radii of the balls are rescaled in time, the same principle still holds which is what is behind the proof of Theorem 4.

Removing the plane containing the ball gives

n,mcncmϕn(t,),ϕm(t,)L2(ω)n,mcncmϕ~n(t,),ϕ~m(t,)L2(d[R2,)d).\displaystyle\sum\limits_{n,m}c_{n}c_{m}\langle\phi_{n}(t,\cdot),\phi_{m}(t,\cdot)\rangle_{L^{2}(\omega)}\geq\sum\limits_{n,m}c_{n}c_{m}\langle\tilde{\phi}_{n}(t,\cdot),\tilde{\phi}_{m}(t,\cdot)\rangle_{L^{2}(\mathbb{R}^{d}\setminus[R_{2},\infty)^{d})}. (6.19)

We will show that

2e|[R2,)dϕ~n(t,x)ϕ~m(t,x)¯𝑑x|2e[R2,)d|ϕ~n(t,x)ϕ~m(t,x)¯|𝑑x\displaystyle 2e|\int\limits_{[R_{2},\infty)^{d}}\tilde{\phi}_{n}(t,x)\overline{\tilde{\phi}_{m}(t,x)}\,dx|\leq 2e\int\limits_{[R_{2},\infty)^{d}}|\tilde{\phi}_{n}(t,x)\overline{\tilde{\phi}_{m}(t,x)}|\,dx (6.20)
ϕ~n(0,),ϕ~m(0,)L2(d)=ϕn(0,),ϕm(0,)L2(d).\displaystyle\leq\langle\tilde{\phi}_{n}(0,\cdot),\tilde{\phi}_{m}(0,\cdot)\rangle_{L^{2}(\mathbb{R}^{d})}=\langle\phi_{n}(0,\cdot),\phi_{m}(0,\cdot)\rangle_{L^{2}(\mathbb{R}^{d})}.

The main result will then follow since

n,mcncmϕ~n(t,),ϕ~m(t,)L2(d[R2,)d)=\displaystyle\sum\limits_{n,m}c_{n}c_{m}\langle\tilde{\phi}_{n}(t,\cdot),\tilde{\phi}_{m}(t,\cdot)\rangle_{L^{2}(\mathbb{R}^{d}\setminus[R_{2},\infty)^{d})}= (6.21)
n,mcncmϕ~n(t,),ϕ~m(t,)L2(d)n,mcncmϕ~n(t,),ϕ~m(t,)L2([R2,)d)\displaystyle\sum\limits_{n,m}c_{n}c_{m}\langle\tilde{\phi}_{n}(t,\cdot),\tilde{\phi}_{m}(t,\cdot)\rangle_{L^{2}(\mathbb{R}^{d})}-\sum\limits_{n,m}c_{n}c_{m}\langle\tilde{\phi}_{n}(t,\cdot),\tilde{\phi}_{m}(t,\cdot)\rangle_{L^{2}([R_{2},\infty)^{d})}

combined with (6.19) and (6.20) will give the result after integrating all the inequalities over tt. In order to compute

[R2,)d|ϕ~n(t,x)ϕ~m(t,x)¯|𝑑x\displaystyle\int\limits_{[R_{2},\infty)^{d}}|\tilde{\phi}_{n}(t,x)\overline{\tilde{\phi}_{m}(t,x)}|\,dx (6.22)

we recall that

R2eD0|xc~|2D0|xd~|2𝑑x\displaystyle\int\limits_{R_{2}}^{\infty}e^{-D_{0}|x-\tilde{c}|^{2}-D_{0}|x-\tilde{d}|^{2}}\,dx
=(π22D0)di=1d(erfc(2D0(R2ci~+di~2)))eD0|c~d~|22.\displaystyle=\left(\frac{\sqrt{\pi}}{2\sqrt{2D_{0}}}\right)^{d}\prod\limits_{i=1}^{d}\left(\mathrm{erfc}\left(\sqrt{2D_{0}}\left(R_{2}-\frac{\tilde{c_{i}}+\tilde{d_{i}}}{2}\right)\right)\right)e^{-\frac{D_{0}|\tilde{c}-\tilde{d}|^{2}}{2}}.

