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

\stackMath

The Airy equation with nonlocal conditions

B. Normatov* and D. A. Smith*†
* Division of Science, Yale-NUS College, Singapore
Department of Mathematics, National University of Singapore Singapore,
[email protected]
Abstract

We study a third order dispersive linear evolution equation on the finite interval subject to an initial condition and inhomogeneous boundary conditions but, in place of one of the three boundary conditions that would typically be imposed, we use a nonlocal condition, which specifies a weighted integral of the solution over the spatial interval. Via adaptations of the Fokas transform method (or unified transform method), we obtain a solution representation for this problem. We also study the time periodic analogue of this problem, and thereby obtain long time asymptotics for the original problem with time periodic boundary and nonlocal data.

1 Introduction

We study the following third order initial nonlocal value problem (INVP).

Problem (Finite interval problem for the Airy equation with one nonlocal condition).
[t+x3]u(x,t)\displaystyle[\partial_{t}+\partial_{x}^{3}]u(x,t) =0\displaystyle=0 (x,t)\displaystyle(x,t) (0,1)×(0,),\displaystyle\in(0,1)\times(0,\infty), (\theparentequation.PDE)
u(x,0)\displaystyle u(x,0) =U(x)\displaystyle=U(x) x\displaystyle x [0,1],\displaystyle\in[0,1], (\theparentequation.IC)
u(1,t)\displaystyle u(1,t) =h0(t)\displaystyle=h_{0}(t) t\displaystyle t [0,),\displaystyle\in[0,\infty), (\theparentequation.BC0)
ux(1,t)\displaystyle u_{x}(1,t) =h1(t)\displaystyle=h_{1}(t) t\displaystyle t [0,),\displaystyle\in[0,\infty), (\theparentequation.BC1)
01K(y)u(y,t)dy\displaystyle\int_{0}^{1}K(y)u(y,t)\,\mathrm{d}y =h2(t)\displaystyle=h_{2}(t) t\displaystyle t [0,),\displaystyle\in[0,\infty), (\theparentequation.NC)
in which UU, hkh_{k}, and KK are sufficiently smooth and UU is compatible with hkh_{k} and KK in the sense that the boundary and nonlocal conditions hold when u(,t)u({}\cdot{},t) is replaced with UU on the left and the right sides are evaluated at t=0t=0.

We consider this problem to be of primarily theoretical interest in understanding the applicability of the Fokas transform method [9] to problems with nonlocal conditions. Indeed, problem (1.1) is a third order generalization of the INVP for the heat equation already studied by one of the authors [22], but this problem is of higher spatial order and is dispersive instead of dissipative.

The Fokas transform method (unified transform method) was developed by Fokas and collaborators in the late 1990s and early 2000s, initially as an inverse scattering transform method for integrable nonlinear equations on domains with a spatial boundary. However, it was soon understood that a version of the method for linear evolution equations yielded novel results, including the solution of the finite interval initial boundary value problem (IBVP) for the Airy equation (\theparentequation.PDE[25]. The method works by constructing an integral transform pair tailored specifically to the particular IBVP of interest. The derivation requires a complex contour deformation applied to the exponential Fourier inversion theorem, followed by the implementation in spectral space of a map from boundary data to unknown boundary values, a D to N map. See [5] for a good pedagogical introduction to the Fokas transform method, including its application to IBVP for equation (\theparentequation.PDE). The Fokas transform method owes its success to the diagonalization property of the transform pair. The transforms diagonalize the IBVP’s spatial differential operator in a weaker sense than does the Fourier sine transform diagonalize the half line Dirichlet heat operator, but in a sense that is suficient to enable the integral transform method to work. See [33] for a fuller explanation of the method’s applicability from the point of view of classical spectral theory.

In the field of linear evolution equations, the domain of applicability of the Fokas transform method has been extended to problems of higher spatial order [11], problems with general linear two point boundary conditions [31, 14, 26, 1], problems for systems of equations [12, 7, 16], and problems with mixed partial derivatives [6]. In the past decade, there has been interest in problems with more complex kinds of boundary conditions, such as interface problems [2, 3, 30, 28, 4], oblique Robin problems [21], multipoint problems [27], and general interface problems on networks [29, 1]. Some of this work reproduced earlier results, often with alternative solution representations more amenable to numerical evaluation, but many of the results were novel, especially for equations of third or higher spatial order. As multipoint conditions can represent a discretized weighted mean of uu over the spatial interval, it is natural to ask whether their continuous analogue, nonlocal conditions such as (\theparentequation.NC), may also be treated via the Fokas transform method. In [22], it was shown that this can indeed be done for the heat equation, but the question remained open for equations of higher spatial order or dispersive equations. This work addresses both of those cases.

Specifically for the dispersive third order INVP (1.1), we obtain three main results:

  1. (i)

    in theorem 3, a solution representation;

  2. (ii)

    in prop 16, long time decay of the solution where all boundary and nonlocal conditions are homogeneous;

  3. (iii)

    and, in theorem 17, long time asymptotics for the solution in the case that all boundary and nonlocal conditions are periodic.

In service of the first of these aims, §2 is dedicated to an implementation of the Fokas transform method, with adaptations appropriate to this problem. In §3, we study a problem related to problem (1.1) in which the data are time periodic with common period, but the initial condition is replaced with a time periodicity condition on uu. We use the “QQ equation” method [10, 15, 13] to obtain a solution representation and criteria on its validity. The greater part of §4 is dedicated to an asymptotic analysis of the solution representation for uu, whereby the long time decay result is derived. We then use the principle of linear superposition to decompose uu into a part satisfying the periodic problem and another part satisfying the homogeneous problem, obtaining the final theorem.

2 Solution of the Airy equation with nonlocal conditions

In this section, we derive an explicit integral representation of the solution of problem (1.1). We use the Fokas transform method, adapting the arguments presented in [22].

2.1 Global relations and Ehrenpreis form

Suppose that uu is a sufficiently smooth function satisfying the partial differential equation (\theparentequation.PDE). For j{0,1,2}j\in\{0,1,2\}, y[0,1]y\in[0,1], and λ\lambda\in\mathbb{C}, define

fj(λ;y,t)=0teiλ3sxju(y,s)ds.f_{j}(\lambda;y,t)=\int_{0}^{t}\mathrm{e}^{-\mathrm{i}\lambda^{3}s}\partial_{x}^{j}u(y,s)\,\mathrm{d}s. (2.1)

In the case that y{0,1}y\in\{0,1\}, fj(λ;y,t)f_{j}(\lambda;y,t) represents a time transform of a boundary value, which we refer to as a spectral boundary value. We use a hat ^\hat{}\cdot{} to signify the spatial exponential Fourier transform of function {}\cdot{} so that, for example,

u^(λ;t)=01eiλxu(x,t)dx.\hat{u}(\lambda;t)=\int_{0}^{1}\mathrm{e}^{-\mathrm{i}\lambda x}u(x,t)\,\mathrm{d}x.

Note that, in this Fourier transform, the zero extension of uu is taken beyond its spatial domain of definition. Fix 0y<z10\leqslant y<z\leqslant 1 and denote the exponential Fourier transform of the [y,z][y,z] restriction of a function by appending the arguments y,zy,z. For example,

u^(λ;t;y,z)=yzeiλxu(x,t)dx.\hat{u}(\lambda;t;y,z)=\int_{y}^{z}\mathrm{e}^{-\mathrm{i}\lambda x}u(x,t)\,\mathrm{d}x.

We apply the [y,z][y,z] restricted spatial Fourier transform to equation (\theparentequation.PDE) to obtain

0=\savestack\tmpbox\stretchto\scaleto\scalerel[[ut+uxxx]] 0.5ex\stackon[1pt][ut+uxxx]\tmpbox(λ;t;y,z).0=\savestack{\tmpbox}{\stretchto{\scaleto{\scalerel*[\widthof{\left[u_{t}+u_{xxx}\right]}]{\kern-0.6pt\bigwedge\kern-0.6pt}{\rule[-505.89pt]{4.30554pt}{505.89pt}}}{}}{0.5ex}}\stackon[1pt]{\left[u_{t}+u_{xxx}\right]}{\tmpbox}(\lambda;t;y,z).

By linearity of the Fourier transform, smoothness of uu in time, and spatial integration by parts, we obtain

0=[tiλ3]u^(λ;t;y,z)+eiλz[xxu(z,t)+iλxu(z,t)λ2u(z,t)]eiλy[xxu(y,t)+iλxu(y,t)λ2u(y,t)].0=\left[\partial_{t}-\mathrm{i}\lambda^{3}\right]\hat{u}(\lambda;t;y,z)+\mathrm{e}^{-\mathrm{i}\lambda z}\left[\partial_{xx}u(z,t)+\mathrm{i}\lambda\partial_{x}u(z,t)-\lambda^{2}u(z,t)\right]\\ -\mathrm{e}^{-\mathrm{i}\lambda y}\left[\partial_{xx}u(y,t)+\mathrm{i}\lambda\partial_{x}u(y,t)-\lambda^{2}u(y,t)\right]. (2.2)

Suppose further that uu satisfies the initial condition (\theparentequation.IC). Solving the temporal first order linear ordinary differential equation (2.2), we derive

eiλ3tu^(λ;t;y,z)=U^(λ;y,z)+eiλy[f2(λ;y,t)+iλf1(λ;y,t)λ2f0(λ;y,t)]eiλz[f2(λ;z,t)+iλf1(λ;z,t)λ2f0(λ;z,t)],\mathrm{e}^{-\mathrm{i}\lambda^{3}t}\hat{u}(\lambda;t;y,z)=\widehat{U}(\lambda;y,z)+\mathrm{e}^{-\mathrm{i}\lambda y}\left[f_{2}(\lambda;y,t)+\mathrm{i}\lambda f_{1}(\lambda;y,t)-\lambda^{2}f_{0}(\lambda;y,t)\right]\\ -\mathrm{e}^{-\mathrm{i}\lambda z}\left[f_{2}(\lambda;z,t)+\mathrm{i}\lambda f_{1}(\lambda;z,t)-\lambda^{2}f_{0}(\lambda;z,t)\right], (2.3)

an equation we refer to as the global relation on [y,z][y,z].

To the global relation on [0,1][0,1] we apply the inverse Fourier transform to find that

2πu(x,t)=eiλx+iλ3tU^(λ)dλ+eiλx+iλ3t[f2(λ;0,t)+iλf1(λ;0,t)λ2f0(λ;0,t)]dλeiλ(x1)+iλ3t[f2(λ;1,t)+iλf1(λ;1,t)λ2f0(λ;1,t)]dλ.2\pi u(x,t)=\int_{-\infty}^{\infty}\mathrm{e}^{\mathrm{i}\lambda x+\mathrm{i}\lambda^{3}t}\widehat{U}(\lambda)\,\mathrm{d}\lambda+\int_{-\infty}^{\infty}\mathrm{e}^{\mathrm{i}\lambda x+\mathrm{i}\lambda^{3}t}\left[f_{2}(\lambda;0,t)+\mathrm{i}\lambda f_{1}(\lambda;0,t)-\lambda^{2}f_{0}(\lambda;0,t)\right]\,\mathrm{d}\lambda\\ -\int_{-\infty}^{\infty}\mathrm{e}^{\mathrm{i}\lambda(x-1)+\mathrm{i}\lambda^{3}t}\left[f_{2}(\lambda;1,t)+\mathrm{i}\lambda f_{1}(\lambda;1,t)-\lambda^{2}f_{0}(\lambda;1,t)\right]\,\mathrm{d}\lambda. (2.4)

Note that the integrals in equation (2.4) are properly interpreted as Cauchy principal values, as they represent inverse Fourier transforms, and we cannot expect that UU or any of the other functions may be extended continuously by zero to \mathbb{R}. This Cauchy principal value interpretation of integrals shall be tacitly maintained for all formulae derived from equation (2.4).

We define the regions, each comprising a union of one or two open sectors,

D±={λ±:Re(iλ3)<0},E±={λ±:Re(iλ3)>0},D^{\pm}=\left\{\lambda\in\mathbb{C}^{\pm}:\operatorname*{Re}(-\mathrm{i}\lambda^{3})<0\right\},\qquad\qquad E^{\pm}=\left\{\lambda\in\mathbb{C}^{\pm}:\operatorname*{Re}(-\mathrm{i}\lambda^{3})>0\right\},

and adopt the convention that their boundaries, unions of rays eijπ/3[0,)\mathrm{e}^{\mathrm{i}j\pi/3}[0,\infty) for integer jj, are oriented so that the region lies to the left of the ray. We define further, for R>0R>0,

DR±={λD±:|λ|>R},ER±={λE±:|λ|>R},D^{\pm}_{R}=\left\{\lambda\in D^{\pm}:\left\lvert\lambda\right\rvert>R\right\},\qquad\qquad E^{\pm}_{R}=\left\{\lambda\in E^{\pm}:\left\lvert\lambda\right\rvert>R\right\}, (2.5)

also with their boundaries oriented so that the regions lie to the left. Integration by parts in the definition of fjf_{j} establishes that

eiλ3t[f2(λ;0,t)+iλf1(λ;0,t)λ2f0(λ;0,t)]=𝒪(λ1),\mathrm{e}^{\mathrm{i}\lambda^{3}t}\left[f_{2}(\lambda;0,t)+\mathrm{i}\lambda f_{1}(\lambda;0,t)-\lambda^{2}f_{0}(\lambda;0,t)\right]=\mathcal{O}\left(\lambda^{-1}\right),

uniformly in arg(λ)\arg(\lambda), as λ\lambda\to\infty within clos(E+)\operatorname{clos}(E^{+}). Hence, by Jordan’s lemma and Cauchy’s theorem,

E+eiλx+iλ3t[f2(λ;0,t)+iλf1(λ;0,t)λ2f0(λ;0,t)]dλ=0,\int_{\partial E^{+}}\mathrm{e}^{\mathrm{i}\lambda x+\mathrm{i}\lambda^{3}t}\left[f_{2}(\lambda;0,t)+\mathrm{i}\lambda f_{1}(\lambda;0,t)-\lambda^{2}f_{0}(\lambda;0,t)\right]\,\mathrm{d}\lambda=0,

and the second integral of equation (2.4) may have its contour deformed from the real line to DR+\partial D_{R}^{+}, for any R>0R>0. Similarly, the third integral of equation (2.4) may have its contour deformed from the real line to DR\partial D_{R}^{-}, but in the opposite direction. Therefore,

