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

Optimal Runge approximation for damped nonlocal wave equations and simultaneous determination results

Philipp Zimmermann Departament de Matemàtiques i Informàtica, Universitat de Barcelona, Barcelona, Spain [email protected]
Abstract.

The main purpose of this article is to establish new uniqueness results for Calderón type inverse problems related to damped nonlocal wave equations. To achieve this goal we extend the theory of very weak solutions to our setting, which allows to deduce an optimal Runge approximation theorem. With this result at our disposal, we can prove simultaneous determination results in the linear and semilinear regime.

Keywords. Fractional Laplacian, wave equations, nonlinear PDEs, inverse problems, Runge approximation, very weak solutions.

Mathematics Subject Classification (2020): Primary 35R30; secondary 26A33, 42B37

1. Introduction

In recent years, inverse problems for nonlocal partial differential equations (PDEs) of elliptic, parabolic and hyperbolic type have been studied. This line of research was initiated by Ghosh, Salo and Uhlman [GSU20], in which they have considered the (partial data) Calderón problem related to the fractional Schrödinger equation

(1.1) {((Δ)s+q)u=0 in Ω,u=φ in Ωe,\begin{cases}((-\Delta)^{s}+q)u=0&\text{ in }\Omega,\\ u=\varphi&\text{ in }\Omega_{e},\end{cases}

where Ωn\Omega\subset{\mathbb{R}}^{n} is a bounded domain, Ωe=nΩ¯\Omega_{e}={\mathbb{R}}^{n}\setminus\overline{\Omega}, 0<s<10<s<1, qq is a suitable potential and (Δ)s(-\Delta)^{s} is the fractional Laplacian which is the operator with Fourier symbol |ξ|2s|\xi|^{2s}. In this problem one asks whether the knowledge of the (partial) Dirichlet to Neumann (DN) map

(1.2) Λqφ=(Δ)suφ|W2,φCc(W1),\Lambda_{q}\varphi=(-\Delta)^{s}u_{\varphi}|_{W_{2}},\quad\varphi\in C_{c}^{\infty}(W_{1}),

where W1,W2ΩeW_{1},W_{2}\subset\Omega_{e} are given measurement sets (i.e. nonempty open sets) and uφu_{\varphi} denotes the unique solution to (1.1), uniquely determines the potential qq. The overall strategy to establish unique determination results for the above Calderón problem is as follows (see [GSU20, RS20, RZ23]):

  1. (i)

    Integral identity: Assume that the potentials qjq_{j} are suitably regular, then one can write

    (1.3) (Λq1Λq2)φ1,φ2=Ω(q1q2)(uφ1φ1),(uφ2φ2)dx,\langle(\Lambda_{q_{1}}-\Lambda_{q_{2}})\varphi_{1},\varphi_{2}\rangle=\int_{\Omega}(q_{1}-q_{2})(u_{\varphi_{1}}-\varphi_{1}),(u_{\varphi_{2}}-\varphi_{2})\,dx,

    when the right hand side is interpreted accordingly.

  2. (ii)

    Establish one of the following Runge approximation theorems:

    1. (I)

      W={uf|Ω;fCc(W)}\mathcal{R}_{W}=\{u_{f}|_{\Omega}\,;\,f\in C_{c}^{\infty}(W)\} is dense in L2(Ω)L^{2}(\Omega) (see [GSU20] for qL(Ω)q\in L^{\infty}(\Omega)).

    2. (II)

      W={uff;fCc(W)}\mathscr{R}_{W}=\{u_{f}-f\,;\,f\in C_{c}^{\infty}(W)\} is dense in H~s(Ω)\widetilde{H}^{s}(\Omega) (see [RS20] for Sobolev multipliers qq or [RZ23] for local, bounded bilinear forms).

  3. (iii)

    If the potentials qjq_{j} for j=1,2j=1,2 have suitable continuity properties, then Λq1=Λq2\Lambda_{q_{1}}=\Lambda_{q_{2}} together with ii ensure that there holds q1=q2q_{1}=q_{2} in Ω\Omega.

In iiII, the space H~s(Ω)\widetilde{H}^{s}(\Omega) is the closure of Cc(Ω)C_{c}^{\infty}(\Omega) in the energy space

Hs(n)={u𝒮(n);uHs(n):=DsuL2(n)<},H^{s}({\mathbb{R}}^{n})=\{u\in\mathscr{S}^{\prime}({\mathbb{R}}^{n})\,;\,\|u\|_{H^{s}({\mathbb{R}}^{n})}\vcentcolon=\|\langle D\rangle^{s}u\|_{L^{2}({\mathbb{R}}^{n})}<\infty\},

where Ds\langle D\rangle^{s} is the Bessel potential operator. Observe the similarity of the above strategy to the one of [SU87] for showing unique determination for the classical Calderón problem, where instead of the Runge approximation theorem suitable geometric optics solutions are used. Moreover, the Runge approximation ii relies on a Hahn–Banach argument and the unique continuation property (UCP) of the fractional Laplacian (Δ)s(-\Delta)^{s}. For more results on Calderón problems for elliptic nonlocal PDEs, we refer the interested reader to [GLX17, CLR20, CLL19, LL22, LL23, LZ23, KLZ24, KLW22, LRZ22, LTZ24b, LLU23, CGRU23, LLU23, RZ23, RZ24, CRTZ22, LZ24, FGKU21, Fei21, FKU24, LNZ24] and the references therein.

1.1. Mathematical model and main results

Recently, the above approach for solving elliptic nonlocal inverse problems has also been adapted to deduce uniqueness results for the Calderón problem of nonlocal hyperbolic equations. Let us next describe some of these results in more detail and for this purpose consider the problem

(1.4) {t2u+λ(Δ)stu+(Δ)su+f(u)=0 in ΩT,u=φ on (Ωe)T,u(0)=0,tu(0)=0 on Ω,\begin{cases}\partial_{t}^{2}u+\lambda(-\Delta)^{s}\partial_{t}u+(-\Delta)^{s}u+f(u)=0&\text{ in }\Omega_{T},\\ u=\varphi&\text{ on }(\Omega_{e})_{T},\\ u(0)=0,\,\partial_{t}u(0)=0&\text{ on }\Omega,\end{cases}

where λ\lambda\in{\mathbb{R}} and f:Ω×f\colon\Omega\times{\mathbb{R}}\to{\mathbb{R}} is a possibly nonlinear function. If the problem (1.4) is well-posed in the energy class Hs(n)H^{s}({\mathbb{R}}^{n}), then for any two given measurement sets W1,W2ΩeW_{1},W_{2}\subset\Omega_{e} we may introduce the DN map Λfλ\Lambda^{\lambda}_{f} via

Λfλφ=(λ(Δ)stuφ+(Δ)suφ)|(W2)T,\Lambda^{\lambda}_{f}\varphi=\left.(\lambda(-\Delta)^{s}\partial_{t}u_{\varphi}+(-\Delta)^{s}u_{\varphi})\right|_{(W_{2})_{T}},

whenever φ\varphi is supported in (W1)T(W_{1})_{T} and uφu_{\varphi} is the solution of (1.4). The Calderón problem for (1.4) reads as follows:

Question 1.

Does the DN map Λfλ\Lambda^{\lambda}_{f} uniquely determine the function ff?

A suitable class of nonlinearities are the so-called weak nonlinearities, which are defined next.

Definition 1.1.

We call a Carathéodory function f:Ω×f\colon\Omega\times{\mathbb{R}}\to{\mathbb{R}} weak nonlinearity, if it satisfies the following conditions:

  1. (i)

    ff has partial derivative τf\partial_{\tau}f, which is a Carathéodory function, and there exists aLp(Ω)a\in L^{p}(\Omega) such that

    (1.5) |τf(x,τ)|a(x)+|τ|r\left|\partial_{\tau}f(x,\tau)\right|\lesssim a(x)+|\tau|^{r}

    for all τ\tau\in{\mathbb{R}} and a.e. xΩx\in\Omega. Here the exponents pp and rr satisfy the restrictions

    (1.6) {n/sp,if  2s<n,2<p,if  2s=n,2p,if  2sn.\begin{cases}n/s\leq p\leq\infty,&\,\text{if }\,2s<n,\\ 2<p\leq\infty,&\,\text{if }\,2s=n,\\ 2\leq p\leq\infty,&\,\text{if }\,2s\geq n.\end{cases}

    and

    (1.7) {0r<,if  2sn,0r2sn2s,if  2s<n,\begin{cases}0\leq r<\infty,&\,\text{if }\,2s\geq n,\\ 0\leq r\leq\frac{2s}{n-2s},&\,\text{if }\,2s<n,\end{cases}

    respectively. Moreover, ff fulfills the integrability condition f(,0)L2(Ω)f(\cdot,0)\in L^{2}(\Omega).

  2. (ii)

    The function F:Ω×F\colon\Omega\times{\mathbb{R}}\to{\mathbb{R}} defined via

    F(x,τ)=0τf(x,ρ)𝑑ρF(x,\tau)=\int_{0}^{\tau}f(x,\rho)\,d\rho

    satisfies F(x,τ)0F(x,\tau)\geq 0 for all τ\tau\in{\mathbb{R}} and xΩx\in\Omega.

Let us note that f(x,τ)=q(x)|τ|rτf(x,\tau)=q(x)|\tau|^{r}\tau with rr satisfying (1.7) and 0qL(Ω)0\leq q\in L^{\infty}(\Omega) is a weak nonlinearity. Using the above notions, we can now discuss some of the existing results.

  1. (a)

    The article [KLW22] gives a positive answer for λ=0,f(x,τ)=q(x)τ\lambda=0,f(x,\tau)=q(x)\tau with qL(Ω)q\in L^{\infty}(\Omega). Their proof relied on the observation that the related nonlocal wave equation (1.4) satisfies an L2(ΩT)L^{2}(\Omega_{T}) Runge approximation theorem.

  2. (b)

    The work [Zim24] deals on the one hand with the linear case λ=1\lambda=1, f(x,τ)=q(x)τf(x,\tau)=q(x)\tau with qL(0,T;Lp(Ω))q\in L^{\infty}(0,T;L^{p}(\Omega)), where pp satisfies the restrictions (1.6), and qq is weakly continuous in tt and on the other hand with the nonlinear case λ=1\lambda=1 and ff is a r+1r+1 homogeneous, weak nonlinearity. The uniqueness proofs use substantially that due to presence of the viscosity term (Δ)st(-\Delta)^{s}\partial_{t} solutions uu to (1.4) satisfy tuL2(0,T;Hs(n))\partial_{t}u\in L^{2}(0,T;H^{s}({\mathbb{R}}^{n})) and as a consequence the linearized equations have the Runge approximation property in L2(0,T;H~s(Ω))L^{2}(0,T;\widetilde{H}^{s}(\Omega)).

  3. (c)

    In [LTZ24c], uniqueness is proved in the case λ=0\lambda=0 and ff satisfies the same properties as in b, but with the additional restriction r1r\leq 1. This article only uses an L2(ΩT)L^{2}(\Omega_{T}) Runge approximation result for the linearized nonlocal wave equation.

  4. (d)

    By establishing a theory for very weak solutions of linear nonlocal wave equations with λ=0\lambda=0, the authors of [LTZ24a] could deduce an optimal L2(0,T;H~s(Ω))L^{2}(0,T;\widetilde{H}^{s}(\Omega)) Runge approximation theorem for these equations. This allowed to extend the results in c to the cases r>1r>1 and additionally showed that one can recover any linear perturbation qLp(Ω)q\in L^{p}(\Omega) with pp satisfying the restrictions (1.6). Furthermore, by this improved Runge approximation theorem the authors could also treat the case of serially or asymptotically polyhomogeneous nonlinearities (see [LTZ24a, Theorem 1.5]).

In this context, let us also mention the recent article [FY24] which deals with the Calderón problem for a third order semilinear, nonlocal, viscous wave equation.