Let 2D0=A2D_{0}=A with AA as in (6.3) and the vectors c~\tilde{c} and d~\tilde{d} we set equal to c~=a~n/(y2(t))\tilde{c}=-\tilde{a}_{n}/(y_{2}(t)) and d~=a~m/y2(t)\tilde{d}=-\tilde{a}_{m}/y_{2}(t) to match with (6.22). Using that erfc(x)ex2\mathrm{erfc}(x)\leq e^{-x^{2}} for all x>0x>0 if we have that

e2D0(R2ci~+di~2)2eD0|c~id~i|22D0e1e(aniami)22\displaystyle e^{-2D_{0}\left(R_{2}-\frac{\tilde{c_{i}}+\tilde{d_{i}}}{2}\right)^{2}}e^{-\frac{D_{0}|\tilde{c}_{i}-\tilde{d}_{i}|^{2}}{2}}\leq\sqrt{D_{0}}e^{-1}e^{-\frac{(a_{n_{i}}-a_{m_{i}})^{2}}{2}} (6.23)

then (6.20) follows immediately. Whenever D01D_{0}\geq 1 we can find R2R_{2} such that

2D0(R2ci~+di~2)2>1log(D0).\displaystyle 2D_{0}\left(R_{2}-\frac{\tilde{c_{i}}+\tilde{d_{i}}}{2}\right)^{2}>1-\log(\sqrt{D_{0}}). (6.24)

However it is not guaranteed that D01D_{0}\geq 1. Therefore we take

4D0(R2ci~+di~2)2>(anam)2+22log(D0)\displaystyle 4D_{0}\left(R_{2}-\frac{\tilde{c_{i}}+\tilde{d_{i}}}{2}\right)^{2}>(a_{n}-a_{m})^{2}+2-2\log(\sqrt{D_{0}})

as our condition instead. We now have c~=anR1y2(t)-\tilde{c}=\frac{a_{n}-R_{1}}{y_{2}(t)} and d~=amR1y2(t)-\tilde{d}=\frac{a_{m}-R_{1}}{y_{2}(t)} by definition. This results in

R2>|αN|2+2log(D0)4D0+|αN|y2(t)\displaystyle R_{2}>\sqrt{\frac{|\alpha_{N}|^{2}+2-\log(\sqrt{D_{0}})}{4D_{0}}}+\frac{|\alpha_{N}|}{y_{2}(t)}

We note that at t=0,D0=1,y2(0)=1t=0,D_{0}=1,y_{2}(0)=1 from (3.2) and this would mean the expression for R2R_{2} found from (6.24) would exactly coincide with the value found pictorially in Figure 1.

Plugging in D00D_{0}\neq 0 in terms of the Riccati equation solutions we see that this implies

R1>|αN|2+22log(|y2(t)|(1+16(y3(t))2)1)4(y2(t))2(1+16(y3(t))2)1+|αN|y2(t)+diam(Ω)2\displaystyle R_{1}>\sqrt{\frac{|\alpha_{N}|^{2}+2-2\log(\sqrt{|y_{2}(t)|(1+16(y_{3}(t))^{2})^{-1}})}{4(y_{2}(t))^{2}(1+16(y_{3}(t))^{2})^{-1}}}+\frac{|\alpha_{N}|}{y_{2}(t)}+\frac{\mathrm{diam}(\Omega)}{2} (6.25)

Taking the maximum of the right hand side in t[0,T)t\in[0,T) with TT as in Definition 2, this finishes the proof by combining (6.20) and (6.21) and integrating over time. In this case CT=e(e1)TC_{T}=\frac{e}{(e-1)T} in Theorem 4. ∎

Demonstration of Counterexample 2.

The proof consists of two counterexamples. Let δ>0\delta>0 be a constant. We consider as the initial data to (2.1) the Gaussian

e|x+δ|2\displaystyle e^{-|x+\delta|^{2}}

By direct computation in the notation of Lemma 6 we obtain

[R,)d|u(t,x)|2𝑑x=(π4A)d2|γ(t)|2(erfc(A(δy2+R)))d\displaystyle\int\limits_{[R,\infty)^{d}}|u(t,x)|^{2}\,dx=\left(\frac{\pi}{4A}\right)^{\frac{d}{2}}|\gamma(t)|^{2}\left(\mathrm{erfc}\left(\sqrt{A}\left(\frac{\delta}{y_{2}}+R\right)\right)\right)^{d}

as δ\delta\rightarrow\infty, this goes to 0 if y2>0y_{2}>0. As a result, no lower bound on uL2((0,T)×ω)\|u\|_{L^{2}((0,T)\times\omega)} exists. Therefore it is necessary that y2y_{2} change sign if ω\omega is consisting of the right or left half space minus a compact set. The same set up can be use if ω\omega is instead a bounded domain, which violates the condition of Definition 4. For fixed δ\delta and some finite R1,R2R_{1},R_{2}