2πu(x,t)=eiλx+iλ3tU^(λ)dλ+DR+eiλx+iλ3t[f2(λ;0,t)+iλf1(λ;0,t)λ2f0(λ;0,t)]dλ+DReiλ(x1)+iλ3t[f2(λ;1,t)+iλf1(λ;1,t)λ2f0(λ;1,t)]dλ.2\pi u(x,t)=\int_{-\infty}^{\infty}\mathrm{e}^{\mathrm{i}\lambda x+\mathrm{i}\lambda^{3}t}\widehat{U}(\lambda)\,\mathrm{d}\lambda+\int_{\partial D_{R}^{+}}\mathrm{e}^{\mathrm{i}\lambda x+\mathrm{i}\lambda^{3}t}\left[f_{2}(\lambda;0,t)+\mathrm{i}\lambda f_{1}(\lambda;0,t)-\lambda^{2}f_{0}(\lambda;0,t)\right]\,\mathrm{d}\lambda\\ +\int_{\partial D_{R}^{-}}\mathrm{e}^{\mathrm{i}\lambda(x-1)+\mathrm{i}\lambda^{3}t}\left[f_{2}(\lambda;1,t)+\mathrm{i}\lambda f_{1}(\lambda;1,t)-\lambda^{2}f_{0}(\lambda;1,t)\right]\,\mathrm{d}\lambda. (2.6)

By a similar Jordan’s lemma argument, for all j{0,1,2}j\in\{0,1,2\} and all t>tt^{\prime}>t,

DR+eiλx+iλ3tλ2j[fj(λ;0,t)fj(λ;0,t)]dλ=0,\int_{\partial D_{R}^{+}}\mathrm{e}^{\mathrm{i}\lambda x+\mathrm{i}\lambda^{3}t}\lambda^{2-j}\Big{[}f_{j}(\lambda;0,t^{\prime})-f_{j}(\lambda;0,t)\Big{]}\,\mathrm{d}\lambda=0,

and similarly for the integral over DR\partial D_{R}^{-}. Therefore, equation (2.6) may alternatively be expressed with any ttt^{\prime}\geqslant t substituted for each tt appearing as an argument of an fjf_{j}, but keeping the tt in the exponential kernels unchanged:

2πu(x,t)=eiλx+iλ3tU^(λ)dλ+DR+eiλx+iλ3t[f2(λ;0,t)+iλf1(λ;0,t)λ2f0(λ;0,t)]dλ+DReiλ(x1)+iλ3t[f2(λ;1,t)+iλf1(λ;1,t)λ2f0(λ;1,t)]dλ,2\pi u(x,t)=\int_{-\infty}^{\infty}\mathrm{e}^{\mathrm{i}\lambda x+\mathrm{i}\lambda^{3}t}\widehat{U}(\lambda)\,\mathrm{d}\lambda\\ +\int_{\partial D_{R}^{+}}\mathrm{e}^{\mathrm{i}\lambda x+\mathrm{i}\lambda^{3}t}\left[f_{2}(\lambda;0,t^{\prime})+\mathrm{i}\lambda f_{1}(\lambda;0,t^{\prime})-\lambda^{2}f_{0}(\lambda;0,t^{\prime})\right]\,\mathrm{d}\lambda\\ +\int_{\partial D_{R}^{-}}\mathrm{e}^{\mathrm{i}\lambda(x-1)+\mathrm{i}\lambda^{3}t}\left[f_{2}(\lambda;1,t^{\prime})+\mathrm{i}\lambda f_{1}(\lambda;1,t^{\prime})-\lambda^{2}f_{0}(\lambda;1,t^{\prime})\right]\,\mathrm{d}\lambda, (2.7)

which is known as the Ehrenpreis form. This is particularly convenient for efficiency of numerical computation when studying problem (1.1) on a finite time interval, using tt^{\prime} the final time, but may be inappropriate for studying long time asymptotics of solutions because hkh_{k} need not be absolutely integrable on [0,)[0,\infty).

Thusfar, because we have not made use of any boundary or nonlocal conditions, the results obtained above are similar to those one requires in the study of IBVP for the Airy equation on the finite interval. The only difference is that the global relation (2.3) is presented for general subintervals of the spatial interval [0,1][0,1]. This slight generalization of the global relation will be crucial in the following arguments.

2.2 D to N map

Equation (2.7) is not an effective solution representation, because we do not have expressions for any of the six spectral boundary values fjf_{j}. Note that this issue appears even in the case of IBVP; any wellposed two point IBVP for the Airy equation must specify exactly three of the six fjf_{j} appearing in the Ehrenpreis form (2.7) or, more generally, exactly three linear combinations of the six. One must, whether studying an INVP or an IBVP construct a map from the data hkh_{k} to the presently unknown spectral boundary values fjf_{j}. The fact that problem (1.1) features a nonlocal condition in place of a boundary condition adds complexity to this D to N map, but is not the reason that such a map is required.

We seek expressions for each of the six spectral boundary values

f0(λ;0,t),f1(λ;0,t),f2(λ;0,t),f0(λ;1,t),f1(λ;1,t),f2(λ;1,t)f_{0}(\lambda;0,t^{\prime}),\qquad f_{1}(\lambda;0,t^{\prime}),\qquad f_{2}(\lambda;0,t^{\prime}),\qquad f_{0}(\lambda;1,t^{\prime}),\qquad f_{1}(\lambda;1,t^{\prime}),\qquad f_{2}(\lambda;1,t^{\prime}) (2.8)

in problem (1.1). We apply the time transform

ϕ(s)0teiλ3sϕ(s)ds\phi(s)\mapsto\int_{0}^{t^{\prime}}\mathrm{e}^{-\mathrm{i}\lambda^{3}s}\phi(s)\,\mathrm{d}s

that was used in equation (2.1) to the boundary conditions (\theparentequation.BC0)–(\theparentequation.BC1) to obtain expressions for two of these:

f0(λ;1,t)=0teiλ3sh0(s)ds=:h~0(λ;t),f1(λ;1,t)=0teiλ3sh1(s)ds=:h~1(λ;t).f_{0}(\lambda;1,t^{\prime})=\int_{0}^{t^{\prime}}\mathrm{e}^{-\mathrm{i}\lambda^{3}s}h_{0}(s)\,\mathrm{d}s=:\tilde{h}_{0}(\lambda;t^{\prime}),\qquad f_{1}(\lambda;1,t^{\prime})=\int_{0}^{t^{\prime}}\mathrm{e}^{-\mathrm{i}\lambda^{3}s}h_{1}(s)\,\mathrm{d}s=:\tilde{h}_{1}(\lambda;t^{\prime}). (2.9)

Let α=ei2π/3\alpha=\mathrm{e}^{\mathrm{i}2\pi/3}, a primitive cube root of unity. Observe that each of the functions fjf_{j} are symmetric under the transformations λαλ\lambda\mapsto\alpha^{\ell}\lambda for {0,1,2}\ell\in\{0,1,2\}. Using these substitutions in the global relation on [0,1][0,1] (2.3), we could obtain three equations relating the remaining four unknown spectral boundary values. If there were a third boundary condition in problem (1.1), then we might attempt to solve the resulting linear system for the remaining spectral boundary values, but there is no such third boundary condition, so this approach must be modified for the INVP. The nonlocal condition (\theparentequation.NC) must play a role in the construction of the D to N map; if it did not, then problem (1.1) could be solved without the nonlocal condition, which is false because problem (\theparentequation.PDE)–(\theparentequation.BC1) is an IBVP known to be underspecified hence illposed [25]. Guided by these cogitations, we adapt the nonlocal condition and global relation so that they feature some common terms before employing the aforementioned λαλ\lambda\mapsto\alpha^{\ell}\lambda symmetries.

We apply the same time transform from equation (2.1) to (\theparentequation.NC), yielding

01K(y)f0(λ;y,t)dy=0teiλ3sh2(s)ds=:h~2(λ;t).\int_{0}^{1}K(y)f_{0}(\lambda;y,t^{\prime})\,\mathrm{d}y=\int_{0}^{t^{\prime}}\mathrm{e}^{-\mathrm{i}\lambda^{3}s}h_{2}(s)\,\mathrm{d}s=:\tilde{h}_{2}(\lambda;t^{\prime}). (2.10)

Henceforth, for efficiency of presentation, we suppress the tt^{\prime} dependence of fjf_{j} and h~j\tilde{h}_{j}. Instead of using the global relation on [0,1][0,1], we use the global relation on [y,1][y,1] (2.3) at time tt^{\prime}, multiply each term by eiλyK(y)\mathrm{e}^{\mathrm{i}\lambda y}K(y), and integrate over y[0,1]y\in[0,1], obtaining

eiλK^(λ)f2(λ;1)iλ01K(y)f1(λ;y)dy01K(y)f2(λ;y)dy=N1(λ)eiλ3tν1(λ),\mathrm{e}^{-\mathrm{i}\lambda}\widehat{K}(-\lambda)f_{2}(\lambda;1)-\mathrm{i}\lambda\int_{0}^{1}K(y)f_{1}(\lambda;y)\,\mathrm{d}y-\int_{0}^{1}K(y)f_{2}(\lambda;y)\,\mathrm{d}y=N_{1}(\lambda)-\mathrm{e}^{-\mathrm{i}\lambda^{3}t^{\prime}}\nu_{1}(\lambda), (2.11)

which we refer to as the nonlocal global relation and in which

N1(λ)\displaystyle N_{1}(\lambda) =λ2eiλK^(λ)h~0(λ)iλeiλK^(λ)h~1(λ)λ2h~2(λ)+01K(y)eiλyU^(λ;y,1)dy,\displaystyle=\lambda^{2}\mathrm{e}^{-\mathrm{i}\lambda}\widehat{K}(-\lambda)\tilde{h}_{0}(\lambda)-\mathrm{i}\lambda\mathrm{e}^{-\mathrm{i}\lambda}\widehat{K}(-\lambda)\tilde{h}_{1}(\lambda)-\lambda^{2}\tilde{h}_{2}(\lambda)+\int_{0}^{1}K(y)\mathrm{e}^{\mathrm{i}\lambda y}\widehat{U}(\lambda;y,1)\,\mathrm{d}y,
ν1(λ)\displaystyle\nu_{1}(\lambda) =01K(y)eiλyu^(λ;t;y,1)dy\displaystyle=\int_{0}^{1}K(y)\mathrm{e}^{\mathrm{i}\lambda y}\hat{u}(\lambda;t^{\prime};y,1)\,\mathrm{d}y

also have their dependence on tt^{\prime} suppressed. Note that the terms in N1N_{1} are all expressed explicitly using the data of the problem, while ν1\nu_{1} involves u^\hat{u}, which is not a datum of the problem. We beg the reader to tolerate the slight notational inconvenience of carrying around these two terms instead of combining them because of the benefit in emphasizing the separation of data and nondata. Using the maps λαjλ\lambda\mapsto\alpha^{j}\lambda for j{0,1,2}j\in\{0,1,2\}, we obtain the linear system

(11eiλK^(λ)α1eiαλK^(αλ)α21eiα2λK^(α2λ))(iλ01K(y)f1(λ;y)dy01K(y)f2(λ;y)dyf2(λ;1))=(N1(λ)N1(αλ)N1(α2λ))eiλ3t(ν1(λ)ν1(αλ)ν1(α2λ)).\begin{pmatrix}1&1&\mathrm{e}^{-\mathrm{i}\lambda}\widehat{K}(-\lambda)\\ \alpha&1&\mathrm{e}^{-\mathrm{i}\alpha\lambda}\widehat{K}(-\alpha\lambda)\\ \alpha^{2}&1&\mathrm{e}^{-\mathrm{i}\alpha^{2}\lambda}\widehat{K}(-\alpha^{2}\lambda)\end{pmatrix}\begin{pmatrix}-\mathrm{i}\lambda\int_{0}^{1}K(y)f_{1}(\lambda;y)\,\mathrm{d}y\\ -\int_{0}^{1}K(y)f_{2}(\lambda;y)\,\mathrm{d}y\\ f_{2}(\lambda;1)\end{pmatrix}=\begin{pmatrix}N_{1}(\lambda)\\ N_{1}(\alpha\lambda)\\ N_{1}(\alpha^{2}\lambda)\end{pmatrix}-\mathrm{e}^{-\mathrm{i}\lambda^{3}t^{\prime}}\begin{pmatrix}\nu_{1}(\lambda)\\ \nu_{1}(\alpha\lambda)\\ \nu_{1}(\alpha^{2}\lambda)\end{pmatrix}.

We solve this system to obtain an expression for f2(λ;1)f_{2}(\lambda;1). We could also determine expressions for the other two entries in the vector of unknowns, but that is unnecessary because they do not appear in the Ehrenpreis form (2.7). Via Cramer’s rule, we find

f2(λ;1)=1Δ(λ)j=02αjN1(αjλ)eiλ3tΔ(λ)j=02αjν1(αjλ),f_{2}(\lambda;1)=\frac{1}{\Delta(\lambda)}\sum_{j=0}^{2}\alpha^{j}N_{1}(\alpha^{j}\lambda)-\frac{\mathrm{e}^{-\mathrm{i}\lambda^{3}t^{\prime}}}{\Delta(\lambda)}\sum_{j=0}^{2}\alpha^{j}\nu_{1}(\alpha^{j}\lambda),

in which the determinant of the system is, up to multiplication by i3\mathrm{i}\sqrt{3},

Δ(λ)=j=02αjeiαjλK^(αjλ).\Delta(\lambda)=\sum_{j=0}^{2}\alpha^{j}\mathrm{e}^{-\mathrm{i}\alpha^{j}\lambda}\widehat{K}(-\alpha^{j}\lambda).

We now have expressions for three of the six spectral boundary values, albeit with one depending on ν1\nu_{1}. The linear combination of spectral boundary values that appears in the integral along DR\partial D_{R}^{-} of the Ehrenpreis form (2.7) is

f2(λ;1)+iλf1(λ;1)λ2f0(λ;1)=1Δ(λ)j=02αjN1(αjλ)+iλh~1(λ;t)λ2h~0(λ;t)eiλ3tΔ(λ)j=02αjν1(αjλ).f_{2}(\lambda;1)+\mathrm{i}\lambda f_{1}(\lambda;1)-\lambda^{2}f_{0}(\lambda;1)=\frac{1}{\Delta(\lambda)}\sum_{j=0}^{2}\alpha^{j}N_{1}(\alpha^{j}\lambda)+\mathrm{i}\lambda\tilde{h}_{1}(\lambda;t)-\lambda^{2}\tilde{h}_{0}(\lambda;t)\\ -\frac{\mathrm{e}^{-\mathrm{i}\lambda^{3}t^{\prime}}}{\Delta(\lambda)}\sum_{j=0}^{2}\alpha^{j}\nu_{1}(\alpha^{j}\lambda). (2.12)

The global relation on [0,1][0,1] (2.3) is

f2(λ;0)+iλf1(λ;0)λ2f0(λ;0)=N0(λ)eiλ3tν0(λ),f_{2}(\lambda;0)+\mathrm{i}\lambda f_{1}(\lambda;0)-\lambda^{2}f_{0}(\lambda;0)=N_{0}(\lambda)-\mathrm{e}^{-\mathrm{i}\lambda^{3}t^{\prime}}\nu_{0}(\lambda), (2.13)

where

N0(λ)\displaystyle N_{0}(\lambda) =eiλ[1Δ(λ)j=02αjN1(αjλ)+iλh~1(λ)λ2h~0(λ)]U^(λ),\displaystyle=\mathrm{e}^{-\mathrm{i}\lambda}\left[\frac{1}{\Delta(\lambda)}\sum_{j=0}^{2}\alpha^{j}N_{1}(\alpha^{j}\lambda)+\mathrm{i}\lambda\tilde{h}_{1}(\lambda)-\lambda^{2}\tilde{h}_{0}(\lambda)\right]-\widehat{U}(\lambda),
ν0(λ)\displaystyle\nu_{0}(\lambda) =eiλΔ(λ)j=02αjν1(αjλ)u^(λ;t).\displaystyle=\frac{\mathrm{e}^{-\mathrm{i}\lambda}}{\Delta(\lambda)}\sum_{j=0}^{2}\alpha^{j}\nu_{1}(\alpha^{j}\lambda)-\hat{u}(\lambda;t^{\prime}).