The goal of this paper us to present an extension of the models described in b and c, which we discuss next. Let 0<s<10<s<1 and suppose that we have given coefficients (γ,q)C0,α(n)×Lp(Ω)(\gamma,q)\in C^{0,\alpha}({\mathbb{R}}^{n})\times L^{p}(\Omega), where 0<s<α10<s<\alpha\leq 1 and 1p1\leq p\leq\infty satisfies the restrictions in (1.6). Then we define the following damped, nonlocal wave operator

(1.8) Lγ:=t2+γt+(Δ)sL_{\gamma}\vcentcolon=\partial_{t}^{2}+\gamma\partial_{t}+(-\Delta)^{s}

and consider the problem

(1.9) {Lγu+f(u)=F in ΩT,u=φ on (Ωe)T,u(0)=u0,tu(0)=u1 on Ω,\begin{cases}L_{\gamma}u+f(u)=F&\text{ in }\Omega_{T},\\ u=\varphi&\text{ on }(\Omega_{e})_{T},\\ u(0)=u_{0},\,\partial_{t}u(0)=u_{1}&\text{ on }\Omega,\end{cases}

where ff is a weak nonlinearity or f(u)=q(x)uf(u)=q(x)u. In fact, this is a possibly nonlinear generalization of the model (G2) with s1=0s_{1}=0 in [Zim24, Section 1.1]. By [LTZ24c, Proposition 3.7] (see Section 2.1 for the linear case), we know that the problem (1.9) is well-posed, whenever the source FF, exterior condition φ\varphi and initial conditions u0,u1u_{0},u_{1} are sufficiently regular. Thus, we can introduce the related (partial) DN map via

(1.10) Λγ,fφ:=(Δ)suφ|(W2)T,\Lambda_{\gamma,f}\varphi\vcentcolon=\left.(-\Delta)^{s}u_{\varphi}\right|_{(W_{2})_{T}},

where W1,W2ΩeW_{1},W_{2}\subset\Omega_{e} are some measurement sets, φ\varphi is supported in (W1)T(W_{1})_{T} and uφu_{\varphi} is the unique solution of (1.9) with u0=u1=0u_{0}=u_{1}=0. Then we ask the following question:

Question 2.

Does the partial DN map Λγ,f\Lambda_{\gamma,f} uniquely determine the damping coefficient γ\gamma and the function ff?

In this work we establish the following affirmative answers to this question, whereas the first result discusses the linear case and the second one the semilinear perturbations.

Theorem 1.2 (Uniqueness for linear perturbations).

Let Ωn\Omega\subset{\mathbb{R}}^{n} be a bounded Lipschitz domain, T>0T>0, 0<s<α10<s<\alpha\leq 1 and suppose that 1p1\leq p\leq\infty satisfies (1.6). Assume that for j=1,2j=1,2 we have given coefficients (γj,qj)C0,α(n)×Lp(Ω)(\gamma_{j},q_{j})\in C^{0,\alpha}({\mathbb{R}}^{n})\times L^{p}(\Omega) and let Λγj,qj\Lambda_{\gamma_{j},q_{j}} be the DN map associated to the problem

(1.11) {(Lγj+qj)u=0 in ΩT,u=φ on (Ωe)T,u(0)=0,tu(0)=0 on Ω\begin{cases}(L_{\gamma_{j}}+q_{j})u=0&\text{ in }\Omega_{T},\\ u=\varphi&\text{ on }(\Omega_{e})_{T},\\ u(0)=0,\,\partial_{t}u(0)=0&\text{ on }\Omega\end{cases}

for j=1,2j=1,2. If W1,W2ΩeW_{1},W_{2}\subset\Omega_{e} are two measurement sets such that

(1.12) Λγ1,q1φ|(W2)T=Λγ2,q2φ|(W2)T\left.\Lambda_{\gamma_{1},q_{1}}\varphi\right|_{(W_{2})_{T}}=\left.\Lambda_{\gamma_{2},q_{2}}\varphi\right|_{(W_{2})_{T}}

for all φCc((W1)T)\varphi\in C_{c}^{\infty}((W_{1})_{T}), then there holds

(1.13) γ1=γ2 and q1=q2 in Ω.\gamma_{1}=\gamma_{2}\text{ and }q_{1}=q_{2}\text{ in }\Omega.
Theorem 1.3 (Uniqueness for semilinear perturbations).

Let Ωn\Omega\subset{\mathbb{R}}^{n} be a bounded Lipschitz domain, T>0T>0 and 0<s<α10<s<\alpha\leq 1. Assume that for j=1,2j=1,2 we have given coefficients γjC0,α(n)\gamma_{j}\in C^{0,\alpha}({\mathbb{R}}^{n}) and r+1r+1 homogeneous, weak nonlinearities fjf_{j}, where r>0r>0 satisfies (1.7). Let Λγj,fj\Lambda_{\gamma_{j},f_{j}} be the DN map associated to the problem

(1.14) {Lγju+fj(u)=0 in ΩT,u=φ on (Ωe)T,u(0)=0,tu(0)=0 on Ω\begin{cases}L_{\gamma_{j}}u+f_{j}(u)=0&\text{ in }\Omega_{T},\\ u=\varphi&\text{ on }(\Omega_{e})_{T},\\ u(0)=0,\,\partial_{t}u(0)=0&\text{ on }\Omega\end{cases}

for j=1,2j=1,2. If W1,W2ΩeW_{1},W_{2}\subset\Omega_{e} are two measurement sets such that

(1.15) Λγ1,f1φ|(W2)T=Λγ2,f2φ|(W2)T\left.\Lambda_{\gamma_{1},f_{1}}\varphi\right|_{(W_{2})_{T}}=\left.\Lambda_{\gamma_{2},f_{2}}\varphi\right|_{(W_{2})_{T}}

for all φCc((W1)T)\varphi\in C_{c}^{\infty}((W_{1})_{T}), then there holds

(1.16) γ1=γ2 in Ω and f1=f2 in Ω×.\gamma_{1}=\gamma_{2}\text{ in }\Omega\text{ and }f_{1}=f_{2}\text{ in }\Omega\times{\mathbb{R}}.
Remark 1.4.

For simplicity we restrict our attention to homogeneous nonlinearities ff, but the unique determination remains valid in some polyhomogeneous cases as described in [LTZ24a] for γ=0\gamma=0.

1.2. Organization of the article

The rest of this article is structured as follows. In Section 2, we establish the existence of unique weak and very weak solutions to damped nonlocal wave equations. In Section 3 we then move on to the inverse problem part of this work. First, in Section 3.1 we establish the optimal Runge approximation theorem. Afterwards, in Section 3.2 we prove a suitable integral identity that allows us to recover simultaneously the damping coefficient γ\gamma and potential qq in Section 3.3. Finally, Section 3.4 contains the proof of the simultaneous determination of the damping coefficient γ\gamma and the homogeneous nonlinearity ff.

2. Weak and very weak solutions to damped, nonlocal wave equations

The main purpose of this section is to show existence of unique weak and very weak solutions to damped, nonlocal wave equations (DNWEQ)

(2.1) {(Lγ+q)u=F in ΩT,u=φ on (Ωe)T,u(0)=u0,tu(0)=u1 on Ω,\begin{cases}(L_{\gamma}+q)u=F&\text{ in }\Omega_{T},\\ u=\varphi&\text{ on }(\Omega_{e})_{T},\\ u(0)=u_{0},\,\partial_{t}u(0)=u_{1}&\text{ on }\Omega,\end{cases}

where LγL_{\gamma} is given by (1.8) and only the case φ=0\varphi=0 is considered for very weak solutions.

2.1. Weak solutions

This section deals with the well-posedness of (2.1) for regular sources, exterior conditions and initial data. We also prove well-posedness for the case when instead of initial values the values at t=Tt=T are specified, which will be needed for the development of the theory of very weak solutions.

Theorem 2.1 (Weak solutions to homogeneous DNWEQ).

Let Ωn\Omega\subset{\mathbb{R}}^{n} be a bounded Lipschitz domain, T>0T>0, 0<s<10<s<1 and suppose that 1p1\leq p\leq\infty satisfies (1.6). Assume that we have given coefficients (γ,q)L(Ω)×Lp(Ω)(\gamma,q)\in L^{\infty}(\Omega)\times L^{p}(\Omega). Then for any FL2(0,T;L~2(Ω))F\in L^{2}(0,T;\widetilde{L}^{2}(\Omega))111Here and below we set L~2(Ω):=H~0(Ω)\widetilde{L}^{2}(\Omega)\vcentcolon=\widetilde{H}^{0}(\Omega). and initial conditions (u0,u1)H~s(Ω)×L~2(Ω)(u_{0},u_{1})\in\widetilde{H}^{s}(\Omega)\times\widetilde{L}^{2}(\Omega), there exists a unique weak solution uC([0,T];H~s(Ω))C1([0,T];L~2(Ω))u\in C([0,T];\widetilde{H}^{s}(\Omega))\cap C^{1}([0,T];\widetilde{L}^{2}(\Omega)) of

(2.2) {(Lγ+q)u=F in ΩT,u=0 on (Ωe)T,u(0)=u0,tu(0)=u1 on Ω,\begin{cases}(L_{\gamma}+q)u=F&\text{ in }\Omega_{T},\\ u=0&\text{ on }(\Omega_{e})_{T},\\ u(0)=u_{0},\,\partial_{t}u(0)=u_{1}&\text{ on }\Omega,\end{cases}

which means that (u(0),tu(0))=(u0,u1)(u(0),\partial_{t}u(0))=(u_{0},u_{1}) in H~s(Ω)×L~2(Ω)\widetilde{H}^{s}(\Omega)\times\widetilde{L}^{2}(\Omega) and there holds

(2.3) ddttu,vL2(Ω)+γtu,vL2(Ω)+(Δ)s/2u,(Δ)s/2vL2(n)+qu,vL2(Ω)=F,vL2(Ω)\begin{split}&\frac{d}{dt}\langle\partial_{t}u,v\rangle_{L^{2}(\Omega)}+\langle\gamma\partial_{t}u,v\rangle_{L^{2}(\Omega)}+\langle(-\Delta)^{s/2}u,(-\Delta)^{s/2}v\rangle_{L^{2}({\mathbb{R}}^{n})}+\langle qu,v\rangle_{L^{2}(\Omega)}\\ &=\langle F,v\rangle_{L^{2}(\Omega)}\end{split}

for all vH~s(Ω)v\in\widetilde{H}^{s}(\Omega) in the sense of distributions on (0,T)(0,T). Moreover, the unique solution uu obeys the energy identity

(2.4) tu(t)L2(Ω)2+(Δ)s/2u(t)L2(n)2+20tγtu(τ)+qu(τ),tu(τ)L2(Ω)𝑑τ=(Δ)s/2u0L2(n)2+u1L2(Ω)2+20tF(τ),tu(τ)L2(Ω)𝑑τ\begin{split}&\|\partial_{t}u(t)\|_{L^{2}(\Omega)}^{2}+\|(-\Delta)^{s/2}u(t)\|_{L^{2}({\mathbb{R}}^{n})}^{2}+2\int_{0}^{t}\langle\gamma\partial_{t}u(\tau)+qu(\tau),\partial_{t}u(\tau)\rangle_{L^{2}(\Omega)}\,d\tau\\ &=\|(-\Delta)^{s/2}u_{0}\|_{L^{2}({\mathbb{R}}^{n})}^{2}+\|u_{1}\|_{L^{2}(\Omega)}^{2}+2\int_{0}^{t}\langle F(\tau),\partial_{t}u(\tau)\rangle_{L^{2}(\Omega)}\,d\tau\end{split}

which implies

(2.5) tu(t)L2(Ω)+(Δ)s/2u(t)L2(n)C(u1L2(Ω)+(Δ)s/2u0L2(n)+FL2(0,t;L2(Ω)))\begin{split}&\|\partial_{t}u(t)\|_{L^{2}(\Omega)}+\|(-\Delta)^{s/2}u(t)\|_{L^{2}({\mathbb{R}}^{n})}\\ &\leq C(\|u_{1}\|_{L^{2}(\Omega)}+\|(-\Delta)^{s/2}u_{0}\|_{L^{2}({\mathbb{R}}^{n})}+\|F\|_{L^{2}(0,t;L^{2}(\Omega))})\end{split}

for all 0tT0\leq t\leq T and some C>0C>0 only depending on qLp(Ω)\|q\|_{L^{p}(\Omega)}, γL(Ω)\|\gamma\|_{L^{{\infty}(\Omega)}} and T>0T>0.