ω|u(t,x)|2𝑑x[R1,R2]d|u(t,x)|2𝑑x\displaystyle\int\limits_{\omega}|u(t,x)|^{2}\,dx\leq\int\limits_{[R_{1},R_{2}]^{d}}|u(t,x)|^{2}\,dx
(π4A)d2|γ(t)|2(erf(A(δy2+R1))erf(A(δy2R2)))d\displaystyle\leq\left(\frac{\pi}{4A}\right)^{\frac{d}{2}}|\gamma(t)|^{2}\left(\mathrm{erf}\left(\sqrt{A}\left(\frac{\delta}{y_{2}}+R_{1}\right)\right)-\mathrm{erf}\left(\sqrt{A}\left(\frac{\delta}{y_{2}}-R_{2}\right)\right)\right)^{d}

which as δ\delta goes to infinity is again 0. ∎

7. Applications

Many of the examples in this section are taken from the physics literature, see [23] for more background on these examples and others. The collection of these examples is the proof of Proposition 1.

Example 3 (The harmonic oscillator and free Schrödinger revisited).

For the free Schrödinger equation, TD=T_{D}=\infty. Using (6.2), the conditions for (6.10) and (2.10) for y2(t)=1y_{2}(t)=1 and y3(t)=ty_{3}(t)=-t are seen hold under the stronger conditions

εN((1+16t2)eR02+NR08t+(4N+4)eN28t)110e4R021+16t2\displaystyle\varepsilon N((\sqrt{1+16t^{2}})e^{\frac{R_{0}^{2}+NR_{0}}{8t}}+\left(4N+4\right)e^{\frac{N^{2}}{8t}})\leq\frac{1}{10}e^{-\frac{4R_{0}^{2}}{1+16t^{2}}} (7.1)

and

R2(|αN|2+2+2ln1+16t2)(1+16t2)2+|αN|\displaystyle R_{2}\geq\frac{\sqrt{(|\alpha_{N}|^{2}+2+2\mathrm{ln}\sqrt{1+16t^{2}})(1+16t^{2})}}{2}+|\alpha_{N}| (7.2)

respectively for all finite tt. Notice that in (6.3)

A=21+16t2A2\displaystyle A=\frac{2}{\sqrt{1+16t^{2}}}\quad A\leq 2 (7.3)

and therefore min0tTϵ(t,R)\min\limits_{0\leq t\leq T}\epsilon(t,R) and min0tTδ(t,R0)\min\limits_{0\leq t\leq T}\delta(t,R_{0}) as defined by (6.5) and (6.12) are positive, so CTC_{T} in (6.16) is bounded.

For κ1(t)=κ2(t)=12\kappa_{1}(t)=\kappa_{2}(t)=\frac{1}{2}, the standard harmonic oscillator, we have the hypothesis of Theorem 2 are satisfied for all time 0t<π20\leq t<\frac{\pi}{2}. Notice that the construction and stability estimates could be extended past the cusp in this case using exercise 11.1 p.129 in [15] to include all times tt in (0,π)(0,\pi) which is sufficient since it is known the operator has periodic L2((0,T)×ω)L^{2}((0,T)\times\omega) norm. Using (6.2), notice that for this example the condition (6.10) would certainly be satisfied for ε\varepsilon satisfying

2εN(e2R02+2NR0+(2N+2)e2N2)110e4R02\displaystyle 2\varepsilon N\left(e^{2R_{0}^{2}+2NR_{0}}+(2N+2)e^{2N^{2}}\right)\leq\frac{1}{10}e^{-4R_{0}^{2}} (7.4)

uniformly for t[0,π2)t\in[0,\frac{\pi}{2}). Also R1R_{1} satisfies the inequality (2.10) if

R22|αN|2+6+|αN|\displaystyle R_{2}\geq 2\sqrt{|\alpha_{N}|^{2}+6}+|\alpha_{N}| (7.5)

which is finite for finite |αN||\alpha_{N}|. Notice that in (6.3)

A=21+15sin2(t)12A2\displaystyle A=\frac{2}{\sqrt{1+15\sin^{2}(t)}}\quad\quad\frac{1}{2}\leq A\leq 2 (7.6)

and therefore min0tTϵ(t,R)\min\limits_{0\leq t\leq T}\epsilon(t,R) and min0tTδ(t,R0)\min\limits_{0\leq t\leq T}\delta(t,R_{0}) are positive, so CTC_{T} in (6.16) is bounded.