As above, N0N_{0} contains the data of the problem and ν0\nu_{0} contains nondata, and both have their dependence on tt^{\prime} suppressed. Equation (2.13) provides precisely the linear combination of spectral boundary values that appears in the integral along DR+\partial D_{R}^{+} of the Ehrenpreis form.

2.3 Solution of the INVP

By substituting formulae (2.12) and (2.13) into the Ehrenpreis form (2.7), one obtains an expression for the solution u(x,t)u(x,t), but it depends on both the data N0,N1N_{0},N_{1} and the nondata ν0,ν1\nu_{0},\nu_{1}. We aim to show that the terms involving nondata contribute nothing to the solution. The main tools are the following lemmata.

Lemma 1.

Suppose that KK has bounded total variation and KK is continuous and nonzero at 0. There exists a finite R>0R>0 such that there are no zeros of Δ\Delta in clos(DR+DR)\operatorname{clos}(D_{R}^{+}\cup D_{R}^{-}).

Lemma 2.

Suppose that KK has bounded total variation and KK is continuous and nonzero at 0. For all tt>0t^{\prime}\geqslant t>0,

eiλ3(tt)ν0(λ;t)=𝒪(1),\mathrm{e}^{-\mathrm{i}\lambda^{3}(t^{\prime}-t)}\nu_{0}(\lambda;t^{\prime})=\mathcal{O}(1),

uniformly in arg(λ)\arg(\lambda), as λ\lambda\to\infty within clos(DR+)\operatorname{clos}(D^{+}_{R}). Similarly, for all tt>0t^{\prime}\geqslant t>0,

eiλ3(tt)1Δ(λ)j=02αjν1(αjλ)=𝒪(λ1),\mathrm{e}^{-\mathrm{i}\lambda^{3}(t^{\prime}-t)}\frac{1}{\Delta(\lambda)}\sum_{j=0}^{2}\alpha^{j}\nu_{1}(\alpha^{j}\lambda)=\mathcal{O}\left(\lambda^{-1}\right),

uniformly in arg(λ)\arg(\lambda), as λ\lambda\to\infty within clos(DR)\operatorname{clos}(D^{-}_{R}).

In the Ehrenpreis form (2.7), we make substitutions for the spectral boundary values using formulae (2.12) and (2.13), to obtain

2πu(x,t)=[data]DR+eiλx2iλ3(tt)ν0(λ;t)eiλx2dλDReiλ(x1)iλ3(tt)1Δ(λ)j=02αjν1(αjλ;t)dλ.2\pi u(x,t)=[\mbox{data}]-\int_{\partial D_{R}^{+}}\mathrm{e}^{\mathrm{i}\lambda\frac{x}{2}-\mathrm{i}\lambda^{3}(t^{\prime}-t)}\nu_{0}(\lambda;t^{\prime})\mathrm{e}^{\mathrm{i}\lambda\frac{x}{2}}\,\mathrm{d}\lambda\\ -\int_{\partial D_{R}^{-}}\mathrm{e}^{\mathrm{i}\lambda(x-1)-\mathrm{i}\lambda^{3}(t^{\prime}-t)}\frac{1}{\Delta(\lambda)}\sum_{j=0}^{2}\alpha^{j}\nu_{1}(\alpha^{j}\lambda;t^{\prime})\,\mathrm{d}\lambda. (2.14)

By lemmata 1 and 2, Jordan’s lemma, and Cauchy’s theorem, the two displayed integrals on the right of equation (2.14) both evaluate to zero. Note that, to justify the application of Jordan’s lemma in the first integral, we used

eiλx2=𝒪(exp[x|λ|32]),\mathrm{e}^{\mathrm{i}\lambda\frac{x}{2}}=\mathcal{O}\left(\exp\left[-x\left\lvert\lambda\right\rvert\frac{\sqrt{3}}{2}\right]\right),

uniformly in arg(λ)\arg(\lambda), as λ\lambda\to\infty within clos(DR+)\operatorname{clos}(D^{+}_{R}), and the other eiλx2\mathrm{e}^{\mathrm{i}\lambda\frac{x}{2}} factor is used as the kernel for Jordan’s lemma. This justifies the following theorem.

Theorem 3.

Suppose that problem (1.1) has solution u(x,t)u(x,t), that UU and hkh_{k} are piecewise continuous, that KK has bounded total variation, and that KK is continuous and nonzero at 0. Then, for all ttt^{\prime}\geqslant t,

2πu(x,t)=eiλx+iλ3tU^(λ)dλ+DR+eiλx+iλ3tN0(λ;t)dλ+DReiλ(x1)+iλ3t[1Δ(λ)j=02αjN1(αjλ;t)+iλh~1(λ;t)λ2h~0(λ;t)]dλ.2\pi u(x,t)=\int_{-\infty}^{\infty}\mathrm{e}^{\mathrm{i}\lambda x+\mathrm{i}\lambda^{3}t}\widehat{U}(\lambda)\,\mathrm{d}\lambda+\int_{\partial D_{R}^{+}}\mathrm{e}^{\mathrm{i}\lambda x+\mathrm{i}\lambda^{3}t}N_{0}(\lambda;t^{\prime})\,\mathrm{d}\lambda\\ +\int_{\partial D_{R}^{-}}\mathrm{e}^{\mathrm{i}\lambda(x-1)+\mathrm{i}\lambda^{3}t}\left[\frac{1}{\Delta(\lambda)}\sum_{j=0}^{2}\alpha^{j}N_{1}(\alpha^{j}\lambda;t^{\prime})+\mathrm{i}\lambda\tilde{h}_{1}(\lambda;t^{\prime})-\lambda^{2}\tilde{h}_{0}(\lambda;t^{\prime})\right]\,\mathrm{d}\lambda. (2.15)

By their definition (2.5), ER+ERE_{R}^{+}\cup E_{R}^{-} and DR+DRD_{R}^{+}\cup D_{R}^{-} have the same boundaries, except on the circle C(0,R)C(0,R), but the boundaries are oppositely oriented. Note that two of the four semiinfinite components of DR\partial D_{R}^{-} are oppositely oriented but coincident with two of the four semiinfinite components of ER+\partial E_{R}^{+}, and these lie along the real line. It follows immediately from the definition of N0N_{0} that, for all λ\lambda\in\mathbb{R},

U^(λ)=N0(λ;t)+eiλ[1Δ(λ)j=02αjN1(αjλ;t)+iλh~1(λ;t)λ2h~0(λ;t)].\widehat{U}(\lambda)=-N_{0}(\lambda;t^{\prime})+\mathrm{e}^{-\mathrm{i}\lambda}\left[\frac{1}{\Delta(\lambda)}\sum_{j=0}^{2}\alpha^{j}N_{1}(\alpha^{j}\lambda;t^{\prime})+\mathrm{i}\lambda\tilde{h}_{1}(\lambda;t^{\prime})-\lambda^{2}\tilde{h}_{0}(\lambda;t^{\prime})\right].

making this substitution in the first integral of equation (2.15) and perturbing the contour away from the real line around C(0,R)C(0,R), we arrive at the following corollary.

Corollary 4.

Under the criteria of theorem 3, for all ttt^{\prime}\geqslant t,

2πu(x,t)=ER+eiλx+iλ3tN0(λ;t)dλEReiλ(x1)+iλ3t[1Δ(λ)j=02αjN1(αjλ;t)+iλh~1(λ;t)λ2h~0(λ;t)]dλ.2\pi u(x,t)=-\int_{\partial E_{R}^{+}}\mathrm{e}^{\mathrm{i}\lambda x+\mathrm{i}\lambda^{3}t}N_{0}(\lambda;t^{\prime})\,\mathrm{d}\lambda\\ -\int_{\partial E_{R}^{-}}\mathrm{e}^{\mathrm{i}\lambda(x-1)+\mathrm{i}\lambda^{3}t}\left[\frac{1}{\Delta(\lambda)}\sum_{j=0}^{2}\alpha^{j}N_{1}(\alpha^{j}\lambda;t^{\prime})+\mathrm{i}\lambda\tilde{h}_{1}(\lambda;t^{\prime})-\lambda^{2}\tilde{h}_{0}(\lambda;t^{\prime})\right]\,\mathrm{d}\lambda. (2.16)
Proof of lemma 1.

This proof follows the arguments of [18]. For notational convenience, we define κ(y)=K(1y)\kappa(y)=K(1-y) so that κ^(λ)=eiλK^(λ)\hat{\kappa}(\lambda)=\mathrm{e}^{-\mathrm{i}\lambda}\widehat{K}(-\lambda).

It is immediate from the definition that Δ(αkλ)=αkΔ(λ)\Delta(\alpha^{k}\lambda)=\alpha^{-k}\Delta(\lambda), so the zeros of Δ\Delta are arranged symmetrically according to rotation by 2π/32\pi/3. Consider λ\lambda such that iλ=a+bi\mathrm{i}\lambda=a+b\mathrm{i} with a,b0a,b\gg 0. In this region, we will show that the term ακ^(αλ)\alpha\hat{\kappa}(\alpha\lambda) is nonzero and dominates the other two terms of Δ\Delta, from which it follows that Δ\Delta has no zeros in this region. The argument is very similar in the a,b0a,-b\gg 0 region, but with the α2κ^(α2λ)\alpha^{2}\hat{\kappa}(\alpha^{2}\lambda) dominant instead. Therefore, outside some disc B(0,R)B(0,R), the zeros of Δ\Delta are confined to semistrips of some finite width 2w2w about the rays iαj+-\mathrm{i}\alpha^{j}\mathbb{R}^{+}, which proves the lemma.

Suppose iλ=a+bi\mathrm{i}\lambda=a+b\mathrm{i} with a,b0a,b\gg 0 and insist 0<δ10<\delta\ll 1 so that κ\kappa is continuous on at least [1δ,1][1-\delta,1]. Then

ακ^(αλ)=iλ(κ(1)eiαλκ(1δ)eiαλ(1δ)1δ1eiαλydκ(y))+α01δeiαλyκ(y)dy.\alpha\hat{\kappa}(\alpha\lambda)=\frac{\mathrm{i}}{\lambda}\left(\kappa(1)\mathrm{e}^{-\mathrm{i}\alpha\lambda}-\kappa(1-\delta)\mathrm{e}^{-\mathrm{i}\alpha\lambda(1-\delta)}-\int_{1-\delta}^{1}\mathrm{e}^{-\mathrm{i}\alpha\lambda y}\,\mathrm{d}\kappa(y)\right)+\alpha\int_{0}^{1-\delta}\mathrm{e}^{-\mathrm{i}\alpha\lambda y}\kappa(y)\,\mathrm{d}y.

Hence

|ακ^(αλ)|ea+3b2a2+b2[κ(1)eδa+3b2κ(1δ)max1δy1<y21|κ(y2)κ(y1)|eδa+3b2a2+b2maxy[0,1δ]κ(y)].\left\lvert\alpha\hat{\kappa}(\alpha\lambda)\right\rvert\geqslant\frac{\mathrm{e}^{\frac{a+\sqrt{3}b}{2}}}{\sqrt{a^{2}+b^{2}}}\left[\kappa(1)-\mathrm{e}^{-\delta\frac{a+\sqrt{3}b}{2}}\kappa(1-\delta)-\max_{1-\delta\leqslant y_{1}<y_{2}\leqslant 1}\left\lvert\kappa(y_{2})-\kappa(y_{1})\right\rvert\right.\\ \left.-\mathrm{e}^{-\delta\frac{a+\sqrt{3}b}{2}}\sqrt{a^{2}+b^{2}}\max_{y\in[0,1-\delta]}\kappa(y)\right].

Because κ\kappa is of bounded variation, the second maximum exists. Because κ\kappa is continuous at 11, the first maximum approaches 0 as δ0\delta\to 0. Hence, there exists R,w>0R,w>0 such that, for all a>Ra>R and b>wb>w, there exists δ>0\delta>0 for which the κ(1)\kappa(1) term dominates the others:

|ακ^(αλ)|ea+3b2a2+b2[κ(1)2],\left\lvert\alpha\hat{\kappa}(\alpha\lambda)\right\rvert\geqslant\frac{\mathrm{e}^{\frac{a+\sqrt{3}b}{2}}}{\sqrt{a^{2}+b^{2}}}\left[\frac{\kappa(1)}{2}\right],

say. But

|Δ(λ)||ακ^(αλ)||κ^(λ)||α2κ^(α2λ)|,\left\lvert\Delta(\lambda)\right\rvert\geqslant\left\lvert\alpha\hat{\kappa}(\alpha\lambda)\right\rvert-\left\lvert\hat{\kappa}(\lambda)\right\rvert-\left\lvert\alpha^{2}\hat{\kappa}(\alpha^{2}\lambda)\right\rvert,

and the latter two terms are bounded by

maxy[0,1]|κ(y)|andmaxy[0,1]|κ(y)|max{ea3b2,1},\max_{y\in[0,1]}\left\lvert\kappa(y)\right\rvert\qquad\mbox{and}\qquad\max_{y\in[0,1]}\left\lvert\kappa(y)\right\rvert\max\left\{\mathrm{e}^{\frac{a-\sqrt{3}b}{2}},1\right\},

respectively. Hence, possibly after further increasing one or both of RR and ww, it must be that Δ(λ)0\Delta(\lambda)\neq 0. ∎

Proof of lemma 2.

Note that, if ttt^{\prime}\geqslant t and λclos(D±)\lambda\in\operatorname{clos}(D^{\pm}), then Re[iλ3(tt)]0\operatorname*{Re}[-\mathrm{i}\lambda^{3}(t^{\prime}-t)]\leqslant 0, so the exponential factors exp[iλ3(tt)]=𝒪(1)\exp[-\mathrm{i}\lambda^{3}(t^{\prime}-t)]=\mathcal{O}(1), uniformly in arg(λ)\arg(\lambda) and may be discounted.