Proof.

Throughout the proof, we endow H~s(Ω)\widetilde{H}^{s}(\Omega) with the equivalent norm uH~s(Ω)=(Δ)s/2uL2(n)\|u\|_{\widetilde{H}^{s}(\Omega)}=\|(-\Delta)^{s/2}u\|_{L^{2}({\mathbb{R}}^{n})} (see [LTZ24c, Lemma 2.3]) and we introduce the following continuous sesquilinear forms

(2.6) a0(u,v)=(Δ)s/2u,(Δ)s/2vL2(n),a1(u,v)=qu,vL2(Ω)a_{0}(u,v)=\langle(-\Delta)^{s/2}u,(-\Delta)^{s/2}v\rangle_{L^{2}({\mathbb{R}}^{n})},\quad a_{1}(u,v)=\langle qu,v\rangle_{L^{2}(\Omega)}

for u,vH~s(Ω)u,v\in\widetilde{H}^{s}(\Omega) and

(2.7) b(u,v)=γu,vL2(Ω)b(u,v)=\langle\gamma u,v\rangle_{L^{2}(\Omega)}

for u,vL~2(Ω)u,v\in\widetilde{L}^{2}(\Omega). Next, recall that by [LTZ24c, eq. (3.7)] one has

(2.8) quL2(Ω)CqLp(Ω)uH~s(Ω)\|qu\|_{L^{2}(\Omega)}\leq C\|q\|_{L^{p}(\Omega)}\|u\|_{\widetilde{H}^{s}(\Omega)}

for all uH~s(Ω)u\in\widetilde{H}^{s}(\Omega). It is not hard to see that we can invoke the existence and uniqueness results [DL92, Chapter XVIII, §5, Theorem 3 & 4] (see [DL92, p. 571]), which ensure the existence of a unique, real-valued solution uC([0,T];H~s(Ω))C1([0,T];L~2(Ω))u\in C([0,T];\widetilde{H}^{s}(\Omega))\cap C^{1}([0,T];\widetilde{L}^{2}(\Omega)) to (2.2). Furthermore, by [DL92, Chapter XVIII, §5, Lemma 7] the solution uu satisfies the following energy identity

(2.9) tu(t)L2(Ω)2+(Δ)s/2u(t)L2(n)2+20tγtu(τ)+qu(τ),tu(τ)L2(Ω)𝑑τ=(Δ)s/2u0L2(n)2+u1L2(Ω)2+20tF(τ),tu(τ)L2(Ω)𝑑τ\begin{split}&\|\partial_{t}u(t)\|_{L^{2}(\Omega)}^{2}+\|(-\Delta)^{s/2}u(t)\|_{L^{2}({\mathbb{R}}^{n})}^{2}+2\int_{0}^{t}\langle\gamma\partial_{t}u(\tau)+qu(\tau),\partial_{t}u(\tau)\rangle_{L^{2}(\Omega)}\,d\tau\\ &=\|(-\Delta)^{s/2}u_{0}\|_{L^{2}({\mathbb{R}}^{n})}^{2}+\|u_{1}\|_{L^{2}(\Omega)}^{2}+2\int_{0}^{t}\langle F(\tau),\partial_{t}u(\tau)\rangle_{L^{2}(\Omega)}\,d\tau\end{split}

for 0tT0\leq t\leq T. Hence, we have shown the identity (2.4). Let us define ΨC([0,T])\Psi\in C([0,T]) by

Ψ(t):=tu(t)L2(Ω)2+(Δ)s/2u(t)L2(n)2\Psi(t)\vcentcolon=\|\partial_{t}u(t)\|_{L^{2}(\Omega)}^{2}+\|(-\Delta)^{s/2}u(t)\|_{L^{2}({\mathbb{R}}^{n})}^{2}

for 0tT0\leq t\leq T. Using (2.8), (2.9) and γL(Ω)\gamma\in L^{\infty}(\Omega), we get

Ψ(t)Ψ(0)+0tF(τ)L2(Ω)2𝑑τ+C0t(1+γL(Ω)+qLp(Ω)2)Ψ(τ)𝑑τ\begin{split}&\Psi(t)\leq\Psi(0)+\int_{0}^{t}\|F(\tau)\|_{L^{2}(\Omega)}^{2}\,d\tau+C\int_{0}^{t}(1+\|\gamma\|_{L^{\infty}(\Omega)}+\|q\|^{2}_{L^{p}(\Omega)})\Psi(\tau)\,d\tau\end{split}

and via Gronwall’s inequality we deduce the energy estimate

(2.10) Ψ(t)C(Ψ(0)+FL2(0,t;L2(Ω))2)\begin{split}&\Psi(t)\leq C(\Psi(0)+\|F\|_{L^{2}(0,t;L^{2}(\Omega))}^{2})\end{split}

for all 0tT0\leq t\leq T and some C>0C>0 only depending on qLp(Ω)\|q\|_{L^{p}(\Omega)}, γL(Ω)\|\gamma\|_{L^{\infty}(\Omega)} and T>0T>0. This establishes the estimate (2.5). ∎

As a consequence we have the following result:

Proposition 2.2 (Weak solutions to inhomogeneous DNWEQ).

Let Ωn\Omega\subset{\mathbb{R}}^{n} be a bounded Lipschitz domain, T>0T>0, 0<s<10<s<1 and suppose that 1p1\leq p\leq\infty satisfies (1.6). Assume that we have given coefficients (γ,q)L(Ω)×Lp(Ω)(\gamma,q)\in L^{\infty}(\Omega)\times L^{p}(\Omega). Then for any FL2(0,T;L~2(Ω))F\in L^{2}(0,T;\widetilde{L}^{2}(\Omega)), exterior condition φC2([0,T];H2s(n))\varphi\in C^{2}([0,T];H^{2s}({\mathbb{R}}^{n})) and initial conditions (u0,u1)Hs(n)×L2(n)(u_{0},u_{1})\in H^{s}({\mathbb{R}}^{n})\times L^{2}({\mathbb{R}}^{n}) satisfying the compatibility conditions u0φ(0)H~s(Ω)u_{0}-\varphi(0)\in\widetilde{H}^{s}(\Omega) and u1tφ(0)L~2(Ω)u_{1}-\partial_{t}\varphi(0)\in\widetilde{L}^{2}(\Omega), there exists a unique weak solution uC([0,T];Hs(n))C1([0,T];L2(n))u\in C([0,T];H^{s}({\mathbb{R}}^{n}))\cap C^{1}([0,T];L^{2}({\mathbb{R}}^{n})) of

(2.11) {(Lγ+q)u=F in ΩT,u=φ on (Ωe)T,u(0)=u0,tu(0)=u1 on Ω,\begin{cases}(L_{\gamma}+q)u=F&\text{ in }\Omega_{T},\\ u=\varphi&\text{ on }(\Omega_{e})_{T},\\ u(0)=u_{0},\,\partial_{t}u(0)=u_{1}&\text{ on }\Omega,\end{cases}

which means that uu satisfies (2.3), the (u(0),tu(0))=(u0,u1)(u(0),\partial_{t}u(0))=(u_{0},u_{1}) in Hs(n)×L2(n)H^{s}({\mathbb{R}}^{n})\times L^{2}({\mathbb{R}}^{n}) and u=φu=\varphi in (Ωe)T(\Omega_{e})_{T} means that u(t)=φ(t)u(t)=\varphi(t) a.e. in Ωe\Omega_{e} for any 0<t<T0<t<T. Furthermore, the following energy estimate holds

(2.12) tu(t)L2(n)+(Δ)s/2u(t)L2(n)C(u1L2(n)+(Δ)s/2u0L2(n)+φC2([0,t];H2s(n))+FL2(0,t;L2(Ω)))\begin{split}&\|\partial_{t}u(t)\|_{L^{2}({\mathbb{R}}^{n})}+\|(-\Delta)^{s/2}u(t)\|_{L^{2}({\mathbb{R}}^{n})}\\ &\leq C(\|u_{1}\|_{L^{2}({\mathbb{R}}^{n})}+\|(-\Delta)^{s/2}u_{0}\|_{L^{2}({\mathbb{R}}^{n})}+\|\varphi\|_{C^{2}([0,t];H^{2s}({\mathbb{R}}^{n}))}+\|F\|_{L^{2}(0,t;L^{2}(\Omega))})\end{split}

for any 0tT0\leq t\leq T.

Proof.

Observe, under the current regularity assumptions and compatibility conditions, that uu solves (2.11) if and only if w:=uφw\vcentcolon=u-\varphi solves

(2.13) {(Lγ+q)w=F(Lγ+q)φ in ΩT,w=0 on (Ωe)T,w(0)=u0φ(0),tw(0)=u1tφ(0) on Ω.\begin{cases}(L_{\gamma}+q)w=F-(L_{\gamma}+q)\varphi&\text{ in }\Omega_{T},\\ w=0&\text{ on }(\Omega_{e})_{T},\\ w(0)=u_{0}-\varphi(0),\,\partial_{t}w(0)=u_{1}-\partial_{t}\varphi(0)&\text{ on }\Omega.\end{cases}

The only fact to keep in mind is that if uC([0,T];Hs(n))u\in C([0,T];H^{s}({\mathbb{R}}^{n})), then the condition u(t)=φ(t)u(t)=\varphi(t) a.e. in Ωe\Omega_{e} is equivalent to u(t)φ(t)H~s(Ω)u(t)-\varphi(t)\in\widetilde{H}^{s}(\Omega) as Ωn\Omega\subset{\mathbb{R}}^{n} is a bounded Lipschitz domain. So, the assertions of Propsition 2.2 follow immediately from Theorem 2.1. ∎

Next, let us define for any gLloc1(VT)g\in L^{1}_{loc}(V_{T}), VnV\subset{\mathbb{R}}^{n} open, its time-reversal

(2.14) g(x,t)=g(x,Tt).g^{\star}(x,t)=g(x,T-t).

Then, we have the following lemma.

Lemma 2.3.