Example 4 (Potentials).

Consider the Hamiltonian with σ\sigma a constant

H(t)=|p|22(t+d)a+σ2(t+d)b|x|22,a,b,d>0\displaystyle H(t)=\frac{|p|^{2}}{2(t+d)^{a}}+\frac{\sigma^{2}(t+d)^{b}|x|^{2}}{2},\quad a,b,d>0

The equations of motion give

d2x(t)dt2+at+ddxdt+σ2(t+d)bax(t)=0.\displaystyle\frac{d^{2}x(t)}{dt^{2}}+\frac{a}{t+d}\frac{dx}{dt}+\sigma^{2}(t+d)^{b-a}x(t)=0.

The solutions are

x(t)=C1(t+d)αJν(z)+C2(t+d)αYν(z)\displaystyle x(t)=C_{1}(t+d)^{\alpha}J_{\nu}(z)+C_{2}\quad(t+d)^{\alpha}Y_{\nu}(z)

with

z=β(t+d)γα=(1a)/2γ=1+(ba)/2β=σ/|γ|ν=|α/γ|.\displaystyle z=\beta(t+d)^{\gamma}\quad\alpha=(1-a)/2\quad\gamma=1+(b-a)/2\quad\beta=\sigma/|\gamma|\quad\nu=|\alpha/\gamma|.

where C1C_{1}, C2C_{2} are chosen so x(0)=1,x(0)=0x(0)=1,x^{\prime}(0)=0. However if γ=0\gamma=0, which occurs when a=b+2a=b+2 then in this case there are power law solutions

x(t)=C1(t+d)α+C2(t+d)α+\displaystyle x(t)=C_{1}(t+d)^{\alpha_{-}}+C_{2}(t+d)^{\alpha_{+}}

where

α±=(a1)±(a1)24σ22\displaystyle\alpha_{\pm}=\frac{-(a-1)\pm\sqrt{(a-1)^{2}-4\sigma^{2}}}{2}

unless a=1±2σa=1\pm 2\sigma.

Example 5 (Caldirola-Kanai oscillator [6, 22]).

The Caldirola-Kanai oscillator which has an exponentially increasing mass and frequency, has been studied through various methods is one of the most typical time-dependent quantum systems whose exact quantum states are known. Let κ1(t)=e2at2,κ2(t)=e2atσ22\kappa_{1}(t)=\frac{e^{-2at}}{2},\kappa_{2}(t)=\frac{e^{2at}\sigma^{2}}{2}, a>0a>0, corresponding to the Hamiltonian

H=|p|22e2at+σ2e2at|x|22\displaystyle H=\frac{|p|^{2}}{2e^{2at}}+\frac{\sigma^{2}e^{2at}|x|^{2}}{2}

then the solutions to the Hamiltonian flow are given by

x(t)=eatcos(Λt)+aΛeatsin(Λt)\displaystyle x(t)=e^{-at}\cos(\Lambda t)+\frac{a}{\Lambda}e^{-at}\sin(\Lambda t)
p(t)=a2Λeatsin(Λt)Λeatsin(Λt)\displaystyle p(t)=-\frac{a^{2}}{\Lambda}e^{at}\sin(\Lambda t)-\Lambda e^{at}\sin(\Lambda t)

where Λ=σ2a2\Lambda=\sqrt{\sigma^{2}-a^{2}}, with σ2a2>0\sigma^{2}-a^{2}>0. If Λ\Lambda real, then TD=π2ΛT_{D}=\frac{\pi}{2\Lambda}, for example. The conditions (6.10) and (2.10) are then also nonempty using trigonometric identities.