Note that DR+DRD_{R}^{+}\cup D_{R}^{-} has three connected components, of which one is DR+D_{R}^{+} and the other two comprise DRD_{R}^{-}. We label these components with the subscript 1,2,31,2,3 counting anticlockwise from the positive real axis so that DR 1+=DR+D_{R\;1}^{+}=D_{R}^{+}. With κ\kappa as defined in the proof of lemma 1, using the criteria on KK to integrate by parts,

Δ(λ)=iλ[κ(1)eiλκ(1δ)eiλ(1δ)1δ1eiλydκ(y)]+01δeiλyκ(y)dy+j=12αj01eiαjλyκ(y)dy,\Delta(\lambda)=\frac{\mathrm{i}}{\lambda}\left[\kappa(1)\mathrm{e}^{-\mathrm{i}\lambda}-\kappa(1-\delta)\mathrm{e}^{-\mathrm{i}\lambda(1-\delta)}-\int_{1-\delta}^{1}\mathrm{e}^{-\mathrm{i}\lambda y}\,\mathrm{d}\kappa(y)\right]+\int_{0}^{1-\delta}\mathrm{e}^{-\mathrm{i}\lambda y}\kappa(y)\,\mathrm{d}y\\ +\sum_{j=1}^{2}\alpha^{j}\int_{0}^{1}\mathrm{e}^{-\mathrm{i}\alpha^{j}\lambda y}\kappa(y)\,\mathrm{d}y,

where 0<δ10<\delta\ll 1 is such that κ\kappa is continuous on at least [1δ,1][1-\delta,1]. Consider λ\lambda\to\infty from within clos(D+)\operatorname{clos}(D^{+}) in this expression. Both the integrals in the final sum are 𝒪(1)\mathcal{O}(1) and both the second term of the bracket and the integral from 0 to 1δ1-\delta are 𝒪(exp(iλ[1δ]))\mathcal{O}(\exp(-\mathrm{i}\lambda[1-\delta])). The third term in the bracket is bounded by

|eiλ|max1δy1<y21|κ(y2)κ(y1)|,\left\lvert\mathrm{e}^{-\mathrm{i}\lambda}\right\rvert\max_{1-\delta\leqslant y_{1}<y_{2}\leqslant 1}\left\lvert\kappa(y_{2})-\kappa(y_{1})\right\rvert,

and the maximum has limit zero as δ0\delta\to 0 because κ\kappa is continuous on [1δ,1][1-\delta,1]. Therefore, by fixing δ>0\delta>0 small enough, we can ensure that this maximum is no greater than κ(1)/2\kappa(1)/2. Then the leading order term has nonzero coefficient, and all other terms are relatively decaying. Using the rotational symmetry of Δ\Delta, we can obtain similar estimates for the behaviour of Δ\Delta in other sectors. Indeed,

Δ(λ)={Θ(eiλ/λ)as λ from within clos(DR+),Θ(eiα2λ/λ)as λ from within clos(DR 2),Θ(eiαλ/λ)as λ from within clos(DR 3).\Delta(\lambda)=\begin{cases}\Theta\left(\mathrm{e}^{-\mathrm{i}\lambda}/\lambda\right)&\mbox{as }\lambda\to\infty\mbox{ from within }\operatorname{clos}\left(D_{R}^{+}\right),\\ \Theta\left(\mathrm{e}^{-\mathrm{i}\alpha^{2}\lambda}/\lambda\right)&\mbox{as }\lambda\to\infty\mbox{ from within }\operatorname{clos}\left(D_{R\;2}^{-}\right),\\ \Theta\left(\mathrm{e}^{-\mathrm{i}\alpha\lambda}/\lambda\right)&\mbox{as }\lambda\to\infty\mbox{ from within }\operatorname{clos}\left(D_{R\;3}^{-}\right).\end{cases}

The dominant term in

j=02αjν1(αjλ)\sum_{j=0}^{2}\alpha^{j}\nu_{1}(\alpha^{j}\lambda) (2.17)

is the j=2j=2 term in clos(DR 2)\operatorname{clos}\left(D_{R\;2}^{-}\right) and the j=1j=1 term in clos(DR 3)\operatorname{clos}\left(D_{R\;3}^{-}\right). However, the symmetry of expression (2.17) means that we may instead (and more notationally conveniently so) show decay of only ν1(λ)/Δ(λ)\nu_{1}(\lambda)/\Delta(\lambda) as λ\lambda\to\infty from within clos(DR+)\operatorname{clos}\left(D_{R}^{+}\right) and thereby conclude the latter claim of the lemma.

Select any 0<δ10<\delta\ll 1 for which KK is continuous on [0,δ][0,\delta]. Integration by parts implies

ν1(λ)=1iλ[K(δ)eiλδu^(λ;t;δ,1)K(0)u^(λ;t;0,1)0δeiλyu^(λ;t;y,1)dK(y)+0δK(y)u(y;t)dy]+δ1K(y)eiλyu^(λ;t;y,1)dy.\nu_{1}(\lambda)=\frac{1}{\mathrm{i}\lambda}\bigg{[}K(\delta)\mathrm{e}^{\mathrm{i}\lambda\delta}\hat{u}(\lambda;t^{\prime};\delta,1)-K(0)\hat{u}(\lambda;t^{\prime};0,1)-\int_{0}^{\delta}\mathrm{e}^{\mathrm{i}\lambda y}\hat{u}(\lambda;t^{\prime};y,1)\,\mathrm{d}K(y)\\ +\int_{0}^{\delta}K(y)u(y;t^{\prime})\,\mathrm{d}y\bigg{]}+\int_{\delta}^{1}K(y)\mathrm{e}^{\mathrm{i}\lambda y}\hat{u}(\lambda;t^{\prime};y,1)\,\mathrm{d}y. (2.18)

Integrating by parts, for all y[0,1)y\in[0,1),

|u^(λ;t;y,1)|\displaystyle\left\lvert\hat{u}(\lambda;t^{\prime};y,1)\right\rvert =|iλ[eiλu(1,t)eiλyu(y,t)y1eiλzux(z,t)dz]|\displaystyle=\left\lvert\frac{\mathrm{i}}{\lambda}\left[\mathrm{e}^{-\mathrm{i}\lambda}u(1,t^{\prime})-\mathrm{e}^{-\mathrm{i}\lambda y}u(y,t^{\prime})-\int_{y}^{1}\mathrm{e}^{-\mathrm{i}\lambda z}u_{x}(z,t^{\prime})\,\mathrm{d}z\right]\right\rvert
1|λ|[|u(1,t)||eiλ|+|u(y,t)||eiλy|+(1y)|eiλ|maxz[y,1]|ux(z,t)|]\displaystyle\leqslant\frac{1}{\left\lvert\lambda\right\rvert}\left[\left\lvert u(1,t^{\prime})\right\rvert\left\lvert\mathrm{e}^{-\mathrm{i}\lambda}\right\rvert+\left\lvert u(y,t^{\prime})\right\rvert\left\lvert\mathrm{e}^{-\mathrm{i}\lambda y}\right\rvert+(1-y)\left\lvert\mathrm{e}^{-\mathrm{i}\lambda}\right\rvert\max_{z\in[y,1]}\left\lvert u_{x}(z,t^{\prime})\right\rvert\right]
M|eiλλ|,\displaystyle\leqslant M\left\lvert\frac{\mathrm{e}^{-\mathrm{i}\lambda}}{\lambda}\right\rvert, (2.19)

for some M>0M>0, which can be chosen uniformly in yy. This immediately implies that the first and second terms in the bracket on the right of equation (2.18) are 𝒪(λ1eiλ)\mathcal{O}(\lambda^{-1}\mathrm{e}^{-\mathrm{i}\lambda}). But also

|0δeiλyu^(λ;t;y,1)dK(y)|M|eiλλ|V0δ(K),\left\lvert\int_{0}^{\delta}\mathrm{e}^{\mathrm{i}\lambda y}\hat{u}(\lambda;t^{\prime};y,1)\,\mathrm{d}K(y)\right\rvert\leqslant M\left\lvert\frac{\mathrm{e}^{-\mathrm{i}\lambda}}{\lambda}\right\rvert V_{0}^{\delta}(K),

where V0δ(K)V_{0}^{\delta}(K) represents the total variation of KK over [0,δ][0,\delta], so the third term in the bracket is also 𝒪(λ1eiλ)\mathcal{O}(\lambda^{-1}\mathrm{e}^{-\mathrm{i}\lambda}). The fourth term in the bracket is independent of λ\lambda.111This term is known by equation (\theparentequation.NC) to be h2(t)h_{2}(t^{\prime}), but we need not use that fact here. The final term on the right of equation (2.18) is 𝒪(eiλ(1δ))\mathcal{O}(\mathrm{e}^{-\mathrm{i}\lambda(1-\delta)}). Hence, overall, ν1(λ)/Δ(λ)=𝒪(λ1)\nu_{1}(\lambda)/\Delta(\lambda)=\mathcal{O}(\lambda^{-1}).

It remains only to establish the first claim of the lemma, in which we study λ\lambda\to\infty within clos(DR+)\operatorname{clos}\left(D_{R}^{+}\right). Note that

eiλΔ(λ)01K(y)eiλydy=1+𝒪(λ1).\frac{\mathrm{e}^{-\mathrm{i}\lambda}}{\Delta(\lambda)}\int_{0}^{1}K(y)\mathrm{e}^{\mathrm{i}\lambda y}\,\mathrm{d}y=1+\mathcal{O}(\lambda^{-1}).

Therefore,

ν0(λ;t)\displaystyle\nu_{0}(\lambda;t^{\prime}) =eiλΔ(λ)(j=02αjν1(αjλ)01K(y)eiλyu^(λ;t)dy)+𝒪(λ1)\displaystyle=\frac{\mathrm{e}^{-\mathrm{i}\lambda}}{\Delta(\lambda)}\left(\sum_{j=0}^{2}\alpha^{j}\nu_{1}(\alpha^{j}\lambda)-\int_{0}^{1}K(y)\mathrm{e}^{\mathrm{i}\lambda y}\hat{u}(\lambda;t^{\prime})\,\mathrm{d}y\right)+\mathcal{O}\left(\lambda^{-1}\right)
=eiλΔ(λ)(j=12αj01K(y)eiαjλyu^(αjλ;t;y,1)dy\displaystyle=\frac{\mathrm{e}^{-\mathrm{i}\lambda}}{\Delta(\lambda)}\left(\sum_{j=1}^{2}\alpha^{j}\int_{0}^{1}K(y)\mathrm{e}^{\mathrm{i}\alpha^{j}\lambda y}\hat{u}(\alpha^{j}\lambda;t^{\prime};y,1)\,\mathrm{d}y\right.
01K(y)eiλyu^(λ;t;0,y)dy)+𝒪(λ1).\displaystyle\hskip 130.0002pt\left.\vphantom{\sum_{j=1}^{2}}-\int_{0}^{1}K(y)\mathrm{e}^{\mathrm{i}\lambda y}\hat{u}(\lambda;t^{\prime};0,y)\,\mathrm{d}y\right)+\mathcal{O}\left(\lambda^{-1}\right). (2.20)

Using similar calculations to those justifying inequalities (2.19), we find that

eiαjλyu^(αjλ;t;y,1)\displaystyle\mathrm{e}^{\mathrm{i}\alpha^{j}\lambda y}\hat{u}(\alpha^{j}\lambda;t^{\prime};y,1) =𝒪(λ1),\displaystyle=\mathcal{O}(\lambda^{-1}),
eiλyu^(αjλ;t;0,y)\displaystyle\mathrm{e}^{\mathrm{i}\lambda y}\hat{u}(\alpha^{j}\lambda;t^{\prime};0,y) =𝒪(λ1),\displaystyle=\mathcal{O}(\lambda^{-1}),

both uniformly in yy. Hence both integrals in equation (2.20) are 𝒪(λ1)\mathcal{O}(\lambda^{-1}), and ν0=𝒪(1)\nu_{0}=\mathcal{O}(1). ∎

Remark 5.

If we assume that KK is not just of bounded variation but continuously differentiable, then we can improve on the information lemma 2 provides about the behaviour of ν0\nu_{0}. Firstly, integrating by parts on the whole interval [0,1][0,1] instead of just [1δ,1][1-\delta,1], we find the leading order term in Δ\Delta explicitly:

Δ(λ)=iλK(0)eiλ+𝒪(λ2eiλ)as λ within clos(D+).\Delta(\lambda)=\frac{\mathrm{i}}{\lambda}K(0)\mathrm{e}^{-\mathrm{i}\lambda}+\mathcal{O}\left(\lambda^{-2}\mathrm{e}^{-\mathrm{i}\lambda}\right)\qquad\mbox{as }\lambda\to\infty\mbox{ within }\operatorname{clos}(D^{+}).

We proceed from equation (2.20). Integrating by parts, we find that, for j=1,2j=1,2,

01K(y)eiαjλyy1eiαjλxu(x,t)dxdy=1iαjλ(K(0)01eiαjλxu(x,t)dx+01K(y)u(y,t)dy01eiαjλyK(y)y1eiαjλxu(x,t)dxdy)\int_{0}^{1}K(y)\mathrm{e}^{\mathrm{i}\alpha^{j}\lambda y}\int_{y}^{1}\mathrm{e}^{-\mathrm{i}\alpha^{j}\lambda x}u(x,t^{\prime})\,\mathrm{d}x\,\mathrm{d}y=\frac{1}{\mathrm{i}\alpha^{j}\lambda}\left(-K(0)\int_{0}^{1}\mathrm{e}^{-\mathrm{i}\alpha^{j}\lambda x}u(x,t^{\prime})\,\mathrm{d}x\right.\\ +\left.\int_{0}^{1}K(y)u(y,t^{\prime})\,\mathrm{d}y-\int_{0}^{1}\mathrm{e}^{\mathrm{i}\alpha^{j}\lambda y}K^{\prime}(y)\int_{y}^{1}\mathrm{e}^{-\mathrm{i}\alpha^{j}\lambda x}u(x,t^{\prime})\,\mathrm{d}x\,\mathrm{d}y\right)

The first and third integrals on the right are both 𝒪(λ1)\mathcal{O}\left(\lambda^{-1}\right). The other is h2(t)h_{2}(t^{\prime}) by equation (\theparentequation.NC). Therefore, equation (2.20) simplifies to

ν0(λ;t)=eiλΔ(λ)(2h2(t)iλ01K(y)eiλy0yeiλxu(x,t)dxdy)+𝒪(λ1).\nu_{0}(\lambda;t^{\prime})=\frac{\mathrm{e}^{-\mathrm{i}\lambda}}{\Delta(\lambda)}\left(\frac{2h_{2}(t^{\prime})}{\mathrm{i}\lambda}-\int_{0}^{1}K(y)\mathrm{e}^{\mathrm{i}\lambda y}\int_{0}^{y}\mathrm{e}^{-\mathrm{i}\lambda x}u(x,t^{\prime})\,\mathrm{d}x\,\mathrm{d}y\right)+\mathcal{O}\left(\lambda^{-1}\right). (2.21)

To obtain the leading order behaviour of the remaining integral in equation (2.21), we again integrate by parts:

01K(y)eiλy0yeiλxu(x,t)dxdy=1iλ(K(1)eiλ01eiλxu(x,t)dx01K(y)u(y,t)dy01eiλyK(y)0yeiλxu(x,t)dxdy).\int_{0}^{1}K(y)\mathrm{e}^{\mathrm{i}\lambda y}\int_{0}^{y}\mathrm{e}^{-\mathrm{i}\lambda x}u(x,t^{\prime})\,\mathrm{d}x\,\mathrm{d}y=\frac{1}{\mathrm{i}\lambda}\left(K(1)\mathrm{e}^{\mathrm{i}\lambda}\int_{0}^{1}\mathrm{e}^{-\mathrm{i}\lambda x}u(x,t^{\prime})\,\mathrm{d}x-\int_{0}^{1}K(y)u(y,t^{\prime})\,\mathrm{d}y\right.\\ -\left.\int_{0}^{1}\mathrm{e}^{\mathrm{i}\lambda y}K^{\prime}(y)\int_{0}^{y}\mathrm{e}^{-\mathrm{i}\lambda x}u(x,t^{\prime})\,\mathrm{d}x\,\mathrm{d}y\right).