Let Ωn\Omega\subset{\mathbb{R}}^{n} be a bounded Lipschitz domain, T>0T>0, 0<s<10<s<1 and suppose that 1p1\leq p\leq\infty satisfies (1.6). Assume that we have given coefficients (γ,q)L(Ω)×Lp(Ω)(\gamma,q)\in L^{\infty}(\Omega)\times L^{p}(\Omega). Let FL2(0,T;L~2(Ω))F\in L^{2}(0,T;\widetilde{L}^{2}(\Omega)), φC2([0,T];H2s(n))\varphi\in C^{2}([0,T];H^{2s}({\mathbb{R}}^{n})) and (u0,u1)Hs(n)×L2(n)(u_{0},u_{1})\in H^{s}({\mathbb{R}}^{n})\times L^{2}({\mathbb{R}}^{n}) satisfying the compatibility conditions u0φ(0)H~s(Ω)u_{0}-\varphi(0)\in\widetilde{H}^{s}(\Omega) and u1tφ(0)L~2(Ω)u_{1}-\partial_{t}\varphi(0)\in\widetilde{L}^{2}(\Omega). Then uu solves

(2.15) {(Lγ+q)u=F in ΩT,u=φ on (Ωe)T,u(0)=u0,tu(0)=u1 on Ω,\begin{cases}(L_{\gamma}+q)u=F&\text{ in }\Omega_{T},\\ u=\varphi&\text{ on }(\Omega_{e})_{T},\\ u(0)=u_{0},\,\partial_{t}u(0)=u_{1}&\text{ on }\Omega,\end{cases}

if and only if uu^{\star} solves

(2.16) {(Lγ+q)v=F in ΩT,v=φ on (Ωe)T,v(T)=u0,tv(T)=u1 on Ω.\begin{cases}(L_{-\gamma}+q)v=F^{\star}&\text{ in }\Omega_{T},\\ v=\varphi^{\star}&\text{ on }(\Omega_{e})_{T},\\ v(T)=u_{0},\,\partial_{t}v(T)=u_{1}&\text{ on }\Omega.\end{cases}

In particular, for any FL2(ΩT)F\in L^{2}(\Omega_{T}), φC2([0,T];H2s(n))\varphi\in C^{2}([0,T];H^{2s}({\mathbb{R}}^{n})) and (u0,u1)Hs(n)×L2(n)(u_{0},u_{1})\in H^{s}({\mathbb{R}}^{n})\times L^{2}({\mathbb{R}}^{n}) satisfying the compatibility conditions u0φ(T)H~s(Ω)u_{0}-\varphi(T)\in\widetilde{H}^{s}(\Omega) and u1tφ(T)L~2(Ω)u_{1}-\partial_{t}\varphi(T)\in\widetilde{L}^{2}(\Omega), there exists a unique solution uu^{\star} of

(2.17) {(Lγ+q)v=F in ΩT,v=φ on (Ωe)T,v(T)=u0,tv(T)=u1 on Ω.\begin{cases}(L_{-\gamma}+q)v=F&\text{ in }\Omega_{T},\\ v=\varphi&\text{ on }(\Omega_{e})_{T},\\ v(T)=u_{0},\,\partial_{t}v(T)=u_{1}&\text{ on }\Omega.\end{cases}
Proof.

First, note that by the proof of Proposition 2.2, we can assume without loss of generality that φ=0\varphi=0. Secondly, one easily sees that tu=(tu)\partial_{t}u^{\star}=-(\partial_{t}u)^{\star} and thus a change of variables in (2.3) gives the asserted equivalence. The unique solvability of (2.17) follows from the equivalence and Theorem 2.1. ∎

2.2. Very weak solutions

Let us start by making some simple observations. Suppose that uu and vv are smooth solutions of the problems

(2.18) {(Lγ+q)u=F in ΩT,u=0 on (Ωe)T,u(0)=u0,tu(0)=u1 on Ω,\begin{cases}(L_{\gamma}+q)u=F&\text{ in }\Omega_{T},\\ u=0&\text{ on }(\Omega_{e})_{T},\\ u(0)=u_{0},\,\partial_{t}u(0)=u_{1}&\text{ on }\Omega,\end{cases}

and

(2.19) {(Lγ+q)v=G in ΩT,v=0 on (Ωe)T,v(T)=0,tv(T)=0 on Ω,\begin{cases}(L_{-\gamma}+q)v=G&\text{ in }\Omega_{T},\\ v=0&\text{ on }(\Omega_{e})_{T},\\ v(T)=0,\,\partial_{t}v(T)=0&\text{ on }\Omega,\end{cases}

respectively. If we multiply the PDE (2.18) by vv and integrate over ΩT\Omega_{T}, then we get

(2.20) ΩTFv𝑑x𝑑t=ΩT[(Lγ+q)u]v𝑑x𝑑t=Ωu0tv(0)dxΩu1v(0)𝑑xΩγu0v(0)𝑑x+ΩTu(Lγ+q)v𝑑x𝑑t=Ωu0tv(0)dxΩu1v(0)𝑑xΩγu0v(0)𝑑x+ΩTGu𝑑x𝑑t.\begin{split}&\int_{\Omega_{T}}Fv\,dxdt=\int_{\Omega_{T}}[(L_{\gamma}+q)u]v\,dxdt\\ &=\int_{\Omega}u_{0}\partial_{t}v(0)\,dx-\int_{\Omega}u_{1}v(0)\,dx-\int_{\Omega}\gamma u_{0}v(0)\,dx+\int_{\Omega_{T}}u(L_{-\gamma}+q)v\,dxdt\\ &=\int_{\Omega}u_{0}\partial_{t}v(0)\,dx-\int_{\Omega}u_{1}v(0)\,dx-\int_{\Omega}\gamma u_{0}v(0)\,dx+\int_{\Omega_{T}}Gu\,dxdt.\end{split}

Notice that if GL2(0,T;L~2(Ω))G\in L^{2}(0,T;\widetilde{L}^{2}(\Omega)), vL2(0,T;H~s(Ω))v\in L^{2}(0,T;\widetilde{H}^{s}(\Omega)) and (v(0),tv(0))H~s(Ω)×L~2(Ω)(v(0),\partial_{t}v(0))\in\widetilde{H}^{s}(\Omega)\times\widetilde{L}^{2}(\Omega), then one can make sense of the first integral and the last line in (2.20), even in the case FL2(0,T;Hs(Ω))F\in L^{2}(0,T;H^{-s}(\Omega)), (u0,u1)L~2(Ω)×Hs(Ω)(u_{0},u_{1})\in\widetilde{L}^{2}(\Omega)\times H^{-s}(\Omega) and uL2(0,T;L~2(Ω))u\in L^{2}(0,T;\widetilde{L}^{2}(\Omega)). Here, Hs(Ω)𝒟(Ω)H^{-s}(\Omega)\subset\mathscr{D}^{\prime}(\Omega) is defined by

Hs(Ω)={u|Ω;uHs(n)}H^{-s}(\Omega)=\{u|_{\Omega}\,;\,u\in H^{-s}({\mathbb{R}}^{n})\}

and it can be identified with the dual space of H~s(Ω)\widetilde{H}^{s}(\Omega), when Ω\Omega is Lipschitz. The previous computation motivates the following definition.

Definition 2.4 (Very weak solutions).

Let Ωn\Omega\subset{\mathbb{R}}^{n} be a bounded Lipschitz domain, T>0T>0, 0<s<10<s<1 and suppose that 1p1\leq p\leq\infty satisfies (1.6). Assume that we have given coefficients (γ,q)L(Ω)×Lp(Ω)(\gamma,q)\in L^{\infty}(\Omega)\times L^{p}(\Omega), source FL2(0,T;Hs(Ω))F\in L^{2}(0,T;H^{-s}(\Omega)) and initial conditions (u0,u1)L~2(Ω)×Hs(Ω)(u_{0},u_{1})\in\widetilde{L}^{2}(\Omega)\times H^{-s}(\Omega). Then we say that uC([0,T];L~2(Ω))C1([0,T];Hs(Ω))u\in C([0,T];\widetilde{L}^{2}(\Omega))\cap C^{1}([0,T];H^{-s}(\Omega)) is a very weak solution of

(2.21) {(Lγ+q)u=F in ΩT,u=0 on (Ωe)T,u(0)=u0,tu(0)=u1 on Ω,\begin{cases}(L_{\gamma}+q)u=F&\text{ in }\Omega_{T},\\ u=0&\text{ on }(\Omega_{e})_{T},\\ u(0)=u_{0},\,\partial_{t}u(0)=u_{1}&\text{ on }\Omega,\end{cases}

whenever there holds222Here and below we sometimes write ,\langle\cdot,\cdot\rangle to denote the duality pairing between Hs(Ω)×H~s(Ω)H^{-s}(\Omega)\times\widetilde{H}^{s}(\Omega).

(2.22) 0TG,uL2(Ω)𝑑t=0TF,v𝑑t+u1,v(0)u0,tv(0)L2(Ω)+γu0,v(0)\int_{0}^{T}\langle G,u\rangle_{L^{2}(\Omega)}\,dt=\int_{0}^{T}\langle F,v\rangle\,dt+\langle u_{1},v(0)\rangle-\langle u_{0},\partial_{t}v(0)\rangle_{L^{2}(\Omega)}+\langle\gamma u_{0},v(0)\rangle

for all GL2(0,T;L~2(Ω))G\in L^{2}(0,T;\widetilde{L}^{2}(\Omega)), where vC([0,T];H~s(Ω))C1([0,T];L~2(Ω))v\in C([0,T];\widetilde{H}^{s}(\Omega))\cap C^{1}([0,T];\widetilde{L}^{2}(\Omega)) is the unique weak solution of the adjoint equation

(2.23) {(Lγ+q)v=G in ΩT,v=0 on (Ωe)T,v(T)=0,tv(T)=0 on Ω,\begin{cases}(L_{-\gamma}+q)v=G&\text{ in }\Omega_{T},\\ v=0&\text{ on }(\Omega_{e})_{T},\\ v(T)=0,\,\partial_{t}v(T)=0&\text{ on }\Omega,\end{cases}

(see Theorem 2.1).

Next, let us recall the following well-posedness result of very weak solutions.

Theorem 2.5 (Very weak solutions for γ=q=0\gamma=q=0,[LTZ24a, Theorem 3.6]).

Let Ωn\Omega\subset{\mathbb{R}}^{n} be a bounded Lipschitz domain, T>0T>0 and 0<s<10<s<1. Then for any given source FL2(0,T;Hs(Ω))F\in L^{2}(0,T;H^{-s}(\Omega)) and initial conditions (u0,u1)L~2(Ω)×Hs(Ω)(u_{0},u_{1})\in\widetilde{L}^{2}(\Omega)\times H^{-s}(\Omega), there exists a unique solution to

(2.24) {(t2+(Δ)s)u=F in ΩT,u=0 on (Ωe)T,u(0)=u0,tu(0)=u1 on Ω\begin{cases}(\partial_{t}^{2}+(-\Delta)^{s})u=F&\text{ in }\Omega_{T},\\ u=0&\text{ on }(\Omega_{e})_{T},\\ u(0)=u_{0},\,\partial_{t}u(0)=u_{1}&\text{ on }\Omega\end{cases}

and it satisfies the following energy estimate

(2.25) u(t)L2(Ω)+tu(t)Hs(Ω)C(u0L2(Ω)+u1Hs(Ω)+FL2(0,t;Hs(Ω)))\|u(t)\|_{L^{2}(\Omega)}+\|\partial_{t}u(t)\|_{H^{-s}(\Omega)}\leq C(\|u_{0}\|_{L^{2}(\Omega)}+\|u_{1}\|_{H^{-s}(\Omega)}+\|F\|_{L^{2}(0,t;H^{-s}(\Omega))})

for all 0tT0\leq t\leq T.

Hence, we have a well-defined solution map.

Proposition 2.6 (Solution map).

Let Ωn\Omega\subset{\mathbb{R}}^{n} be a bounded Lipschitz domain, T>0T>0, 0<s<10<s<1 and let Xs:=L~2(Ω)×Hs(Ω)X_{s}\vcentcolon=\widetilde{L}^{2}(\Omega)\times H^{-s}(\Omega) be endowed with the usual product norm

(u,w)Xs:=(uL2(Ω)2+wHs(Ω)2)1/2.\|(u,w)\|_{X_{s}}\vcentcolon=(\|u\|_{L^{2}(\Omega)}^{2}+\|w\|_{H^{-s}(\Omega)}^{2})^{1/2}.