When Λ\Lambda is imaginary, we can recalculate. Let λ=a2σ2\lambda=\sqrt{a^{2}-\sigma^{2}} where a2σ2>0a^{2}-\sigma^{2}>0. Then

x(t)=eatcosh(λt)+aλeatsinh(λt)\displaystyle x(t)=e^{-at}\cosh(\lambda t)+\frac{a}{\lambda}e^{-at}\sinh(\lambda t) (7.7)
p(t)=a2λeatsinh(λt)+Λeatsinh(λt)\displaystyle p(t)=-\frac{a^{2}}{\lambda}e^{at}\sinh(\lambda t)+\Lambda e^{at}\sinh(\lambda t) (7.8)

The equations for the FIO coefficients are then

y1(t)=e2at2(λa2λ)sinh(λt)cosh(λt)+aλsinh(λt)\displaystyle y_{1}(t)=\frac{e^{2at}}{2}\frac{\left(\lambda-\frac{a^{2}}{\lambda}\right)\sinh(\lambda t)}{\cosh(\lambda t)+\frac{a}{\lambda}\sinh(\lambda t)}
y2(t)=λeatλcosh(λt)+asinh(λt)\displaystyle y_{2}(t)=\frac{\lambda e^{at}}{\lambda\cosh(\lambda t)+a\sinh(\lambda t)}
y3(t)=sinh(λt)2asinh(λt)+2λcosh(λt)\displaystyle y_{3}(t)=-\frac{\sinh(\lambda t)}{2a\sinh(\lambda t)+2\lambda\cosh(\lambda t)}

As you can see a(t)=y2(t)a(t)=\sqrt{y_{2}(t)}, and TD=T_{D}=\infty. In this case we have that

λe2ata(1+eat)|y2(t)|2e2at\displaystyle\frac{\lambda e^{2at}}{a(1+e^{at})}\leq|y_{2}(t)|\leq 2e^{2at} (7.9)
14a(eλt+1)|y3(t)|eλt2λ\displaystyle\frac{1}{4a(e^{\lambda t}+1)}\leq|y_{3}(t)|\leq\frac{e^{-\lambda t}}{2\lambda} (7.10)

and

A=λeat(asinh(λt)+λcosh(λt))2+4sinh2(λt)\displaystyle A=\frac{\lambda e^{at}}{\sqrt{(a\sinh(\lambda t)+\lambda\cosh(\lambda t))^{2}+4\sinh^{2}(\lambda t)}} (7.11)

so there exists constants C1,C2C_{1},C_{2} depending on aa and λ\lambda such that

C1eλtAC2eat\displaystyle C_{1}e^{-\lambda t}\leq A\leq C_{2}e^{at} (7.12)

the conditions (6.10) and (2.10) are true for a nonempty set of ϵ0\epsilon_{0} and R1R_{1} for finite TT.

Proof of Proposition 1.

The proof of Proposition 1 follows directly from the examples above. ∎

8. Conclusions

We have the following Corollary

Corollary 2.

The function ϕ(x)\phi(x) constructed in Lemma 3 shifted a distance from Ω\Omega specified in Theorem 4 with 2M=|αN|2M=|\alpha_{N}| is L2LL^{2}\textendash L^{\infty} observable if

η=(2eM24+MΔx)d<e14eϕL2(d).\displaystyle\eta=\left(2e^{-\frac{M^{2}}{4}}+M\Delta x\right)^{d}<\frac{e-1}{4e}{\|\phi\|_{L^{2}(\mathbb{R}^{d})}}. (8.1)
Proof.

We have from the proof of Theorem 4 that η\eta must satisfy the following

η=(2eM24+MΔx)dmin{(CTT),1}<14ϕL2(d)\eta=\left(2e^{-\frac{M^{2}}{4}}+M\Delta x\right)^{d}\min\{(C_{T}T),1\}<\frac{1}{4}{\|\phi\|_{L^{2}(\mathbb{R}^{d})}}

CT=e(e1)TC_{T}=\frac{e}{(e-1)T} to be L2LL^{2}\textendash L^{\infty} observable, that is for the η\eta to be folded into the left hand side of (2.9). By the construction of ϕ\phi, 2M=|αN|2M=|\alpha_{N}| which gives the distance R1R_{1} required by Theorem 4. ∎

More general ϕ𝒜\phi\in\mathcal{A} can also be constructed. Because the FIO construction presented in Lemma 1 here also gives an exact construction for initial data which consists of Hermite functions, there is room for future development and examples of the construction presented here.