The first and third parenthetical terms are both 𝒪(λ1)\mathcal{O}\left(\lambda^{-1}\right) and the second term is known. Substituting into equation (2.21), we find that

ν0(λ;t)=eiλΔ(λ)iλ3h2(t)+𝒪(λ1)=3h2(t)K(0)+𝒪(λ1).\nu_{0}(\lambda;t^{\prime})=\frac{\mathrm{e}^{-\mathrm{i}\lambda}}{\Delta(\lambda)\mathrm{i}\lambda}3h_{2}(t^{\prime})+\mathcal{O}\left(\lambda^{-1}\right)=\frac{-3h_{2}(t^{\prime})}{K(0)}+\mathcal{O}\left(\lambda^{-1}\right).
Remark 6.

The D to N map arguments used, in conjunction with an evaluation of the global relation at [0,1][0,1], an evaluation z=1z=1 and integration over y[0,1]y\in[0,1]. This was arbitrary, in the sense that one could have instead employed an evaluation at y=0y=0 and integration over z[0,1]z\in[0,1]. We emphasize that the three equations derived from the global relation thus form a system of rank 22. However, problem (1.1) admits a simpler D to N map when using our selection of evaluations, because it makes the D to N map separate into a system of three equations instead of the full six.

Remark 7.

It is reasonable to attempt to extend the above presented arguments to a problem with no boundary conditions but three nonlocal conditions, each with a different weight K0,K1,K2K_{0},K_{1},K_{2}. Unfortunately, this is rather difficult, because the direct analogue of lemma 2, appears to be false. Indeed, to obtain formulae for each of f0(λ;1),f1(λ;1),f2(λ;1)f_{0}(\lambda;1),f_{1}(\lambda;1),f_{2}(\lambda;1), requires using versions of the nonlocal global relation (2.11) with each weight KkK_{k}, introducing six new unknowns instead of two. Therefore, the λαjλ\lambda\mapsto\alpha^{j}\lambda maps are required for each nonlocal global relation, resulting in a full rank system of nine equations in nine unknowns. When solving this system via Cramer’s rule, one finds exponentials e2iαjλ\mathrm{e}^{-2\mathrm{i}\alpha^{j}\lambda} in the numerators, but only eiαjλ\mathrm{e}^{-\mathrm{i}\alpha^{j}\lambda} in the denominator, so the ratios are unbounded on the relevant sectors. It is an open question whether problems like this are illposed, wellposed but not amenable to a Fokas transform method approach, or open to anlaysis via an alternative version of the Fokas transfrom method. For example, it may be possible to adapt the method, outlined in [22, §5.1–5.2], of understanding nonlocal value problems as weak-\star limits of multipoint value problems, but then uniqueness of the solution must be proved by other means.

Remark 8.

In [32, 14, 26, 1, 33], it was shown how to understand and even construct the Fokas transform method for IBVP, and in particular the objects analagous to N0,N1N_{0},N_{1}, in terms of the characteristic matrix of the classical (Lagrange) adjoint of the spatial differential operator. Because the spatial differential we study in this paper has a nonlocal condition, it does not have a classical adjoint, so it is not clear what the analagous construction should be. More examples like this one, or an analysis via weak-\star limits of multipoint operators (which do have classical adjoints [19, 1]) may be illustrative.

Remark 9.

Because the Airy equation is third order, it behaves substantially differently if time (equivalently, space) is run in the other direction. Indeed, it is expected that simply replacing equation (\theparentequation.PDE) with [tx3]u(x,t)=0[\partial_{t}-\partial_{x}^{3}]u(x,t)=0 would make INVP (1.1) illposed. It is known that separated IBVP for the Airy equation require exactly one boundary condition specified at the left and two at the right in order to be wellposed, and the alternative must have two at the left and one at the right [24]. In this context, the requirement in theorem 3 that KK be nonzero at 0 is not simply a technical imposition inherited from its lemmata but a fundamental requirement for wellposedness. For an INVP with the alternative PDE to be wellposed, it is expected that not only would the KK nonzero at 0 requirement remain, but one of the boundary conditions at x=1x=1 would have to be substituted for a further boundary condition at x=0x=0. A full characterization of wellposedness for third order IBVP including those with boundary conditions coupling between the two ends has still not been obtained, despite indications that it is related to the criteria for Birkhoff regularity [20, 31, 1]. Therefore, we relegate also to later work an investigation of how wellposedness of INVP for the Airy equation is affected by selection of different boundary and nonlocal conditions.

Remark 10.

This problem can be understood as related to that of a linearization of the physical problem of unidirectional waves in shallow water, where the surface elevation is difficult to measure at a point x=0x=0 but relatively easy to measure on average over an interval, using a measuring device with sensitivity K(x)K(x). This measurement may be related to that obtained from a pressure plate. In the small amplitude linearization of the Korteweg de Vries equation, equation (\theparentequation.PDE) would also have a uxu_{x} term. Including that term would make the formulae above more complicated but, guided by the results for IBVP, we do not expect its inclusion would significantly change the character of the results.

Related problems have KK the sum of a function of bounded variation and a one sided delta distribution or derivative delta distribution at x=0x=0. This effectively transforms the nonlocal condition into a hybrid boundary nonlocal condition. Such problems have applications in the feedback stabilizability and boundary controllability of the system via backstepping; see [17] for a general survey and, including the additional uxu_{x} term in the PDE, [23]. We expect that such problems may be studied via the means presented above, but the analysis in the proofs of results analogous to lemmata 1 and 2 would be slightly simplified because the presence of the xku(0,t)\partial_{x}^{k}u(0,t) term in the hybrid nonlocal boundary condition can provide an extra 𝒪(λ1)\mathcal{O}\left(\lambda^{-1}\right) decay in some of the numerator integrals.

Remark 11.

Having presented the full unified transform method for INVP (1.1), some discussion is warranted of how this relates to the method for the heat equation presented in [22].

To emphasize the ready adaptability of the method, we have been careful to present the broader argument in section 2 in a manner closely paralleling the heat equation paper. The arguments to justify lemmata 1 and 2 require some more careful bounds for the third order problem. This arises from the geometric complications of the exponential sum k=01j=02exp((1)kiαjλ)\sum_{k=0}^{1}\sum_{j=0}^{2}\exp((-1)^{k}\mathrm{i}\alpha^{j}\lambda) having 6 terms which may be dominant as λ\lambda\to\infty in various sectors, while the corresponding k=01j=01exp((1)kiα2jλ)\sum_{k=0}^{1}\sum_{j=0}^{1}\exp((-1)^{k}\mathrm{i}\alpha_{2}^{j}\lambda) (in which α2=1\alpha_{2}=-1 is a primitive square root of unity) arising for the heat equation has only two. But this point of contrast is one of detail rather than representing a fundamental divergence of the argument.

The most striking point is one of comparison rather than contrast. In lemma 1, the first asymptotic statement is of 𝒪(1)\mathcal{O}(1) behaviour rather than 𝒪(λ1)\mathcal{O}\left(\lambda^{-1}\right), and this is reflected in equation (2.3) of [22]. These two asymptotic bounds are very different to the 𝒪(λ1)\mathcal{O}\left(\lambda^{-1}\right) bounds one typically obtains in all relevant sectors in the equivalent lemmata for IBVP. They necessitate the more careful application of Jordan’s lemma in which the exp(iλx)\exp(\mathrm{i}\lambda x) factor is split, which works if and only if the relevant sector boundaries are nonparallel with the real line. It remains to be investigated whether this is a fundamental feature of the unified transform method for INVP, and how it might affect wellposedness for problems with different boundary and nonlocal problems.

3 Time periodic problems for the Airy equation

In this section, we study the time periodic analogue of problem (1.1). We modify the arguments of [15, 13] for the present setting, informed by the nonlocal global relation concepts that were introduced in [22] and adapted to the Airy equation above.

3.1 Time periodic problem

For T>0T>0, we say that function d:[0,)d:[0,\infty)\to\mathbb{C} is TT periodic if it can be represented via its exponential Fourier series

d(t)=nDneinωt,ω=2πT,d(t)=\sum_{n\in\mathbb{Z}}D_{n}\mathrm{e}^{\mathrm{i}n\omega t},\qquad\omega=\frac{2\pi}{T},

almost everywhere. Occasionally, we tacitly extend the domain of dd to \mathbb{R} so that it is periodic in the usual sense. This is not intended to suggest that all of the problems we study have solutions for negative time; it is merely a notational convenience.

We shall study the following problem, which is the time periodic analogue of problem (1.1), because the initial condition has been replaced with a time periodicity assumption.

Problem (Time periodic finite interval problem for the Airy equation with one nonlocal condition).
[t+x3]q(x,t)\displaystyle[\partial_{t}+\partial_{x}^{3}]q(x,t) =0\displaystyle=0 (x,t)\displaystyle(x,t) (0,1)×(0,),\displaystyle\in(0,1)\times(0,\infty), (\theparentequation.PDE)
q(x,t)\displaystyle q(x,t) =q(x,t+T)\displaystyle=q(x,t+T) (x,t)\displaystyle(x,t) [0,1]×[0,),\displaystyle\in[0,1]\times[0,\infty), (\theparentequation.Per)
q(1,t)\displaystyle q(1,t) =h0(t)\displaystyle=h_{0}(t) t\displaystyle t [0,),\displaystyle\in[0,\infty), (\theparentequation.BC0)
qx(1,t)\displaystyle q_{x}(1,t) =h1(t)\displaystyle=h_{1}(t) t\displaystyle t [0,),\displaystyle\in[0,\infty), (\theparentequation.BC1)
01K(y)q(y,t)dy\displaystyle\int_{0}^{1}K(y)q(y,t)\,\mathrm{d}y =h2(t)\displaystyle=h_{2}(t) t\displaystyle t [0,),\displaystyle\in[0,\infty), (\theparentequation.NC)
in which hkh_{k}, and KK are sufficiently smooth, hk(t)=hk(t+T)h_{k}(t)=h_{k}(t+T) for all t[0,)t\in[0,\infty), and T>0T>0 is fixed.

In [13, proposition 2], the asymptotically valid D to N map was derived for an asymptotically time periodic problem similar to (1.1). The differences being that the PDE was allowed to have any linear spatial differential operator, but only boundary conditions were admitted, not nonlocal conditions. Here, we specialise to the Airy equation, but admit the more complicated nonlocal condition (\theparentequation.NC). More significantly, we streamline the argument by focussing directly on the problem for qq, which is assumed to be truly periodic, not just asymptotically so.

3.2 The periodic relations

Because qq satisfies the same PDE as did uu in §2, qq also satisfies equation (2.2), provided 0yz10\leqslant y\leqslant z\leqslant 1. When evaluated at y=0y=0, z=1z=1, this equation is named “the QQ equation” in [15], after a change of notation Q(λ)=q^(λ;t)Q(\lambda)=-\hat{q}(\lambda;t). To emphasize the simpler derivation of this equation presented herein (and to preserve the symbol QQ for later use), we do not adopt this notation. To extend the arguments of [13] to nonlocal problems, we shall employ both the [0,1][0,1] and the [y,1]\int[y,1] versions of equation (2.2):

[tiλ3]q^(λ;t)\displaystyle\left[\partial_{t}-\mathrm{i}\lambda^{3}\right]\hat{q}(\lambda;t) =[xxq(0,t)+iλxq(0,t)λ2q(0,t)]\displaystyle=\left[\partial_{xx}q(0,t)+\mathrm{i}\lambda\partial_{x}q(0,t)-\lambda^{2}q(0,t)\right]
eiλ[xxq(1,t)+iλxq(1,t)λ2q(1,t)],\displaystyle\hskip 40.00006pt-\mathrm{e}^{-\mathrm{i}\lambda}\left[\partial_{xx}q(1,t)+\mathrm{i}\lambda\partial_{x}q(1,t)-\lambda^{2}q(1,t)\right], (3.2)
[tiλ3]01eiλyq^(λ;t;y,1)K(y)dy\displaystyle\left[\partial_{t}-\mathrm{i}\lambda^{3}\right]\int_{0}^{1}\mathrm{e}^{\mathrm{i}\lambda y}\hat{q}(\lambda;t;y,1)K(y)\,\mathrm{d}y =01[xxq(y,t)+iλxq(y,t)λ2q(y,t)]K(y)dy\displaystyle=\int_{0}^{1}\left[\partial_{xx}q(y,t)+\mathrm{i}\lambda\partial_{x}q(y,t)-\lambda^{2}q(y,t)\right]K(y)\,\mathrm{d}y
01eiλ(1y)K(y)dy[xxq(1,t)+iλxq(1,t)λ2q(1,t)],\displaystyle-\int_{0}^{1}\mathrm{e}^{-\mathrm{i}\lambda(1-y)}K(y)\,\mathrm{d}y\left[\partial_{xx}q(1,t)+\mathrm{i}\lambda\partial_{x}q(1,t)-\lambda^{2}q(1,t)\right], (3.3)

mirroring the approach of §2, wherein both [0,1][0,1] and [y,1]\int[y,1] versions of the global relation (2.3) were used.