Then the solution map S:L2(0,T;Hs(Ω))C([0,T];Xs)S\colon L^{2}(0,T;H^{-s}(\Omega))\to C([0,T];X_{s}) defined by

(2.26) S(F):=(u,tu),S(F)\vcentcolon=(u,\partial_{t}u),

where uC([0,T];L~2(Ω))C1([0,T];Hs(Ω))u\in C([0,T];\widetilde{L}^{2}(\Omega))\cap C^{1}([0,T];H^{-s}(\Omega)) is the unique solution of (2.24) with (u0,u1)=0(u_{0},u_{1})=0. Moreover, the solution map is continuous and satisfies the estimate

(2.27) S(F)(t)XsCFL2(0,t;Hs(Ω))\|S(F)(t)\|_{X_{s}}\leq C\|F\|_{L^{2}(0,t;H^{-s}(\Omega))}

for any 0tT0\leq t\leq T.

Proof.

First of all note that the solution map SS is well-defined by Theorem 2.5. The estimate (2.27) follows from (2.25), which together with the linearity of SS gives the continuity of SS. Observe that the linearity of SS is a direct consequence of the unique solvability of (2.24) and the fact that the PDE is linear. ∎

Theorem 2.7.

Let Ωn\Omega\subset{\mathbb{R}}^{n} be a bounded Lipschitz domain, T>0T>0, 0<s<10<s<1 and suppose that :C([0,T];Xs)L2(0,T;Hs(Ω))\mathcal{F}\colon C([0,T];X_{s})\to L^{2}(0,T;H^{-s}(\Omega)) satisfies the Lipschitz estimate

(2.28) (U)(t)(V)(t)Hs(Ω)CU(t)V(t)Xs\|\mathcal{F}(U)(t)-\mathcal{F}(V)(t)\|_{H^{-s}(\Omega)}\leq C\|U(t)-V(t)\|_{X_{s}}

for a.e. 0tT0\leq t\leq T and U,VC([0,T];Xs)U,V\in C([0,T];X_{s}). Then for all (u0,u1)Xs(u_{0},u_{1})\in X_{s}, there exists a unique solution uu of

(2.29) {(t2+(Δ)s)u=(u,tu) in ΩT,u=0 on (Ωe)T,u(0)=u0,tu(0)=u1 on Ω,\begin{cases}(\partial_{t}^{2}+(-\Delta)^{s})u=\mathcal{F}(u,\partial_{t}u)&\text{ in }\Omega_{T},\\ u=0&\text{ on }(\Omega_{e})_{T},\\ u(0)=u_{0},\,\partial_{t}u(0)=u_{1}&\text{ on }\Omega,\end{cases}

that is the formula (2.22) holds with FF replaced by (u,tu)\mathcal{F}(u,\partial_{t}u) in which we test against every weak solution vv of the adjoint equation

(2.30) {(t2+(Δ)s)v=G in ΩT,v=0 on (Ωe)T,v(T)=0,tv(T)=0 on Ω\begin{cases}(\partial_{t}^{2}+(-\Delta)^{s})v=G&\text{ in }\Omega_{T},\\ v=0&\text{ on }(\Omega_{e})_{T},\\ v(T)=0,\,\partial_{t}v(T)=0&\text{ on }\Omega\end{cases}

with GL2(0,T;L~2(Ω))G\in L^{2}(0,T;\widetilde{L}^{2}(\Omega)).

Proof of Theorem 2.7.

Let uhC([0,T];L~2(Ω))C1([0,T];Hs(Ω))u_{h}\in C([0,T];\widetilde{L}^{2}(\Omega))\cap C^{1}([0,T];H^{-s}(\Omega)) be the unique solution to

(2.31) {(t2+(Δ)s)u=0 in ΩT,u=0 on (Ωe)T,u(0)=u0,tu(0)=u1 on Ω\begin{cases}(\partial_{t}^{2}+(-\Delta)^{s})u=0&\text{ in }\Omega_{T},\\ u=0&\text{ on }(\Omega_{e})_{T},\\ u(0)=u_{0},\,\partial_{t}u(0)=u_{1}&\text{ on }\Omega\end{cases}

and let us set Uh:=(uh,tuh)C([0,T];Xs)U_{h}\vcentcolon=(u_{h},\partial_{t}u_{h})\in C([0,T];X_{s}). Furthermore, we define the operator 𝒯:C([0,T];Xs)C([0,T];Xs)\mathcal{T}\colon C([0,T];X_{s})\to C([0,T];X_{s}) as

(2.32) 𝒯(U):=Uh+S((U)),\mathcal{T}(U)\vcentcolon=U_{h}+S(\mathcal{F}(U)),

which is well-defined by (2.6) and the properties of \mathcal{F}. Next, we show that 𝒯\mathcal{T} has a unique fixed point U=(U1,U2)U=(U_{1},U_{2}).

Step 1. Existence. Let U,VC([0,T];Xs)U,V\in C([0,T];X_{s}), then by linearity of SS, (2.27) and (2.28) we get

𝒯(U)(t)𝒯(V)(t)Xs=S((U))(t)S((V))(t)Xs=S((U)(V))(t)XsC(U)(V)L2(0,t;Hs(Ω))CUVL2(0,t;Xs).\begin{split}\|\mathcal{T}(U)(t)-\mathcal{T}(V)(t)\|_{X_{s}}&=\|S(\mathcal{F}(U))(t)-S(\mathcal{F}(V))(t)\|_{X_{s}}\\ &=\|S(\mathcal{F}(U)-\mathcal{F}(V))(t)\|_{X_{s}}\\ &\leq C\|\mathcal{F}(U)-\mathcal{F}(V)\|_{L^{2}(0,t;H^{-s}(\Omega))}\\ &\leq C\|U-V\|_{L^{2}(0,t;X_{s})}.\end{split}

Next, let us define the following norm on XsX_{s}

(2.33) Uθ:=sup0tT(eθtU(t)Xs)\|U\|_{\theta}\vcentcolon=\sup_{0\leq t\leq T}\left(e^{-\theta t}\|U(t)\|_{X_{s}}\right)

for θ>0\theta>0, which will be fixed in a moment. Then we have the estimate

𝒯(U)(t)𝒯(V)(t)XsC(0te2θτ𝑑τ)1/2UVθC(2θ)1/2eθtUVθ\|\mathcal{T}(U)(t)-\mathcal{T}(V)(t)\|_{X_{s}}\leq C\left(\int_{0}^{t}e^{2\theta\tau}\,d\tau\right)^{1/2}\|U-V\|_{\theta}\leq\frac{C}{(2\theta)^{1/2}}e^{\theta t}\|U-V\|_{\theta}

and hence there holds

𝒯(U)(t)𝒯(V)(t)θC(2θ)1/2UVθ.\|\mathcal{T}(U)(t)-\mathcal{T}(V)(t)\|_{\theta}\leq\frac{C}{(2\theta)^{1/2}}\|U-V\|_{\theta}.

Therefore, we deduce that 𝒯\mathcal{T} is a strict contraction from the complete metric space (C([0,T];Xs),θ)(C([0,T];X_{s}),\|\cdot\|_{\theta}) to itself, when θ>0\theta>0 is chosen such that C/(2θ)1/2<1C/(2\theta)^{1/2}<1. Now, we may invoke Banach’s fixed point theorem to obtain a unique fixed point U=(u,w)U=(u,w) of 𝒯\mathcal{T}. Next, observe that the definition of the solution map SS and U=𝒯(U)=Uh+S((U))U=\mathcal{T}(U)=U_{h}+S(\mathcal{F}(U)) imply

u=uh+un and w=tu,u=u_{h}+u_{n}\text{ and }w=\partial_{t}u,

where unu_{n} solves

(2.34) {(t2+(Δ)s)v=(U) in ΩT,v=0 on (Ωe)T,v(0)=0,tv(0)=0 on Ω.\begin{cases}(\partial_{t}^{2}+(-\Delta)^{s})v=\mathcal{F}(U)&\text{ in }\Omega_{T},\\ v=0&\text{ on }(\Omega_{e})_{T},\\ v(0)=0,\,\partial_{t}v(0)=0&\text{ on }\Omega.\end{cases}

Going back to the definition of very weak solutions, we see this implies that uu solves (2.29).

Step 2. Uniqueness. Suppose u~C([0,T];L~2(Ω))C1([0,T];Hs(Ω))\tilde{u}\in C([0,T];\widetilde{L}^{2}(\Omega))\cap C^{1}([0,T];H^{-s}(\Omega)) is any other solution to (2.29), then u¯:=uu~\bar{u}\vcentcolon=u-\widetilde{u} solves

(2.35) {(t2+(Δ)s)v=(u,tu)(u~,tu~) in ΩT,v=0 on (Ωe)T,v(0)=0,tv(0)=0 on Ω.\begin{cases}(\partial_{t}^{2}+(-\Delta)^{s})v=\mathcal{F}(u,\partial_{t}u)-\mathcal{F}(\widetilde{u},\partial_{t}\widetilde{u})&\text{ in }\Omega_{T},\\ v=0&\text{ on }(\Omega_{e})_{T},\\ v(0)=0,\,\partial_{t}v(0)=0&\text{ on }\Omega.\end{cases}

Thus, applying the energy estimate (2.25) together with the Lipschitz assumption on \mathcal{F}, we see that

U(t)U~(t)Xs2C0t(U)(τ)(U~)(τ)Hs(Ω)2𝑑τC0tU(t)U~(t)Xs2𝑑τ,\begin{split}\|U(t)-\widetilde{U}(t)\|^{2}_{X_{s}}&\leq C\int_{0}^{t}\|\mathcal{F}(U)(\tau)-\mathcal{F}(\widetilde{U})(\tau)\|_{H^{-s}(\Omega)}^{2}\,d\tau\\ &\leq C\int_{0}^{t}\|U(t)-\widetilde{U}(t)\|^{2}_{X_{s}}\,d\tau,\end{split}

where U=(u,tu)U=(u,\partial_{t}u) and U~=(u~,tu~)\widetilde{U}=(\widetilde{u},\partial_{t}\widetilde{u}). So, Gronwall’s inequality shows that u=u~u=\widetilde{u}. This establishes the uniqueness assertion and we can conclude the proof. ∎

As an application of Theorem 2.7, we can show the unique solvability of (2.1) for rough source and initial data.

Theorem 2.8 (Very weak solutions to DNWEQ).

Let Ωn\Omega\subset{\mathbb{R}}^{n} be a bounded Lipschitz domain, T>0T>0, 0<s<α10<s<\alpha\leq 1 and suppose that 1p1\leq p\leq\infty satisfies (1.6). Assume that we have given coefficients (γ,q)C0,α(n)×Lp(Ω)(\gamma,q)\in C^{0,\alpha}({\mathbb{R}}^{n})\times L^{p}(\Omega). Then for any FL2(0,T;Hs(Ω))F\in L^{2}(0,T;H^{-s}(\Omega)) and (u0,u1)L~2(Ω)×Hs(Ω)(u_{0},u_{1})\in\widetilde{L}^{2}(\Omega)\times H^{-s}(\Omega), there exists a unique solution of

(2.36) {(Lγ+q)u=F in ΩT,u=0 on (Ωe)T,u(0)=u0,tu(0)=u1 on Ω.\begin{cases}(L_{\gamma}+q)u=F&\text{ in }\Omega_{T},\\ u=0&\text{ on }(\Omega_{e})_{T},\\ u(0)=u_{0},\,\partial_{t}u(0)=u_{1}&\text{ on }\Omega.\end{cases}
Proof.