References

  • [1] L. Baudouin and J.P. Puel, Uniqueness and stability in an inverse problem for the Schrödinger equation. Inverse problems. Vol. 18 No. 6, (2002) pp. 15-37.
  • [2] K. Beauchard, M.  Egidi, and K. Pravda-Starov, Geometric conditions for the null-controllability of hypoelliptic quadratic parabolic equations with moving control supports. accepted for publication in Comptes Rendus - Mathématique.
  • [3] K. Beauchard, P. Jaming, and K. Pravda-Starov, Spectral inequality for finite combinations of Hermite functions and null-controllability of hypoelliptic quadratic equations. preprint (2018), arXiv:1804.04895.
  • [4] J. Boyd, Asymptotic coefficients of Hermite function series. Journal of Computational Physics Vol. 54 (1984) pp 382-410.
  • [5] C. Calcutta, and A. Bolt, Approximating by Gaussians. https://arxiv.org/pdf/0805.3795.pdf
  • [6] P.  Caldirola, Forze non conservative nella meccanica quantistica. Il Nuovo Cimento I8, (1941) pp. 393.
  • [7] S.  Chang; Cosman, C. Pamela, and L. Milstein, Chernoff-Type Bounds for the Gaussian Error Function. IEEE Transactions on Communications. Vol. 59 No. 11. (2011) pp. 2939–2944.
  • [8] M. Combescure, and D. Robert, Semiclassical spreading of quantum wavepackets and applications near unstable fixed points of the classical flow. Asymptotic Analysis, Vol. 14 No. 4. (1997), pp. 377-404.
  • [9] M. Combescure, and D. Robert, Quadratic quantum Hamiltonians revisited. Cubo Vol. 8 No. 1 (2006), pp. 61-86.
  • [10] M. Combescure, and D. Robert, Coherent States and Applications in Mathematical Physics, Theoretical and Mathematical Physics. Springer, Dordrecht (2012).
  • [11] A. Cordoba, and C. Fefferman, Wave packets and Fourier integral operators. Comm. Part. Diff. Eq. Vol. 3, No. 11, (1978).
  • [12] J.-M. Coron, Control and nonlinearity, Mathematical Surveys and Monographs 136. AMS, Providence, RI (2007).
  • [13] J.  Diaz, and F.  Metcalf, A complementary triangle inequality in Hilbert and Banach spaces. Proceedings of the Amer. Math. Soc. Vol. 17 No. 1 (1966) pp 88-97.
  • [14] R. Grahm, D. Knuth, O. Patashnik. Concrete Mathematics. Addison-Wesley. (1994).
  • [15] A.  Grigis and J. Sjöstrand, Microlocal Analysis for Differential Operators, An Introduction. Cambridge University Press (1994).
  • [16] G.A. Hagedorn, Raising and lowering operators for semiclassical wave packets. Ann. Physics, Vol. 269, No. 1, (1998) pp. 77-104.
  • [17] N.R. Hill, Gaussian beam migration, Geophysics 55 (1990) 1416–1428.
  • [18] L. Hörmander, The Analysis of Linear Partial Differential Operators Vol III. Springer-Verlag (1983).
  • [19] L. Hörmander, Quadratic hyperbolic operators, Microlocal Analysis and Applications, Lecture Notes in Math. 1495, Eds. L. Cattabriga, L. Rodino, Springer (1991) pp. 118-160.
  • [20] L. Hörmander, Symplectic classification of quadratic forms and general Mehler formulas. Math. Z. Vol. 219, No. 3, (1995) pp. 413-449.
  • [21] S. Huang, G. Wang, and M. Wang, Observable sets, potentials and Schrödinger equations. arXiv preprint arXiv:2003.11263
  • [22] E. Kanai, On the Quantization of Dissipative Systems. Prog. Theor. Phys. Vol. 3 No. 440, (1948).
  • [23] S. Kim, A class of exactly solved time-dependent quantum harmonic oscillators. J. Phys. A: Math. Gen. Vol. 27 (1994) pp. 3927-3936.
  • [24] R. Killip, M. Visan, and X. Zhang, Quintic NLS in the exterior of a strictly convex obstacle. Amer. J. Math. Vol. 138, No. 5, (2016) pp. 1193-1346.