Because qq is assumed time periodic, so also are its spatial Fourier transform and its boundary values. Hence, we may express all terms in the above equations using their exponential Fourier series over the interval [0,T][0,T].

nGnjeinωt=gj(t)=xjq(0,t),nΓnjeinωt=γj(t)=xjq(1,t),nHnkeinωt=hk(t),nQn(λ)einωt=q^(λ;t),nPn(λ)einωt=p(λ;t)=01eiλyK(y)q^(λ;t,y,1)dy,nAnjeinωt=ajk(t)=01K(y)xjq(y,t)dy,\begin{gathered}\sum_{n\in\mathbb{Z}}G_{n}^{j}\mathrm{e}^{\mathrm{i}n\omega t}=g^{j}(t)=\partial_{x}^{j}q(0,t),\qquad\qquad\sum_{n\in\mathbb{Z}}\Gamma_{n}^{j}\mathrm{e}^{\mathrm{i}n\omega t}=\gamma^{j}(t)=\partial_{x}^{j}q(1,t),\\ \sum_{n\in\mathbb{Z}}H_{n}^{k}\mathrm{e}^{\mathrm{i}n\omega t}=h_{k}(t),\qquad\qquad\sum_{n\in\mathbb{Z}}Q_{n}(\lambda)\mathrm{e}^{\mathrm{i}n\omega t}=\hat{q}(\lambda;t),\\ \sum_{n\in\mathbb{Z}}P_{n}(\lambda)\mathrm{e}^{\mathrm{i}n\omega t}=p(\lambda;t)=\int_{0}^{1}\mathrm{e}^{\mathrm{i}\lambda y}K(y)\hat{q}(\lambda;t,y,1)\,\mathrm{d}y,\\ \sum_{n\in\mathbb{Z}}A_{n}^{j}\mathrm{e}^{\mathrm{i}n\omega t}=a_{\;}^{j\;k}(t)=\int_{0}^{1}K(y)\partial_{x}^{j}q(y,t)\,\mathrm{d}y,\end{gathered} (3.4)

where each Fourier coefficient may be calculated as

Φn=1T0Tϕ(t)einωtdt.\Phi_{n}=\frac{1}{T}\int_{0}^{T}\phi(t)\mathrm{e}^{\mathrm{i}n\omega t}\,\mathrm{d}t.

In the above notation, ω=2π/T\omega=2\pi/T is the angular frequency, gg are the left boundary values, γ\gamma the right boundary values, hh the boundary data, and aa are nonlocal values. In problem (3.1), a0a^{0} is specified via nonlocal condition (\theparentequation.NC), while a1a^{1}, a2a^{2} are unknown. By the Fourier basis property, equality of Fourier series is equivalent to equality of corresponding Fourier coefficients. Therefore, equations (3.2) and (3.3) simplify to, for all nn\in\mathbb{Z} and all λ\lambda\in\mathbb{C},

(inωiλ3)Qn(λ)\displaystyle(\mathrm{i}n\omega-\mathrm{i}\lambda^{3})Q_{n}(\lambda) =[Gn2+iλGn1λ2Gn0]eiλ[Γn2+iλΓn1λ2Γn0],\displaystyle=\left[G_{n}^{2}+\mathrm{i}\lambda G_{n}^{1}-\lambda^{2}G_{n}^{0}\right]-\mathrm{e}^{-\mathrm{i}\lambda}\left[\Gamma_{n}^{2}+\mathrm{i}\lambda\Gamma_{n}^{1}-\lambda^{2}\Gamma_{n}^{0}\right], (3.5)
(inωiλ3)Pn(λ)\displaystyle(\mathrm{i}n\omega-\mathrm{i}\lambda^{3})P_{n}(\lambda) =[An2+iλAn1λ2An0]eiλK^(λ)[Γn2+iλΓn1λ2Γn0].\displaystyle=\left[A_{n}^{2}+\mathrm{i}\lambda A_{n}^{1}-\lambda^{2}A_{n}^{0}\right]-\mathrm{e}^{-\mathrm{i}\lambda}\widehat{K}(-\lambda)\left[\Gamma_{n}^{2}+\mathrm{i}\lambda\Gamma_{n}^{1}-\lambda^{2}\Gamma_{n}^{0}\right]. (3.6)

We refer to these equations as the periodic relation and nonlocal periodic relation, respectively. Other than the notational changes, in the derivation of these relations from equation (2.2), the only step was to calculate exponential Fourier coefficients.

Remark 12.

To aid solution of the INVP in §2, from equation (2.2) was derived the global relation on [0,1][0,1] (2.3) by solution of an ODE in time. The nonlocal global relation (2.11) followed by taking the same integrals against KK. Thus, the (nonlocal) periodic relation is the analogue of the (nonlocal) global relation, but derived via temporal Fourier series expansion instead of solution of a temporal ODE. This is a reasonable approach because, lacking an initial condition in problem (3.1), it is impossible to solve the temporal ODE, but periodicity means that temporal Fourier expansion is possible. Moreover, the fact that the global relations could be used to construct D to N maps for the INVP suggests the periodic relations may be valuable in finding D to N maps for the periodic nonlocal value problems.

3.3 Periodic D to N map

We construct the D to N map from the periodic relation (3.5) and the nonlocal periodic relation (3.6). Along with the boundary values GnkG_{n}^{k} and Γnk\Gamma_{n}^{k}, these equations feature, within QnQ_{n} and PnP_{n}, the unknown spatial Fourier transforms of qq. Construction of the periodic D to N map requires elimination of the terms PnP_{n} and QnQ_{n}. In our study of INVP in §2, u^(λ;t)\hat{u}(\lambda;t) was treated as if it were data and shown only at the conclusion of the argument not to contribute to the derived solution formula. Fortunately, for the periodic nonlocal value problems, a simpler approach is possible.

Because the coefficient (inωiλ3)(\mathrm{i}n\omega-\mathrm{i}\lambda^{3}) of QnQ_{n} and PnP_{n} in equations (3.5)–(3.6) has zeros, QnQ_{n} and PnP_{n} may be eliminated by selecting the particular λ\lambda values

λ=αjλn,whereλn=sgn(n)|n|ω3,j{0,1,2},n,\lambda=\alpha^{j}\lambda_{n},\qquad\mbox{where}\qquad\lambda_{n}=\operatorname{sgn}(n)\sqrt[3]{\left\lvert n\right\rvert\omega},\qquad j\in\{0,1,2\},\qquad n\in\mathbb{Z}, (3.7)

and α=exp(2πi/3)\alpha=\exp(2\pi\mathrm{i}/3). The finitude of Pn(λ)P_{n}(\lambda) and Qn(λ)Q_{n}(\lambda) for all λ\lambda\in\mathbb{C} follows from the fact that they are defined as compact integrals of qq, which is assumed to be integrable in both space and time. Thus, the left sides of the (nonlocal) periodic relations vanish, leaving equations linking only the boundary and nonlocal values. We shall solve the resulting system.

For notational convenience, we define κ(y)=K(1y)\kappa(y)=K(1-y) so that κ^(λ)=eiλK^(λ)\hat{\kappa}(\lambda)=\mathrm{e}^{-\mathrm{i}\lambda}\widehat{K}(-\lambda). The nonlocal periodic relation (3.6) reduces to

(inωiλ3)Pn(λ)=[An2+iλAn1λ2Hn2]κ^(λ)[Γn2+iλHn1λ2Hn0].(\mathrm{i}n\omega-\mathrm{i}\lambda^{3})P_{n}(\lambda)=\left[A_{n\;}^{2}+\mathrm{i}\lambda A_{n\;}^{1}-\lambda^{2}H_{n}^{2}\right]-\hat{\kappa}(\lambda)\left[\Gamma_{n}^{2}+\mathrm{i}\lambda H_{n}^{1}-\lambda^{2}H_{n}^{0}\right]. (3.8)

For each nonzero integer nn, the left side of this equation has three zeros, as given by equation (3.7). Therefore,

(1iλnκ^(λn)1iαλnκ^(αλn)1iα2λnκ^(α2λn))(An2An1Γn2)=(𝒩n(λn)𝒩n(αλn)𝒩n(α2λn)),\begin{pmatrix}1&\mathrm{i}\lambda_{n}&-\hat{\kappa}(\lambda_{n})\\ 1&\mathrm{i}\alpha\lambda_{n}&-\hat{\kappa}(\alpha\lambda_{n})\\ 1&\mathrm{i}\alpha^{2}\lambda_{n}&-\hat{\kappa}(\alpha^{2}\lambda_{n})\end{pmatrix}\begin{pmatrix}A_{n\;}^{2}\\ A_{n\;}^{1}\\ \Gamma_{n}^{2}\end{pmatrix}=\begin{pmatrix}\mathcal{N}_{n}(\lambda_{n})\\ \mathcal{N}_{n}(\alpha\lambda_{n})\\ \mathcal{N}_{n}(\alpha^{2}\lambda_{n})\end{pmatrix}, (3.9)

where

𝒩n(λ)=λ2Hn2+iλκ^(λ)Hn1λ2κ^(λ)Hn0.\mathcal{N}_{n}(\lambda)=\lambda^{2}H_{n}^{2}+\mathrm{i}\lambda\hat{\kappa}(\lambda)H_{n}^{1}-\lambda^{2}\hat{\kappa}(\lambda)H_{n}^{0}.

If instead n=0n=0, then system (3.9) is guaranteed to be rank deficient. However, differentiating equation (3.8) twice with respect to λ\lambda and evaluating at λ=0\lambda=0 yields an equation for Γ02\Gamma_{0}^{2}. Therefore,

Γn2={2(ddλκ^(0)H01κ^(0)H00+H02)/d2dλ2κ^(0)if n=0,j=02αj𝒩n(αjλn)/j=02αjκ^(αjλn),otherwise.\Gamma_{n}^{2}=\begin{cases}2\left(\frac{\,\mathrm{d}}{\,\mathrm{d}\lambda}\hat{\kappa}(0)H_{0}^{1}-\hat{\kappa}(0)H_{0}^{0}+H_{0}^{2}\right)/\frac{\,\mathrm{d}^{2}}{\,\mathrm{d}\lambda^{2}}\hat{\kappa}(0)&\mbox{if }n=0,\\ \sum_{j=0}^{2}\alpha^{j}\mathcal{N}_{n}(\alpha^{j}\lambda_{n})/\sum_{j=0}^{2}\alpha^{j}\hat{\kappa}(\alpha^{j}\lambda_{n}),&\mbox{otherwise.}\end{cases} (3.10)

The exceptional case is if either of the denominators in equation (3.10) is zero. In that case, for generic hkh_{k}, no solution for Γn2\Gamma_{n}^{2} exists. Therefore, we have found a necessary condition for periodic nonlocal value problem (3.1) to have a solution:

Criterion 13.

Both d2dλ2κ^(0)\frac{\,\mathrm{d}^{2}}{\,\mathrm{d}\lambda^{2}}\hat{\kappa}(0) and, for all n{0}n\in\mathbb{Z}\setminus\{0\}, j=02αjκ^(αjλn)\sum_{j=0}^{2}\alpha^{j}\hat{\kappa}(\alpha^{j}\lambda_{n}) are nonzero.

We proceed under our original assumption that qq exists, which implies that equation (3.10) is valid. Evaluating the periodic relation (3.5) at the λ\lambda values (3.7), we determine that

(1iλnλn21iαλnα2λn21iα2λnαλn2)(Gn2Gn1Gn0)=(𝔑n(λ)𝔑n(αλ)𝔑n(α2λ)),\begin{pmatrix}1&\mathrm{i}\lambda_{n}&-\lambda_{n}^{2}\\ 1&\mathrm{i}\alpha\lambda_{n}&-\alpha^{2}\lambda_{n}^{2}\\ 1&\mathrm{i}\alpha^{2}\lambda_{n}&-\alpha\lambda_{n}^{2}\end{pmatrix}\begin{pmatrix}G_{n}^{2}\\ G_{n}^{1}\\ G_{n}^{0}\end{pmatrix}=\begin{pmatrix}\mathfrak{N}_{n}(\lambda)\\ \mathfrak{N}_{n}(\alpha\lambda)\\ \mathfrak{N}_{n}(\alpha^{2}\lambda)\end{pmatrix}, (3.11)

where

𝔑n(λ)=eiλ[Γn2+iλHn1λ2Hn0].\mathfrak{N}_{n}(\lambda)=\mathrm{e}^{-\mathrm{i}\lambda}\left[\Gamma_{n}^{2}+\mathrm{i}\lambda H_{n}^{1}-\lambda^{2}H_{n}^{0}\right].

System (3.11) has determinant 33λn3-3\sqrt{3}\lambda_{n}^{3}, so is full rank for all n0n\neq 0. To determine the zero indexed left boundary Fourier coefficients, we use equations obtained by differentiating the periodic relation (3.5) 0, 1, and 2 times, respectively, and evaluating at λ=0\lambda=0. The solution is

Gn2\displaystyle G_{n}^{2} ={Γ02if n=0,13j=02𝔑n(αjλn)otherwise,\displaystyle=\begin{cases}\Gamma_{0}^{2}&\mbox{if }n=0,\\ \frac{1}{3}\sum_{j=0}^{2}\mathfrak{N}_{n}(\alpha^{j}\lambda_{n})&\mbox{otherwise,}\end{cases} (3.12a)
Gn1\displaystyle G_{n}^{1} ={H01Γ02if n=0,13iλnj=02α2j𝔑n(αjλn)otherwise,\displaystyle=\begin{cases}H_{0}^{1}-\Gamma_{0}^{2}&\mbox{if }n=0,\\ \frac{1}{3\mathrm{i}\lambda_{n}}\sum_{j=0}^{2}\alpha^{2j}\mathfrak{N}_{n}(\alpha^{j}\lambda_{n})&\mbox{otherwise,}\end{cases} (3.12b)
Gn0\displaystyle G_{n}^{0} ={H00H01+12Γ02if n=0,13λn2j=02αj𝔑n(αjλn)otherwise.\displaystyle=\begin{cases}H_{0}^{0}-H_{0}^{1}+\frac{1}{2}\Gamma_{0}^{2}&\mbox{if }n=0,\\ \frac{1}{-3\lambda_{n}^{2}}\sum_{j=0}^{2}\alpha^{j}\mathfrak{N}_{n}(\alpha^{j}\lambda_{n})&\mbox{otherwise.}\end{cases} (3.12c)

3.4 Solution of the time periodic problem

Proposition 14.

Suppose qq satisfies problem (3.1) with generic data piecewise continuous hkh_{k}, and KK is such that criterion 13 holds for κ(y)=K(1y)\kappa(y)=K(1-y). Then

q(x,t)=12πeiλxneinωt[Gn2+iλGn1λ2Gn0]eiλ[Γn2+iλHn1λ2Hn0]inωiλ3dλ,q(x,t)=\frac{1}{2\pi}\int_{-\infty}^{\infty}\mathrm{e}^{\mathrm{i}\lambda x}\sum_{n\in\mathbb{Z}}\mathrm{e}^{\mathrm{i}n\omega t}\frac{\left[G_{n}^{2}+\mathrm{i}\lambda G_{n}^{1}-\lambda^{2}G_{n}^{0}\right]-\mathrm{e}^{-\mathrm{i}\lambda}\left[\Gamma_{n}^{2}+\mathrm{i}\lambda H_{n}^{1}-\lambda^{2}H_{n}^{0}\right]}{\mathrm{i}n\omega-\mathrm{i}\lambda^{3}}\,\mathrm{d}\lambda,

where HnjH_{n}^{j} are defined in equations (3.4), Γn2\Gamma_{n}^{2} are defined in equation (3.10), and GnjG_{n}^{j} are defined in equations (3.12).

Proof.