Let us define the mapping :C([0,T];Xs)L2(0,T;Hs(Ω))\mathcal{F}\colon C([0,T];X_{s})\to L^{2}(0,T;H^{-s}(\Omega)) by

(U)(t):=Fγw(t)qu(t),\mathcal{F}(U)(t)\vcentcolon=F-\gamma w(t)-qu(t),

where U=(u,w)C([0,T];Xs)U=(u,w)\in C([0,T];X_{s}). On the one hand, using the estimate (2.8) we see that for any uL~2(Ω)u\in\widetilde{L}^{2}(\Omega) one has quHs(Ω)qu\in H^{-s}(\Omega) and there holds

(2.37) quHs(Ω)=supvH~s(Ω)1|u,qvL2(Ω)|CqLp(Ω)uL2(Ω).\begin{split}\|qu\|_{H^{-s}(\Omega)}&=\sup_{\|v\|_{\widetilde{H}^{s}(\Omega)}\leq 1}|\langle u,qv\rangle_{L^{2}(\Omega)}|\\ &\leq C\|q\|_{L^{p}(\Omega)}\|u\|_{L^{2}(\Omega)}.\end{split}

On the other hand, by applying [CRTZ24, Lemma 3.1] and ΩC0\partial\Omega\in C^{0} we deduce that for any vH~s(Ω)v\in\widetilde{H}^{s}(\Omega) one has γvH~s(Ω)\gamma v\in\widetilde{H}^{s}(\Omega) and it obeys the estimate

(2.38) γvH~s(Ω)CγC0,α(n)vH~s(Ω).\|\gamma v\|_{\widetilde{H}^{s}(\Omega)}\leq C\|\gamma\|_{C^{0,\alpha}({\mathbb{R}}^{n})}\|v\|_{\widetilde{H}^{s}(\Omega)}.

Thus, we can again infer from a duality argument that Hs(Ω)wγwHs(Ω)H^{-s}(\Omega)\ni w\mapsto\gamma w\in H^{-s}(\Omega) is a continuous map satisfying

(2.39) γwHs(Ω)=supvH~s(Ω)1|γw,v|=supvH~s(Ω)1|w,γv|CγC0,α(n)wHs(Ω).\begin{split}\|\gamma w\|_{H^{-s}(\Omega)}&=\sup_{\|v\|_{\widetilde{H}^{s}(\Omega)}\leq 1}|\langle\gamma w,v\rangle|\\ &=\sup_{\|v\|_{\widetilde{H}^{s}(\Omega)}\leq 1}|\langle w,\gamma v\rangle|\\ &\leq C\|\gamma\|_{C^{0,\alpha}({\mathbb{R}}^{n})}\|w\|_{H^{-s}(\Omega)}.\end{split}

From the estimates (2.37) and (2.39), we easily deduce that \mathcal{F} is well-defined and satisfies the Lipschitz estimate

(2.40) (U)(t)(V)(t)Hs(Ω)C(γC0,α(n)+qLp(Ω))U(t)V(t)Xs\|\mathcal{F}(U)(t)-\mathcal{F}(V)(t)\|_{H^{-s}(\Omega)}\leq C(\|\gamma\|_{C^{0,\alpha}({\mathbb{R}}^{n})}+\|q\|_{L^{p}(\Omega)})\|U(t)-V(t)\|_{X_{s}}

for all U,VC([0,T];Xs)U,V\in C([0,T];X_{s}). Thus, we can apply Theorem 2.7 to get the existence of a unique solution to (2.36) in the sense that for any GL2(0,T;L~2(Ω))G\in L^{2}(0,T;\widetilde{L}^{2}(\Omega)) and corresponding solution vv of (2.30), there holds

(2.41) 0TG,uL2(Ω)𝑑t=0T(Fγtuqu),v𝑑t+u1,v(0)u0,tv(0)L2(Ω).\int_{0}^{T}\langle G,u\rangle_{L^{2}(\Omega)}\,dt=\int_{0}^{T}\langle(F-\gamma\partial_{t}u-qu),v\rangle\,dt+\langle u_{1},v(0)\rangle-\langle u_{0},\partial_{t}v(0)\rangle_{L^{2}(\Omega)}.

It remains to verify that uu is indeed a solution of (2.36) in the sense of Definition 2.4. For this purpose let GL2(0,T;L~2(Ω))G\in L^{2}(0,T;\widetilde{L}^{2}(\Omega)) and suppose that vv is the unique solution to (2.23). Hence, vv solves

{(t2+(Δ)s)v=G~ in ΩT,v=0 on (Ωe)T,v(T)=0,tv(T)=0 on Ω\begin{cases}(\partial_{t}^{2}+(-\Delta)^{s})v=\widetilde{G}&\text{ in }\Omega_{T},\\ v=0&\text{ on }(\Omega_{e})_{T},\\ v(T)=0,\,\partial_{t}v(T)=0&\text{ on }\Omega\end{cases}

with G~=G+γtvqvL2(0,T;L~2(Ω))\widetilde{G}=G+\gamma\partial_{t}v-qv\in L^{2}(0,T;\widetilde{L}^{2}(\Omega)) (see (2.8)). Next, we claim that there holds

(2.42) 0Tγtu,v𝑑t=0Tγtv,uL2(Ω)𝑑tγu0,v(0).\int_{0}^{T}\langle\gamma\partial_{t}u,v\rangle\,dt=-\int_{0}^{T}\langle\gamma\partial_{t}v,u\rangle_{L^{2}(\Omega)}\,dt-\langle\gamma u_{0},v(0)\rangle.

For this purpose, let us consider for ε>0\varepsilon>0 the unique solution vεH1(0,T;H~s(Ω))v_{\varepsilon}\in H^{1}(0,T;\widetilde{H}^{s}(\Omega)) with t2vεL2(0,T;Hs(Ω))\partial_{t}^{2}v_{\varepsilon}\in L^{2}(0,T;H^{-s}(\Omega)) to the following parabolically regularized problem

(2.43) {(t2ε(Δ)st+(Δ)s)v=G~ in ΩT,v=0 in (Ωe)T,v(T)=tv(T)=0 in Ω\begin{cases}(\partial_{t}^{2}-\varepsilon(-\Delta)^{s}\partial_{t}+(-\Delta)^{s})v=\widetilde{G}&\text{ in }\Omega_{T},\\ v=0&\text{ in }(\Omega_{e})_{T},\\ v(T)=\partial_{t}v(T)=0&\text{ in }\Omega\end{cases}

(see [DL92, Chapter XVIII, Section 5.3.1]). By [DL92, Chapter XVIII, Section 5.3.4] we know that there holds

(2.44) vεv in L(0,T;H~s(Ω)),tvεtv in L(0,T;L~2(Ω)),vε(t)v(t) in H~s(Ω) for all 0tT.\begin{split}v_{\varepsilon}&\overset{\ast}{\rightharpoonup}v\text{ in }L^{\infty}(0,T;\widetilde{H}^{s}(\Omega)),\\ \partial_{t}v_{\varepsilon}&\overset{\ast}{\rightharpoonup}\partial_{t}v\text{ in }L^{\infty}(0,T;\widetilde{L}^{2}(\Omega)),\\ v_{\varepsilon}(t)&\to v(t)\text{ in }\widetilde{H}^{s}(\Omega)\text{ for all }0\leq t\leq T.\end{split}

First, note that the conditions uC1([0,T];Hs(Ω))u\in C^{1}([0,T];H^{-s}(\Omega)) and vεC1([0,T];L~2(Ω))v_{\varepsilon}\in C^{1}([0,T];\widetilde{L}^{2}(\Omega)), where the latter follows from the Sobolev embedding, guarantee that γu,vεC1([0,T])\langle\gamma u,v_{\varepsilon}\rangle\in C^{1}([0,T]) with

(2.45) tγu,vε=tu,γvε+u,γtvεL2(Ω).\partial_{t}\langle\gamma u,v_{\varepsilon}\rangle=\langle\partial_{t}u,\gamma v_{\varepsilon}\rangle+\langle u,\gamma\partial_{t}v_{\varepsilon}\rangle_{L^{2}(\Omega)}.

Thus, by the fundamental theorem of calculus we deduce that there holds

γu(T),vε(T)γu0,vε(0)=0Ttu,γvε+u,γtvεL2(Ω)dt.\langle\gamma u(T),v_{\varepsilon}(T)\rangle-\langle\gamma u_{0},v_{\varepsilon}(0)\rangle=\int_{0}^{T}\langle\partial_{t}u,\gamma v_{\varepsilon}\rangle+\langle u,\gamma\partial_{t}v_{\varepsilon}\rangle_{L^{2}(\Omega)}\,dt.

By the convergence assertions (2.44) and vε(T)=0v_{\varepsilon}(T)=0, we get

γu0,v(0)=0Ttu,γv+u,γtvL2(Ω)dt.-\langle\gamma u_{0},v(0)\rangle=\int_{0}^{T}\langle\partial_{t}u,\gamma v\rangle+\langle u,\gamma\partial_{t}v\rangle_{L^{2}(\Omega)}\,dt.

This proves (2.42). Hence, inserting this into (2.41), we obtain

0TG~,uL2(Ω)𝑑t=0T(Fγtuqu),v𝑑t+u1,v(0)u0,tv(0)L2(Ω)=0TF,v𝑑t+0Tu,γtvL2(Ω)𝑑t0Tu,qv𝑑t+u1,v(0)u0,tv(0)L2(Ω)+γu0,v(0).\begin{split}\int_{0}^{T}\langle\widetilde{G},u\rangle_{L^{2}(\Omega)}\,dt&=\int_{0}^{T}\langle(F-\gamma\partial_{t}u-qu),v\rangle\,dt+\langle u_{1},v(0)\rangle-\langle u_{0},\partial_{t}v(0)\rangle_{L^{2}(\Omega)}\\ &=\int_{0}^{T}\langle F,v\rangle\,dt+\int_{0}^{T}\langle u,\gamma\partial_{t}v\rangle_{L^{2}(\Omega)}\,dt-\int_{0}^{T}\langle u,qv\rangle\,dt\\ &\quad+\langle u_{1},v(0)\rangle-\langle u_{0},\partial_{t}v(0)\rangle_{L^{2}(\Omega)}+\langle\gamma u_{0},v(0)\rangle.\end{split}

As G~=G+γtvqv\widetilde{G}=G+\gamma\partial_{t}v-qv this gives

0TG,uL2(Ω)𝑑t=0TF,v𝑑t+u1,v(0)u0,tv(0)L2(Ω)+γu0,v(0).\begin{split}\int_{0}^{T}\langle G,u\rangle_{L^{2}(\Omega)}\,dt&=\int_{0}^{T}\langle F,v\rangle\,dt+\langle u_{1},v(0)\rangle-\langle u_{0},\partial_{t}v(0)\rangle_{L^{2}(\Omega)}+\langle\gamma u_{0},v(0)\rangle.\end{split}

Hence, we observe that uu is indeed a solution of (2.36) in the sense of Definition 2.4. By reversing the above arguments one can also observe that if uu is a solution in the sense of Definition 2.4, then by (2.42) it is a solution in the sense of (2.41) and thus the solution in the sense of Definition 2.4 is unique. ∎

3. The inverse problem

After establishing the theory of very weak solutions to damped, nonlocal wave equations, we now turn our attention to the inverse problem part. First, in Section 3.1 we prove the optimal Runge approximation theorem (Theorem 3.1) and in Section 3.2 a suitable integral identity. Using these results, we then show in Section 3.3 our first main result dealing with linear perturbations (Theorem 1.2). Finally, in Section 3.4 we prove Theorem 1.3 showing that the damping coefficient and the nonlinearity can be determined simultaneously.

3.1. Runge approximation

With the material from Section 2 at our disposal, we can now show the following Runge approximation theorem, whose proof is very similar to the one of [LTZ24a, Theorem 1.2].

Theorem 3.1 (Runge approximation).