  • [25] A. Laptev, and I.M. Sigal, Global Fourier integral operators and semiclassical asymptotics, Rev. Math. Phys. Vol. 12, No. 5, (2000) pp. 749-766.
  • [26] I. Lasiecka and R. Triggiani Optimal regularity, exact controllability and uniform stabilization of Schrödinger equations with Dirichlet control. Differential Integral Equations, Vol. 5, No. 3, (1992) pp. 521–535.
  • [27] I.  Lasiecka, R. Triggiani, and X. Zhang, Nonconservative Schrödinger equations with unobserved Neumann B. C.: Global uniqueness and observability in one shot. Analysis and Optimization of Differential Systems (2002) pp. 235-246.
  • [28] I.  Lasiecka, R. Triggiani, and X. Zhang, Global uniqueness, observability and stabilization of nonconservative Schrödinger equations via pointwise Carleman estimates. J. Inv. Ill-Posed Problems, Vol. 11, No. 3 (2003). pp 1-96.
  • [29] G. Lebeau, Controle de l’equation de Schrödinger. J. Math. Pures Appl., Vol. 71 (1992), pp. 267-291.
  • [30] J.-L. Lions, Controlabilite exacte, perturbations et stabilisation de systemes distribues. Tome 1 et 2, Vol. 8, Recherches en Mathematiques Appliquees, Masson, Paris (1988).
  • [31] H. Liu, O. Runborg and N. M. Tanushev. Error estimates for Gaussian beam superpositions. Math. Comp. Vol 82 (2013) pp. 919-952.
  • [32] F. Macia, The Schrodinger flow on a compact manifold: High-frequency dynamics, and dispersion Modern Aspects of the Theory of Partial Differential Equations, Oper. Theory Adv. Apple., 216, Springer, Basel (2011) pp.275-289.
  • [33] F. Macia, and G. Riviere, Concentration and Non-Concentration for the Schrodinger Evolution on Zoll Manifolds Comm. Math. Phys. 345 (2016) 3, 1019-1054.
  • [34] F. Macia, and G. Riviere, Observability and quantum limits for the Schrodinger equation on the sphere arXiv:1702.02066 (2017).
  • [35] J. Martin and K. Pravda-Starov, Geometric conditions for the exact controllability of fractional free and harmonic Schrödinger equations. preprint: https://perso.univ-rennes1.fr/karel.pravda-starov/Articles/ExactControSchro.pdf.
  • [36] L. Miller, Resolvent conditions for the control of unitary groups and their approximations. J. Spectr. Theory Vol. 2 No. 1, (2012), pp. 1-55.
  • [37] K. Phung, Observability and control of Schrödinger equations. SIAM J. Control Optim. Vol. 21 No. 1, (2001), pp. 211–230.
  • [38] K. Pravda-Starov, Generalized Mehler formula for time-dependent non-selfadjoint quadratic operators and propagation of singularities. Math. Ann. Vol. 372, No. 3-4, (2018) pp. 1335-1382.
  • [39] K. Pravda-Starov, L. Rodino, and P. Wahlberg, Propagation of Gabor singularities for Schrödinger equations with quadratic Hamiltonians. Math. Nachr. Vol. 291 (2018), pp. 128-159.
  • [40] S. Huang, M.  Wang, G. Wang Observable sets, potentials, and Schrödinger equations preprint arXiv preprint arXiv:2003.11263. (2020).
  • [41] G. Wang, M. Wang, and Y. Zhang, Observability and unique continuation inequalities for the Schrödinger equation. JEMS Vol. 21, No. 11, (2019) pp. 3513–3572.
  • [42] G. Wang, M. Wang, C. Zhang, and Y. Zhang, Observable set, observability, interpolation inequality and spectral inequality for the heat equation in n\mathbb{R}^{n}. Journal de Mathematiques Pures et Appliquees Vol. 126 (2019) pp.144-194.
  • [43] N. Wiener, Tauberian Theorems. Annals of Mathematics, Vol. 33 No. 1 (1932) pp. 1–100.
  • [44] E.  Zuazua, Remarks on the controllability of the Schrödinger equation. Quantum Control: Mathematical and Numerical Challenges. CRM Proceedings and Lecture notes Vol. 23 (2003)