The Fourier coefficients on the right of equation (3.5) are determined and the criterion justified above. Formally, the result follows from equation (3.5) by the fourth of equations (3.4) and the usual inversion theorem for the spatial Fourier transform. It remains only to justify the convergence of the series and integral.

By Parseval’s theorem, for each kk, (Hnk)n(H_{n}^{k})_{n\in\mathbb{Z}} is square summable. By definition, Γn2\Gamma_{n}^{2} has dominant term λn2Hn0\lambda_{n}^{2}H_{n}^{0}, and λn=Θ(n1/3)\lambda_{n}=\Theta(n^{1/3}), so (Γn2/n)n(\Gamma_{n}^{2}/n)_{n\in\mathbb{Z}} is also square summable. Similarly, for each jj, (Gnj/n)n(G_{n}^{j}/n)_{n\in\mathbb{Z}} are also square summable. Because G0jG_{0}^{j} and Γ02\Gamma_{0}^{2} were constructed to ensure that the numerator has a zero of order 3 at 0, and because the powers of λ\lambda appearing in the numerator are each no greater than 22, the convergence is uniform in λ\lambda, and the value of the series decays as λ±\lambda\to\pm\infty. Hence, by the Riemann-Lebesgue lemma, the integral also converges. ∎

Remark 15.

We have not investigated the failure of criterion 13. It may be that this case corresponds, as it does in certain IBVP for the linear Schrödinger equation [13, 8], to INVP whose solutions blow up linearly in time, despite having periodic boundary data. We emphasize that it is unsurprising that illposedness of periodic problems like problem (3.1) can occur. For example, in the heat equation, an IBVP with a nonzero Neumann conditions should be expected have a solution that blows up in time, rather than behaving periodically.

To investigate more precisely the illposedness would require a careful asymptotic analysis of formula (2.15); the method of §3 cannot provide any insight on the cause of its own failure, because it is predicated on the existence of a solution to problem (3.1).

4 Long time asymptotics

In this section, we study the long time behaviour of solutions to problem (1.1) with either zero or periodic boundary and nonlocal data. The results are stated in proposition 16 and theorem 17.

We begin by studying the following problem, which is a special case of problem (1.1) with homogeneous boundary and nonlocal conditions.

Problem (Homogeneous finite interval problem for the Airy equation with one nonlocal condition).
[t+x3]v(x,t)\displaystyle[\partial_{t}+\partial_{x}^{3}]v(x,t) =0\displaystyle=0 (x,t)\displaystyle(x,t) (0,1)×(0,),\displaystyle\in(0,1)\times(0,\infty), (\theparentequation.PDE)
v(x,0)\displaystyle v(x,0) =V(x)\displaystyle=V(x) x\displaystyle x [0,1],\displaystyle\in[0,1], (\theparentequation.IC)
v(1,t)\displaystyle v(1,t) =0\displaystyle=0 t\displaystyle t [0,),\displaystyle\in[0,\infty), (\theparentequation.BC0)
vx(1,t)\displaystyle v_{x}(1,t) =0\displaystyle=0 t\displaystyle t [0,),\displaystyle\in[0,\infty), (\theparentequation.BC1)
01K(y)v(y,t)dy\displaystyle\int_{0}^{1}K(y)v(y,t)\,\mathrm{d}y =0\displaystyle=0 t\displaystyle t [0,),\displaystyle\in[0,\infty), (\theparentequation.NC)
in which VV and KK are sufficiently smooth.

This is the same as problem (1.1), except that here we insist h0=h1=h2=0h_{0}=h_{1}=h_{2}=0. In particular, corollary 4 applies, with simplifications of the definitions of N0N_{0} and N1N_{1}. An asymptotic analysis of this formula for v(x,t)v(x,t), presented at the end of this section allows us to prove the following proposition.

Proposition 16.

Suppose vv satisfies problem (4.1) and the criteria of theorem 3 hold. Then, uniformly in xx, v(x,t)=𝒪(t1)v(x,t)=\mathcal{O}(t^{-1}) as tt\to\infty.

Now consider problem (1.1), but with all of h0,h1,h2h_{0},h_{1},h_{2} having common period TT. We seek an expression for u(x,t)u(x,t) valid in the long time regime. We use the principle of linear superposition to express u(x,t)=q(x,t)+v(x,t)u(x,t)=q(x,t)+v(x,t), with qq satisfying problem (3.1) and vv satisfying problem (4.1) in which V(x)=U(x)q(x,0)V(x)=U(x)-q(x,0). By proposition 16, vv decays. From proposition 14, we can obtain an expression for qq, hence an asymptotically valid expression for uu. This justifies the following theorem.

Theorem 17.

Suppose uu satisfies problem (1.1) in which all of h0,h1,h2h_{0},h_{1},h_{2} are generic piecewise continuous functions with common period TT. Suppose that the corresponding homogeneous problem (4.1) satisfies the criteria of theorem 3. Suppose that KK is such that κ(y)=K(1y)\kappa(y)=K(1-y) satisfies criterion 13. Then

u(x,t)=12πeiλxneinωt[Gn2+iλGn1λ2Gn0]eiλ[Γn2+iλHn1λ2Hn0]inωiλ3dλ+𝒪(t1),u(x,t)=\frac{1}{2\pi}\int_{-\infty}^{\infty}\mathrm{e}^{\mathrm{i}\lambda x}\sum_{n\in\mathbb{Z}}\mathrm{e}^{\mathrm{i}n\omega t}\frac{\left[G_{n}^{2}+\mathrm{i}\lambda G_{n}^{1}-\lambda^{2}G_{n}^{0}\right]-\mathrm{e}^{-\mathrm{i}\lambda}\left[\Gamma_{n}^{2}+\mathrm{i}\lambda H_{n}^{1}-\lambda^{2}H_{n}^{0}\right]}{\mathrm{i}n\omega-\mathrm{i}\lambda^{3}}\,\mathrm{d}\lambda+\mathcal{O}(t^{-1}),

where HnjH_{n}^{j} are defined in equations (3.4), Γn2\Gamma_{n}^{2} are defined in equation (3.10), and GnjG_{n}^{j} are defined in equations (3.12).

Proof of proposition 16.

Let

n(λ)=1Δ(λ)j=02αjN1(λ)(αjλ),n(\lambda)=\frac{1}{\Delta(\lambda)}\sum_{j=0}^{2}\alpha^{j}N_{1}(\lambda)(\alpha^{j}\lambda),

and observe that, because the boundary and nonlocal conditions are homogeneous,

N1(λ)=01K(y)eiλyV^(λ;y,1)dy;N_{1}(\lambda)=\int_{0}^{1}K(y)\mathrm{e}^{\mathrm{i}\lambda y}\widehat{V}(\lambda;y,1)\,\mathrm{d}y;

both N1N_{1} and nn are truly independent of tt^{\prime}. By corollary 4, 2πu(x,t)=I+(t)+I(t)2\pi u(x,t)=I^{+}(t)+I^{-}(t), where we have suppressed the xx depenence of I±I^{\pm}, and

I+(t)\displaystyle I^{+}(t) =ER+eiλx+iλ3t(eiλn(λ)V^(λ))dλ,\displaystyle=\int_{\partial E_{R}^{+}}\mathrm{e}^{\mathrm{i}\lambda x+\mathrm{i}\lambda^{3}t}\left(\mathrm{e}^{-\mathrm{i}\lambda}n(\lambda)-\widehat{V}(\lambda)\right)\,\mathrm{d}\lambda,
I(t)\displaystyle I^{-}(t) =EReiλ(x1)+iλ3tn(λ)dλ.\displaystyle=\int_{\partial E_{R}^{-}}\mathrm{e}^{\mathrm{i}\lambda(x-1)+\mathrm{i}\lambda^{3}t}n(\lambda)\,\mathrm{d}\lambda.

We will analyse each of I±I^{\pm} separately.

Interpreting, as we must due to its derivation from equation (2.4), I(t)I^{-}(t) as a Cauchy principal value contour integral, and integrating by parts,

I(t)=13itlimb+{[eiλ3teiλ(x1)n(λ)λ2]λ=α1/2bλ=α2bΓbeiλ3tddλ[eiλ(x1)n(λ)λ2]dλ},I^{-}(t)=\frac{1}{3\mathrm{i}t}\lim_{b\to+\infty}\left\{\left[\mathrm{e}^{\mathrm{i}\lambda^{3}t}\frac{\mathrm{e}^{\mathrm{i}\lambda(x-1)}n(\lambda)}{\lambda^{2}}\right]^{\lambda=\alpha^{2}b}_{\lambda=\alpha^{-1/2}b}-\int_{\Gamma_{b}^{-}}\mathrm{e}^{\mathrm{i}\lambda^{3}t}\frac{\,\mathrm{d}}{\,\mathrm{d}\lambda}\left[\frac{\mathrm{e}^{\mathrm{i}\lambda(x-1)}n(\lambda)}{\lambda^{2}}\right]\,\mathrm{d}\lambda\right\},

where Γb=ERB(0,b)\Gamma_{b}^{-}=\partial E_{R}^{-}\cap B(0,b). Because α2b,α1/2bclos(DR)\alpha^{2}b,\alpha^{-1/2}b\in\operatorname{clos}(D_{R}^{-}), lemma 2 guarantees that n(λ)=𝒪(b1)n(\lambda)=\mathcal{O}(b^{-1}). Along the rays α2+\alpha^{2}\mathbb{R}^{+} and α1/2+\alpha^{-1/2}\mathbb{R}^{+}, |eiλ3t|=1\left\lvert\mathrm{e}^{\mathrm{i}\lambda^{3}t}\right\rvert=1 and |eiλ(x1)|1\left\lvert\mathrm{e}^{\mathrm{i}\lambda(x-1)}\right\rvert\leqslant 1, so the decay of both boundary terms as b+b\to+\infty is justified:

I(t)=13itlimb+{0Γbeiλ3tddλ[eiλ(x1)n(λ)λ2]dλ}.I^{-}(t)=\frac{1}{3\mathrm{i}t}\lim_{b\to+\infty}\left\{0-\int_{\Gamma_{b}^{-}}\mathrm{e}^{\mathrm{i}\lambda^{3}t}\frac{\,\mathrm{d}}{\,\mathrm{d}\lambda}\left[\frac{\mathrm{e}^{\mathrm{i}\lambda(x-1)}n(\lambda)}{\lambda^{2}}\right]\,\mathrm{d}\lambda\right\}. (4.2)

Differentiating,

eiλ3tddλ[eiλ(x1)n(λ)λ2]=1λ2×eiλ3t+iλ(x1)×[(i(x1)2λ)n(λ)+n(λ)].\mathrm{e}^{\mathrm{i}\lambda^{3}t}\frac{\,\mathrm{d}}{\,\mathrm{d}\lambda}\left[\frac{\mathrm{e}^{\mathrm{i}\lambda(x-1)}n(\lambda)}{\lambda^{2}}\right]=\frac{1}{\lambda^{2}}\times\mathrm{e}^{\mathrm{i}\lambda^{3}t+\mathrm{i}\lambda(x-1)}\times\left[\left(\mathrm{i}(x-1)-\frac{2}{\lambda}\right)n(\lambda)+n^{\prime}(\lambda)\right].

The exponential factor is, uniformly in x,tx,t, bounded in modulus by 11, and n(λ)n(\lambda) decays as λ\lambda\to\infty along ER±\partial E_{R}^{\pm}. We shall argue that n(λ)n^{\prime}(\lambda) is also bounded (in fact decaying), whence dominated convergence gives a bound on the b+b\to+\infty limit in equation (4.2) that is uniform in both xx and tt. We calculate

n(λ)=1Δ(λ)j=02α2jN1(αjλ)Δ(λ)[Δ(λ)]2j=02α2jN1(αjλ),n^{\prime}(\lambda)=\frac{1}{\Delta(\lambda)}\sum_{j=0}^{2}\alpha^{2j}N^{\prime}_{1}(\alpha^{j}\lambda)-\frac{\Delta^{\prime}(\lambda)}{[\Delta(\lambda)]^{2}}\sum_{j=0}^{2}\alpha^{2j}N_{1}(\alpha^{j}\lambda),

where

Δ(λ)=ij=02α2j01eiλαjyyκ(y)dy=𝒪(Δ(λ)),\Delta^{\prime}(\lambda)=-\mathrm{i}\sum_{j=0}^{2}\alpha^{2j}\int_{0}^{1}\mathrm{e}^{-\mathrm{i}\lambda\alpha^{j}y}y\kappa(y)\,\mathrm{d}y=\mathcal{O}(\Delta(\lambda)),

and

N1(αjλ)=iαj01K(y)eiαjλyy1eiαjλz(yz)V(z)dzdy=𝒪(λ1Δ(λ)),N^{\prime}_{1}(\alpha^{j}\lambda)=-\mathrm{i}\alpha^{j}\int_{0}^{1}K(y)\mathrm{e}^{\mathrm{i}\alpha^{j}\lambda y}\int_{y}^{1}\mathrm{e}^{-\mathrm{i}\alpha^{j}\lambda z}(y-z)V(z)\,\mathrm{d}z\,\mathrm{d}y=\mathcal{O}\left(\lambda^{-1}\Delta(\lambda)\right),

as λ\lambda\to\infty along the contour ER±\partial E_{R}^{\pm}, hence n(λ)n^{\prime}(\lambda) also decays. This shows that I(t)=𝒪(t1)I^{-}(t)=\mathcal{O}(t^{-1}), uniformly in xx, as tt\to\infty.