Let Ωn\Omega\subset{\mathbb{R}}^{n} be a bounded Lipschitz domain, T>0T>0, 0<s<α10<s<\alpha\leq 1 and suppose that 1p1\leq p\leq\infty satisfies (1.6). Assume that we have given coefficients (γ,q)C0,α(n)×Lp(Ω)(\gamma,q)\in C^{0,\alpha}({\mathbb{R}}^{n})\times L^{p}(\Omega). Then for any measurement set WΩeW\subset\Omega_{e} and initial conditions (u0,u1)H~s(Ω)×L~2(Ω)(u_{0},u_{1})\in\widetilde{H}^{s}(\Omega)\times\widetilde{L}^{2}(\Omega), the Runge set

(3.1) Wu0,u1:={uφφ;φCc(WT)}\mathcal{R}^{u_{0},u_{1}}_{W}\vcentcolon=\{u_{\varphi}-\varphi\,;\,\varphi\in C_{c}^{\infty}(W_{T})\}

is dense in L2(0,T;H~s(Ω))L^{2}(0,T;\widetilde{H}^{s}(\Omega)), where uφu_{\varphi} is the unique solution to

(3.2) {(Lγ+q)u=0 in ΩT,u=φ on (Ωe)T,u(0)=u0,tu(0)=u1 on Ω,\begin{cases}(L_{\gamma}+q)u=0&\text{ in }\Omega_{T},\\ u=\varphi&\text{ on }(\Omega_{e})_{T},\\ u(0)=u_{0},\,\partial_{t}u(0)=u_{1}&\text{ on }\Omega,\end{cases}

(see Proposition 2.2).

Proof.

First of all note that it is enough to consider the case (u0,u1)=0(u_{0},u_{1})=0. To see this assume that the density holds for W:=W0,0\mathcal{R}_{W}\vcentcolon=\mathcal{R}_{W}^{0,0} and let fL2(0,T;H~s(Ω))f\in L^{2}(0,T;\widetilde{H}^{s}(\Omega)). Let v0v_{0} be the unique solution to

(3.3) {(Lγ+q)v=0 in ΩT,v=0 on (Ωe)T,v(0)=u0,tv(0)=u1 on Ω\begin{cases}(L_{\gamma}+q)v=0&\text{ in }\Omega_{T},\\ v=0&\text{ on }(\Omega_{e})_{T},\\ v(0)=u_{0},\,\partial_{t}v(0)=u_{1}&\text{ on }\Omega\end{cases}

and define f~:=fv0L2(0,T;H~s(Ω))\widetilde{f}\vcentcolon=f-v_{0}\in L^{2}(0,T;\widetilde{H}^{s}(\Omega)). By assumption there exists (φk)kCc(WT)(\varphi_{k})_{k\in{\mathbb{N}}}\subset C_{c}^{\infty}(W_{T}) such that ukφkf~u_{k}-\varphi_{k}\to\widetilde{f} in L2(0,T;H~s(Ω))L^{2}(0,T;\widetilde{H}^{s}(\Omega)) as kk\to\infty, where uku_{k} is the unique solution to

(3.4) {(Lγ+q)u=0 in ΩT,u=φ on (Ωe)T,u(0)=0,tu(0)=0 on Ω\begin{cases}(L_{\gamma}+q)u=0&\text{ in }\Omega_{T},\\ u=\varphi&\text{ on }(\Omega_{e})_{T},\\ u(0)=0,\,\partial_{t}u(0)=0&\text{ on }\Omega\end{cases}

with φ=φk\varphi=\varphi_{k}. Then vk:=uk+v0v_{k}\vcentcolon=u_{k}+v_{0} is the unique solution to (3.2) with φ=φk\varphi=\varphi_{k}. The above convergence now implies vkφkfv_{k}-\varphi_{k}\to f in L2(0,T;H~s(Ω))L^{2}(0,T;\widetilde{H}^{s}(\Omega)) as kk\to\infty and we get that Wu0,u1\mathcal{R}_{W}^{u_{0},u_{1}} is dense in L2(0,T;H~s(Ω))L^{2}(0,T;\widetilde{H}^{s}(\Omega)).

Therefore, it remains to show that W\mathcal{R}_{W} is dense in L2(0,T;H~s(Ω))L^{2}(0,T;\widetilde{H}^{s}(\Omega)). As usual, we show this by a Hahn–Banach argument. Thus, suppose that FL2(0,T;Hs(Ω))F\in L^{2}(0,T;H^{-s}(\Omega)) vanishes on W\mathcal{R}_{W}. Let us recall that if φCc(WT)\varphi\in C_{c}^{\infty}(W_{T}) and uu solves (3.4), then by (2.13) and Lemma 2.3 the function v=(uφ)v=(u-\varphi)^{\star} satisfies

(3.5) {(Lγ+q)v=(Δ)sφ in ΩT,v=0 in (Ωe)T,v(T)=tv(T)=0 in Ω.\begin{cases}(L_{-\gamma}+q)v=-(-\Delta)^{s}\varphi^{\star}&\text{ in }\Omega_{T},\\ v=0&\text{ in }(\Omega_{e})_{T},\\ v(T)=\partial_{t}v(T)=0&\text{ in }\Omega.\end{cases}

Next, let ww be the unique solution to

(3.6) {(Lγ+q)w=F in ΩT,w=0 in (Ωe)T,w(0)=tw(0)=0 in Ω\begin{cases}(L_{\gamma}+q)w=F^{\star}&\text{ in }\Omega_{T},\\ w=0&\text{ in }(\Omega_{e})_{T},\\ w(0)=\partial_{t}w(0)=0&\text{ in }\Omega\end{cases}

(see Theorem 2.8). By testing the equation for ww by vv, we get

0T(Δ)sφ,wL2(Ω)𝑑t=0TF,v𝑑t=0TF,uφφ𝑑t=0-\int_{0}^{T}\langle(-\Delta)^{s}\varphi^{\star},w\rangle_{L^{2}(\Omega)}\,dt=\int_{0}^{T}\langle F^{\star},v\rangle\,dt=\int_{0}^{T}\langle F,u_{\varphi}-\varphi\rangle\,dt=0

for any φCc(WT)\varphi\in C_{c}^{\infty}(W_{T}). This ensures that there holds

(Δ)sw=0 in WT.(-\Delta)^{s}w=0\quad\text{ in }W_{T}.

Furthermore, by construction ww vanishes in (Ωe)T(\Omega_{e})_{T} and hence the unique continuation principle for the fractional Laplacian guarantees w=0w=0 in Tn{\mathbb{R}}^{n}_{T} (see [GSU20]). As very weak solutions are distributional solutions, we get

0TF,Φ𝑑t=0T(Lγ+q)Φ,wL2(Ω)𝑑t=0\int_{0}^{T}\langle F^{\star},\Phi\rangle\,dt=\int_{0}^{T}\langle(L_{-\gamma}+q)\Phi,w\rangle_{L^{2}(\Omega)}\,dt=0

for all ΦCc(ΩT)\Phi\in C_{c}^{\infty}(\Omega_{T}). To see that very weak solutions are distributional solutions, one can simply take G=χΩ(Lγ+q)ΦG=\chi_{\Omega}(L_{\gamma}+q)\Phi with ΦCc(Ω×[0,T))\Phi\in C_{c}^{\infty}(\Omega\times[0,T)) in Definition 2.4, where χΩ\chi_{\Omega} denotes the characteristic function of Ω\Omega (see also [LTZ24a, Proposition 3.8]). By density of Cc(ΩT)C_{c}^{\infty}(\Omega_{T}) in L2(0,T;H~s(Ω))L^{2}(0,T;\widetilde{H}^{s}(\Omega)) we deduce that F=0F=0. This concludes the proof. ∎

As a consequence we have the following lemma.

Lemma 3.2 (Convergence of time derivative).

Let Ωn\Omega\subset{\mathbb{R}}^{n} be a bounded Lipschitz domain, T>0T>0, 0<s<α10<s<\alpha\leq 1 and suppose that 1p1\leq p\leq\infty satisfies (1.6). Assume that we have given coefficients (γ,q)C0,α(n)×Lp(Ω)(\gamma,q)\in C^{0,\alpha}({\mathbb{R}}^{n})\times L^{p}(\Omega). Let Φ,ΨL2(0,T;H~s(Ω))H1(0,T;Hs(Ω))\Phi,\Psi\in L^{2}(0,T;\widetilde{H}^{s}(\Omega))\cap H^{1}(0,T;H^{-s}(\Omega)) and suppose (φk)kCc((Ωe)T)(\varphi_{k})_{k\in{\mathbb{N}}}\subset C_{c}^{\infty}((\Omega_{e})_{T}) is such that

(3.7) ukφkΦ in L2(0,T;H~s(Ω)) as k,u_{k}-\varphi_{k}\to\Phi\text{ in }L^{2}(0,T;\widetilde{H}^{s}(\Omega))\text{ as }k\to\infty,

where uku_{k} solves

(3.8) {(Lγ+q)u=0 in ΩT,u=φk on (Ωe)T,u(0)=0,tu(0)=0 on Ω\begin{cases}(L_{\gamma}+q)u=0&\text{ in }\Omega_{T},\\ u=\varphi_{k}&\text{ on }(\Omega_{e})_{T},\\ u(0)=0,\,\partial_{t}u(0)=0&\text{ on }\Omega\end{cases}

for kk\in{\mathbb{N}}. If Φ,Ψ\Phi,\Psi satisfy one of the conditions

  1. (a)

    Ψ(T)=Φ(0)=0\Psi(T)=\Phi(0)=0

  2. (b)

    or Ψ(T)=Ψ(0)=0\Psi(T)=\Psi(0)=0,

then we have

(3.9) limk0Tt(ukφk),Ψ𝑑t=0TtΦ,Ψ𝑑t.\lim_{k\to\infty}\int_{0}^{T}\langle\partial_{t}(u_{k}-\varphi_{k}),\Psi\rangle\,dt=\int_{0}^{T}\langle\partial_{t}\Phi,\Psi\rangle\,dt.
Remark 3.3.

Let us note that the same formula (3.9) holds for second order time derivatives under appropriate conditions.

Proof.

Using the integration by parts formula, we may compute

limk0Tt(ukφk),Ψ𝑑t=limk((ukφk)(T),Ψ(T)L2(Ω)(ukφk)(0),Ψ(0)L2(Ω)0TtΨ,ukφk𝑑t)=limk0TtΨ,ukφk𝑑t=0TtΨ,Φ𝑑t=Φ(0),Ψ(0)L2(Ω)Φ(T),Ψ(T)L2(Ω)+0TtΦ,Ψ𝑑t=0TtΦ,Ψ𝑑t.\small\begin{split}&\lim_{k\to\infty}\int_{0}^{T}\langle\partial_{t}(u_{k}-\varphi_{k}),\Psi\rangle\,dt\\ &=\lim_{k\to\infty}\left(\langle(u_{k}-\varphi_{k})(T),\Psi(T)\rangle_{L^{2}(\Omega)}-\langle(u_{k}-\varphi_{k})(0),\Psi(0)\rangle_{L^{2}(\Omega)}-\int_{0}^{T}\langle\partial_{t}\Psi,u_{k}-\varphi_{k}\rangle\,dt\right)\\ &=-\lim_{k\to\infty}\int_{0}^{T}\langle\partial_{t}\Psi,u_{k}-\varphi_{k}\rangle\,dt\\ &=-\int_{0}^{T}\langle\partial_{t}\Psi,\Phi\rangle\,dt\\ &=\langle\Phi(0),\Psi(0)\rangle_{L^{2}(\Omega)}-\langle\Phi(T),\Psi(T)\rangle_{L^{2}(\Omega)}+\int_{0}^{T}\langle\partial_{t}\Phi,\Psi\rangle\,dt\\ &=\int_{0}^{T}\langle\partial_{t}\Phi,\Psi\rangle\,dt.\end{split}

In the first equality sign we used an integration by parts, in the second equality we used (3.8), Ψ(T)=0\Psi(T)=0 and (3.8), in the third equality the convergence (3.7), in the fourth equality again an integration by parts and finally in the last equality the conditions a or b. ∎

3.2. DN map and integral identities

Next, we define the Dirichlet to Neumann (DN) map Λγ,q\Lambda_{\gamma,q} related to

(3.10) {(Lγ+q)u=0 in ΩT,u=φ on (Ωe)T,u(0)=0,tu(0)=0 on Ω,\begin{cases}(L_{\gamma}+q)u=0&\text{ in }\Omega_{T},\\ u=\varphi&\text{ on }(\Omega_{e})_{T},\\ u(0)=0,\,\partial_{t}u(0)=0&\text{ on }\Omega,\end{cases}

via

(3.11) Λγ,qφ,ψ=Tn(Δ)s/2uφ(Δ)s/2ψ𝑑x\langle\Lambda_{\gamma,q}\varphi,\psi\rangle=\int_{{\mathbb{R}}^{n}_{T}}(-\Delta)^{s/2}u_{\varphi}(-\Delta)^{s/2}\psi\,dx

for all φ,ψCc((Ωe)T)\varphi,\psi\in C_{c}^{\infty}((\Omega_{e})_{T}), where uφu_{\varphi} is the unique solution to (3.10) with exterior condition φ\varphi. Using the above preparation, we now establish the following integral identity.