The argument for I+I^{+} follows the same structure as that for II^{-}. Integrating by parts,

I+(t)=13itlimb+{[eiλ3t+iλxeiλn(λ)V^(λ)λ2]λ=αb,bλ=b,α1/2bΓb+eiλ3tddλ[eiλxeiλn(λ)V^(λ)λ2]dλ},I^{+}(t)=\frac{1}{3\mathrm{i}t}\lim_{b\to+\infty}\left\{\left[\mathrm{e}^{\mathrm{i}\lambda^{3}t+\mathrm{i}\lambda x}\frac{\mathrm{e}^{-\mathrm{i}\lambda}n(\lambda)-\widehat{V}(\lambda)}{\lambda^{2}}\right]^{\lambda=-b,\,\alpha^{1/2}b}_{\lambda=\alpha b,\,b}\right.\\ \left.-\int_{\Gamma_{b}^{+}}\mathrm{e}^{\mathrm{i}\lambda^{3}t}\frac{\,\mathrm{d}}{\,\mathrm{d}\lambda}\left[\mathrm{e}^{\mathrm{i}\lambda x}\frac{\mathrm{e}^{-\mathrm{i}\lambda}n(\lambda)-\widehat{V}(\lambda)}{\lambda^{2}}\right]\,\mathrm{d}\lambda\vphantom{\left[\mathrm{e}^{\mathrm{i}\lambda^{3}t+\mathrm{i}\lambda x}\frac{\mathrm{e}^{-\mathrm{i}\lambda}n(\lambda)-\widehat{V}(\lambda)}{\lambda^{2}}\right]^{\lambda=-b,\,\alpha^{1/2}b}_{\lambda=\alpha b,\,b}}\right\},

where Γb+=ER+B(0,b)\Gamma_{b}^{+}=\partial E_{R}^{+}\cap B(0,b). By lemma 2,

eiλn(λ)V^(λ)=𝒪(1),\mathrm{e}^{-\mathrm{i}\lambda}n(\lambda)-\widehat{V}(\lambda)=\mathcal{O}\left(1\right),

as λ\lambda\to\infty along DR+\partial D_{R}^{+}, and the same argument applies to the limit as λ\lambda\to\infty along \mathbb{R}. Hence, this numerator is bounded on ER+\partial E_{R}^{+}, and the boundary terms decay in the limit b+b\to+\infty;

I+(t)=13itlimb+{0Γb+eiλ3tddλ[eiλxeiλn(λ)V^(λ)λ2]dλ}.I^{+}(t)=\frac{1}{3\mathrm{i}t}\lim_{b\to+\infty}\left\{0-\int_{\Gamma_{b}^{+}}\mathrm{e}^{\mathrm{i}\lambda^{3}t}\frac{\,\mathrm{d}}{\,\mathrm{d}\lambda}\left[\mathrm{e}^{\mathrm{i}\lambda x}\frac{\mathrm{e}^{-\mathrm{i}\lambda}n(\lambda)-\widehat{V}(\lambda)}{\lambda^{2}}\right]\,\mathrm{d}\lambda\vphantom{\left[\mathrm{e}^{\mathrm{i}\lambda^{3}t+\mathrm{i}\lambda x}\frac{\mathrm{e}^{-\mathrm{i}\lambda}n(\lambda)-\widehat{V}(\lambda)}{\lambda^{2}}\right]^{\lambda=-b,\,\alpha^{1/2}b}_{\lambda=\alpha b,\,b}}\right\}. (4.3)

The remaining integrand can be reexpressed as

1λ2×eiλ3t+iλx×[(ix2λ)(eiλn(λ)V^(λ))+ddλ(eiλn(λ)V^(λ))],\frac{1}{\lambda^{2}}\times\mathrm{e}^{\mathrm{i}\lambda^{3}t+\mathrm{i}\lambda x}\times\left[\left(\mathrm{i}x-\frac{2}{\lambda}\right)\left(\mathrm{e}^{-\mathrm{i}\lambda}n(\lambda)-\widehat{V}(\lambda)\right)+\frac{\,\mathrm{d}}{\,\mathrm{d}\lambda}\left(\mathrm{e}^{-\mathrm{i}\lambda}n(\lambda)-\widehat{V}(\lambda)\right)\right], (4.4)

and the first term in the bracket, as above, is 𝒪(1)\mathcal{O}(1). To obtain a bound on the derivative, we apply a similar reexpression to that used in the proof of lemma 2, but keep the lower order terms. Because

eiλΔ(λ)01K(y)eiλydy=11Δ(λ)j=12αjκ^(αjλ),\frac{\mathrm{e}^{-\mathrm{i}\lambda}}{\Delta(\lambda)}\int_{0}^{1}K(y)\mathrm{e}^{\mathrm{i}\lambda y}\,\mathrm{d}y=1-\frac{1}{\Delta(\lambda)}\sum_{j=1}^{2}\alpha^{j}\hat{\kappa}(\alpha^{j}\lambda),

we can rewrite

eiλn(λ)V^(λ)=eiλΔ(λ)(j=12αj01K(y)y1eiαjλ(zy)V(z)dzdy01K(y)0yeiλ(yz)V(z)dzdy)+V^(λ)Δ(λ)j=12αj01eiαjλ(1y)K(y)dy.\mathrm{e}^{-\mathrm{i}\lambda}n(\lambda)-\widehat{V}(\lambda)\\ =\frac{\mathrm{e}^{-\mathrm{i}\lambda}}{\Delta(\lambda)}\left(\sum_{j=1}^{2}\alpha^{j}\int_{0}^{1}K(y)\int_{y}^{1}\mathrm{e}^{-\mathrm{i}\alpha^{j}\lambda(z-y)}V(z)\,\mathrm{d}z\,\mathrm{d}y-\int_{0}^{1}K(y)\int_{0}^{y}\mathrm{e}^{\mathrm{i}\lambda(y-z)}V(z)\,\mathrm{d}z\,\mathrm{d}y\right)\\ +\frac{\widehat{V}(\lambda)}{\Delta(\lambda)}\sum_{j=1}^{2}\alpha^{j}\int_{0}^{1}\mathrm{e}^{-\mathrm{i}\alpha^{j}\lambda(1-y)}K(y)\,\mathrm{d}y.

The derivative of this quantity is

eiλ(i+Δ(λ)Δ(λ))Δ(λ)(j=12αj01K(y)y1eiαjλ(zy)V(z)dzdy01K(y)0yeiλ(yz)V(z)dzdy)+ieiλΔ(λ)(j=12α2j01K(y)y1eiαjλ(zy)(yz)V(z)dzdy01K(y)0yeiλ(yz)(yz)V(z)dzdy)iΔ(λ)j=12α2j01eiαjλ(1y)(1y)K(y)dy+j=12αjκ^(αjλ)[Δ(λ)]2(Δ(λ)V^(λ)Δ(λ)ddλV^(λ)).\frac{-\mathrm{e}^{-\mathrm{i}\lambda}\left(\mathrm{i}+\frac{\Delta^{\prime}(\lambda)}{\Delta(\lambda)}\right)}{\Delta(\lambda)}\left(\sum_{j=1}^{2}\alpha^{j}\int_{0}^{1}K(y)\int_{y}^{1}\mathrm{e}^{-\mathrm{i}\alpha^{j}\lambda(z-y)}V(z)\,\mathrm{d}z\,\mathrm{d}y-\int_{0}^{1}K(y)\int_{0}^{y}\mathrm{e}^{\mathrm{i}\lambda(y-z)}V(z)\,\mathrm{d}z\,\mathrm{d}y\right)\\ +\frac{\mathrm{i}\mathrm{e}^{-\mathrm{i}\lambda}}{\Delta(\lambda)}\left(\sum_{j=1}^{2}\alpha^{2j}\int_{0}^{1}K(y)\int_{y}^{1}\mathrm{e}^{-\mathrm{i}\alpha^{j}\lambda(z-y)}(y-z)V(z)\,\mathrm{d}z\,\mathrm{d}y-\int_{0}^{1}K(y)\int_{0}^{y}\mathrm{e}^{\mathrm{i}\lambda(y-z)}(y-z)V(z)\,\mathrm{d}z\,\mathrm{d}y\right)\\ -\frac{\mathrm{i}}{\Delta(\lambda)}\sum_{j=1}^{2}\alpha^{2j}\int_{0}^{1}\mathrm{e}^{-\mathrm{i}\alpha^{j}\lambda(1-y)}(1-y)K(y)\,\mathrm{d}y+\frac{\sum_{j=1}^{2}\alpha^{j}\hat{\kappa}(\alpha^{j}\lambda)}{[\Delta(\lambda)]^{2}}\left(\Delta^{\prime}(\lambda)\widehat{V}(\lambda)-\Delta(\lambda)\frac{\,\mathrm{d}}{\,\mathrm{d}\lambda}\widehat{V}(\lambda)\right).

Similar arguments to those presented in lemma 2 justify that this quantity is 𝒪(1)\mathcal{O}(1) as λ\lambda\to\infty along the contours of interest. Therefore, the bracketed factor in expression (4.4) is 𝒪(1)\mathcal{O}(1). Because the exponential in that expression is, uniformly in x,tx,t, bounded in modulus, dominated convergence establishes a uniform bound on the remaining integral on the right of equation (4.3). Hence I+(t)=𝒪(t1)I^{+}(t)=\mathcal{O}(t^{-1}), uniformly in xx, as tt\to\infty. ∎

Remark 18.

In proposition 16, we have not attempted to obtain the leading order behaviour of v(x,t)v(x,t) in the long time limit, only that it decays at least linearly. It may even be that further integration by parts reveals uniform quadratic or faster decay of vv. The arguments presented above could be extended to investigate these questions.

Remark 19.

One may extend this linear superposition argument to study the case in which the boundary and nonlocal data are not periodic but only asymptotically periodic. In this case, the problem for qq need not be altered, but the problem for vv would have all hkh_{k} decaying instead of identically zero. How fast must the data decay to their perodic limit to obtain decay of the corresponding solution vv? Questions such as this require a generalisation of proposition 16 to study the extra terms.

Acknowledgement

B. Normatov gratefully acknowledges support from Yale-NUS College summer research programme. D. A. Smith gratefully acknowledges support from Yale-NUS College seed grant IG21-SG101.

References

  • [1] S. A. Aitzhan, S. Bhandari, and D. A. Smith, Fokas diagonalization of piecewise constant coefficient linear differential operators on finite intervals and networks, Acta Appl. Math. 177 (2022), no. 2, 1–66.
  • [2] M Asvestas, EP Papadopoulou, AG Sifalakis, and YG Saridakis, The unified transform for a class of reaction-diffusion problems with discontinuous time dependent parameters, Proceedings of the World Congress on Engineering, vol. 1, 2015.
  • [3] B. Deconinck, B. Pelloni, and N. E. Sheils, Non-steady-state heat conduction in composite walls, Proc. R. Soc. Lond. Ser. A Math. Phys. Eng. Sci. 470 (2014), no. 2165, 20130605.
  • [4] B. Deconinck, N. E. Sheils, and D. A. Smith, The linear KdV equation with an interface, Comm. Math. Phys. 347 (2016), 489–509.
  • [5] B. Deconinck, T. Trogdon, and V. Vasan, The method of Fokas for solving linear partial differential equations, SIAM Rev. 56 (2014), no. 1, 159–186.
  • [6] B. Deconinck and V. Vasan, Well-posedness of boundary-value problems for the linear Benjamin-Bona-Mahony equation, Discrete & Continuous Dyn. Sys. A 33 (2013), no. 7, 3171–3188.
  • [7] Bernard Deconinck, Guo Qi, Eli Shlizerman, and Vishal Vasan, Fokas’s unified transform method for linear systems, Quarterly of Applied Mathematics 76 (2018), 463–488.
  • [8] G. M. Dujardin, Asymptotics of linear initial boundary value problems with periodic boundary data on the half-line and finite intervals, Proc. R. Soc. Lond. Ser. A Math. Phys. Eng. Sci. 465 (2009), no. 2111, 3341–3360.
  • [9] A. S. Fokas, A unified approach to boundary value problems, CBMS-SIAM, 2008.
  • [10] A. S. Fokas and J. Lenells, The unified method: II NLS on the half-line with t-periodic boundary conditions, J. Phys. A 45 (2012), 195202.
  • [11] A. S. Fokas and B. Pelloni, Two-point boundary value problems for linear evolution equations, Math. Proc. Cambridge Philos. Soc. 131 (2001), 521–543.
  • [12]  , Boundary value problems for Boussinesq type systems, Math. Phys. Anal. Geom. 8 (2005), no. 1, 59–96.
  • [13] A. S. Fokas, B. Pelloni, and D. A. Smith, Time-periodic linear boundary value problems on a finite interval, Quart. Appl. Math. 80 (2022), 481–506.
  • [14] A. S. Fokas and D. A. Smith, Evolution PDEs and augmented eigenfunctions. Finite interval, Adv. Differential Equations 21 (2016), no. 7/8, 735–766.
  • [15] A. S. Fokas and M. C. van der Weele, The unified transform for evolution equations on the half-line with time-periodic boundary conditions, Stud. Appl. Math. 147 (2021), 1339–1368.
  • [16] C. M. Johnston, C. T. Gartman, and D. Mantzavinos, The linearized classical Boussinesq system on the half-line, Stud. Appl. Math. 145 (2021), no. 3, 635–657.
  • [17] M. Krstic and A. Smyshlyaev, Boundary control of PDEs: a course on backstepping designs, Advances in Design and Control, SIAM, Philadelphia, PA, 2008.
  • [18] R. E. Langer, The zeros of exponential sums and integrals, Bull. Amer. Math. Soc. 37 (1931), 213–239.
  • [19] J. Locker, Self-adjointness for multi-point differential operators, Pacific J. Math. 45 (1973), no. 2, 561–570.
  • [20]  , Spectral theory of non-self-adjoint two-point differential operators, Mathematical Surveys and Monographs, vol. 73, American Mathematical Society, Providence, Rhode Island, 2000.
  • [21] D. Mantzavinos, Initial-boundary value problems for linear and integrable nonlinear evolution PDEs, Ph.D. thesis, University of Cambridge, 2012.
  • [22] P. D. Miller and D. A. Smith, The diffusion equation with nonlocal data, J. Math. Anal. Appl. 466 (2018), no. 2, 1119–1143.
  • [23] T. Özsarı and A. Batal, Pseudo-backstepping and its application to the control of Korteweg–de Vries equation from the right endpoint on a finite domain, SIAM Journal on Control and Optimization 57 (2019), no. 2, 1255–1283.
  • [24] B. Pelloni, Well-posed boundary value problems for linear evolution equations on a finite interval, Math. Proc. Cambridge Philos. Soc. 136 (2004), 361–382.
  • [25]  , The spectral representation of two-point boundary-value problems for third-order linear evolution partial differential equations, Proc. R. Soc. Lond. Ser. A Math. Phys. Eng. Sci. 461 (2005), 2965–2984.
  • [26] B. Pelloni and D. A. Smith, Evolution PDEs and augmented eigenfunctions. Half line, J. Spectr. Theory 6 (2016), 185–213.
  • [27]  , Nonlocal and multipoint boundary value problems for linear evolution equations, Stud. Appl. Math. 141 (2018), no. 1, 46–88.
  • [28] N. Sheils, Multilayer diffusion in a composite medium with imperfect contact, Applied Mathematical Modelling 46 (2017), 450–464.
  • [29] N. E. Sheils and D. A. Smith, Heat equation on a network using the Fokas method, Journal of Physics A: Mathematical and Theoretical 48 (2015), no. 33, 21 pp.
  • [30] N.E. Sheils and B. Deconinck, Interface problems for dispersive equations, Stud. in Appl. Math. 134 (2015), no. 3, 253–275.
  • [31] D. A. Smith, Well-posed two-point initial-boundary value problems with arbitrary boundary conditions, Math. Proc. Cambridge Philos. Soc. 152 (2012), 473–496.
  • [32]  , The unified transform method for linear initial-boundary value problems: a spectral interpretation, Unified transform method for boundary value problems: applications and advances, SIAM, Philadelphia, PA, 2015.
  • [33]  , Fokas diagonalization, Chaos, Fractals and Complexity (Cham) (T. Bountis, F. Vallianatos, A. Provata, D. Kugiumtzis, and Y. Kominis, eds.), Springer, 2022, (submitted) arXiv:2211.10392 [math.SP], pp. 1–16.