Proposition 3.4 (Integral identity for linear perturbations).

Let Ωn\Omega\subset{\mathbb{R}}^{n} be a bounded Lipschitz domain, T>0T>0, 0<s<α10<s<\alpha\leq 1 and suppose that 1p1\leq p\leq\infty satisfies (1.6). Assume that we have given coefficients (γj,qj)C0,α(n)×Lp(Ω)(\gamma_{j},q_{j})\in C^{0,\alpha}({\mathbb{R}}^{n})\times L^{p}(\Omega) for j=1,2j=1,2. Let φjCc((Ωe)T)\varphi_{j}\in C_{c}^{\infty}((\Omega_{e})_{T}) and denote by uju_{j} the corresponding solution of (3.10) with (γ,q)=(γj,qj)(\gamma,q)=(\gamma_{j},q_{j}). Then there holds

(3.12) (Λγ1,q1Λγ2,q2)φ1,φ2=ΩT{[(γ1γ2)t+q1q2](u1φ1)}(u2φ2)𝑑x𝑑t.\begin{split}&\langle(\Lambda_{\gamma_{1},q_{1}}-\Lambda_{\gamma_{2},q_{2}})\varphi_{1},\varphi_{2}^{\star}\rangle\\ &\quad=\int_{\Omega_{T}}\{[(\gamma_{1}-\gamma_{2})\partial_{t}+q_{1}-q_{2}](u_{1}-\varphi_{1})\}(u_{2}-\varphi_{2})^{\star}\,dxdt.\end{split}
Proof.

Let (Γj,Qj)C0,α(n)×Lp(Ω)(\Gamma_{j},Q_{j})\in C^{0,\alpha}({\mathbb{R}}^{n})\times L^{p}(\Omega), j=1,2j=1,2, and suppose UjU_{j} is the unique solutions of (3.10) with (γ,q)=(Γj,Qj)(\gamma,q)=(\Gamma_{j},Q_{j}) and exterior condition φ=ψj\varphi=\psi_{j}. Then we may compute

(3.13) ΩT{[(Γ1Γ2)t+Q1Q2](U1ψ1)}(U2ψ2)𝑑x𝑑t=ΩT{[Γ1t+Q1](U1ψ1)}(U2ψ2)𝑑x𝑑tΩT(U1ψ1)[Γ2t+Q2](U2ψ2)𝑑x𝑑t=0TLΓ1,Q1(U1ψ1),(U2ψ2)𝑑t0T(t2+(Δ)s)(U1ψ1),(U2ψ2)𝑑t0TLΓ2,Q2(U2ψ2),(U1ψ1)𝑑t+0T(t2+(Δ)s)(U2ψ2),(U1ψ1)𝑑t=0T(Δ)sψ1,(U2ψ2)L2(Ω)𝑑t+0T(Δ)sψ2,(U1ψ1)L2(Ω)𝑑t=Tn((Δ)sψ2)U1𝑑x𝑑tTn((Δ)sψ1)U2𝑑x𝑑t=ΛΓ1,Q1ψ1,ψ2ΛΓ2,Q2ψ2,ψ1.\begin{split}&\int_{\Omega_{T}}\{[(\Gamma_{1}-\Gamma_{2})\partial_{t}+Q_{1}-Q_{2}](U_{1}-\psi_{1})\}(U_{2}-\psi_{2})^{\star}\,dxdt\\ &=\int_{\Omega_{T}}\{[\Gamma_{1}\partial_{t}+Q_{1}](U_{1}-\psi_{1})\}(U_{2}-\psi_{2})^{\star}\,dxdt\\ &\quad-\int_{\Omega_{T}}(U_{1}-\psi_{1})[-\Gamma_{2}\partial_{t}+Q_{2}](U_{2}-\psi_{2})^{\star}\,dxdt\\ &=\int_{0}^{T}\langle L_{\Gamma_{1},Q_{1}}(U_{1}-\psi_{1}),(U_{2}-\psi_{2})^{\star}\rangle\,dt\\ &\quad-\int_{0}^{T}\langle(\partial_{t}^{2}+(-\Delta)^{s})(U_{1}-\psi_{1}),(U_{2}-\psi_{2})^{\star}\rangle\,dt\\ &\quad-\int_{0}^{T}\langle L_{-\Gamma_{2},Q_{2}}(U_{2}-\psi_{2})^{\star},(U_{1}-\psi_{1})\rangle\,dt\\ &\quad+\int_{0}^{T}\langle(\partial_{t}^{2}+(-\Delta)^{s})(U_{2}-\psi_{2})^{\star},(U_{1}-\psi_{1})\rangle\,dt\\ &=-\int_{0}^{T}\langle(-\Delta)^{s}\psi_{1},(U_{2}-\psi_{2})^{\star}\rangle_{L^{2}(\Omega)}\,dt+\int_{0}^{T}\langle(-\Delta)^{s}\psi_{2}^{\star},(U_{1}-\psi_{1})\rangle_{L^{2}(\Omega)}\,dt\\ &=\int_{{\mathbb{R}}^{n}_{T}}((-\Delta)^{s}\psi_{2}^{\star})U_{1}\,dxdt-\int_{{\mathbb{R}}^{n}_{T}}((-\Delta)^{s}\psi_{1})U_{2}^{\star}\,dxdt\\ &=\langle\Lambda_{\Gamma_{1},Q_{1}}\psi_{1},\psi_{2}^{\star}\rangle-\langle\Lambda_{\Gamma_{2},Q_{2}}\psi_{2},\psi_{1}^{\star}\rangle.\end{split}

In the first equality we used that U1ψ1U_{1}-\psi_{1} has vanishing initial conditions, (U2ψ2)(U_{2}-\psi_{2})^{\star} has vanishing terminal conditions and an integration by parts. In the third equality we used that the PDEs for U1ψ1U_{1}-\psi_{1} and (U2ψ2)(U_{2}-\psi_{2})^{\star} hold in the sense of L2(0,T;Hs(Ω))=(L2(0,T;H~s(Ω))L^{2}(0,T;H^{-s}(\Omega))=(L^{2}(0,T;\widetilde{H}^{s}(\Omega))^{\prime} (see Lemma 2.3). In the fourth equality, we used the PDEs for U1U_{1} and U2U_{2}, Lemma 2.3 and that there holds

0T(t2+(Δ)s)(U2ψ2),(U1ψ1)𝑑t=0T(t2+(Δ)s)(U1ψ1),(U2ψ2)𝑑t,\begin{split}&\int_{0}^{T}\langle(\partial_{t}^{2}+(-\Delta)^{s})(U_{2}-\psi_{2})^{\star},(U_{1}-\psi_{1})\rangle\,dt\\ &=\int_{0}^{T}\langle(\partial_{t}^{2}+(-\Delta)^{s})(U_{1}-\psi_{1}),(U_{2}-\psi_{2})^{\star}\rangle\,dt,\end{split}

which can be established similarly as [LTZ24c, Claim 4.2] (see also the proof of Theorem 2.8). In the last equality, we have made the change of variables τ=Tt\tau=T-t for the second integral. On the one hand, using (3.13) with

Γ1=Γ2=γj and Q1=Q2=qj,\Gamma_{1}=\Gamma_{2}=\gamma_{j}\text{ and }Q_{1}=Q_{2}=q_{j},

we observe that

(3.14) Λγj,qjψ1,ψ2=Λγj,qjψ2,ψ1\langle\Lambda_{\gamma_{j},q_{j}}\psi_{1},\psi_{2}^{\star}\rangle=\langle\Lambda_{\gamma_{j},q_{j}}\psi_{2},\psi_{1}^{\star}\rangle

for all ψjCc((Ωe)T)\psi_{j}\in C_{c}^{\infty}((\Omega_{e})_{T}), j=1,2j=1,2. On the other hand, choosing

Γj=γj,Qj=qj and ψj=φj\Gamma_{j}=\gamma_{j},\,Q_{j}=q_{j}\text{ and }\psi_{j}=\varphi_{j}

in (3.13) and taking into account the self-adjointness (3.14), we get (3.12). ∎

3.3. Simultaneous determination of damping coefficient and linear perturbations

Proof of Theorem 1.2.

First note that by the integral identity in Proposition 3.4, we may deduce from the condition (1.12) that there holds

(3.15) ΩT{[(γ1γ2)t+q1q2](u1φ1)}(u2φ2)𝑑x𝑑t=0\int_{\Omega_{T}}\{[(\gamma_{1}-\gamma_{2})\partial_{t}+q_{1}-q_{2}](u_{1}-\varphi_{1})\}(u_{2}-\varphi_{2})^{\star}\,dxdt=0

for all φjCc((Wj)T)\varphi_{j}\in C_{c}^{\infty}((W_{j})_{T}), where uju_{j} denotes the unique solution to

(3.16) {(Lγj+qj)u=0 in ΩT,u=φj on (Ωe)T,u(0)=0,tu(0)=0 on Ω.\begin{cases}(L_{\gamma_{j}}+q_{j})u=0&\text{ in }\Omega_{T},\\ u=\varphi_{j}&\text{ on }(\Omega_{e})_{T},\\ u(0)=0,\,\partial_{t}u(0)=0&\text{ on }\Omega.\end{cases}

Let ωΩ\omega\Subset\Omega and choose a cutoff function Φ1Cc(Ω)\Phi_{1}\in C_{c}^{\infty}(\Omega) satisfying Φ1=1\Phi_{1}=1 on ω\omega. Moreover, let Φ2Cc(ωT)\Phi_{2}\in C_{c}^{\infty}(\omega_{T}). By the Runge approximation (Theorem 3.1), there exist sequences (φjk)kCc((Wj)T)(\varphi_{j}^{k})_{k\in{\mathbb{N}}}\subset C_{c}^{\infty}((W_{j})_{T}) with corresponding solutions ujku_{j}^{k} of (3.16) with φj=φjk\varphi_{j}=\varphi_{j}^{k} such that ujkφjkΦju_{j}^{k}-\varphi_{j}^{k}\to\Phi_{j} in L2(0,T;H~s(Ω))L^{2}(0,T;\widetilde{H}^{s}(\Omega)). Taking φ1=φ1k\varphi_{1}=\varphi_{1}^{k} and φ2=φ2\varphi_{2}=\varphi_{2}^{\ell} in (3.15) gives

ΩT{[(γ1γ2)t+q1q2](u1kφ1k)}(u2φ2)𝑑x𝑑t=0\int_{\Omega_{T}}\{[(\gamma_{1}-\gamma_{2})\partial_{t}+q_{1}-q_{2}](u^{k}_{1}-\varphi^{k}_{1})\}(u^{\ell}_{2}-\varphi^{\ell}_{2})^{\star}\,dxdt=0

for all k,k,\ell\in{\mathbb{N}}. First, we let \ell\to\infty to deduce

(3.17) ΩT{[(γ1γ2)t+q1q2](u1kφ1k)}Φ2𝑑x𝑑t=0\int_{\Omega_{T}}\{[(\gamma_{1}-\gamma_{2})\partial_{t}+q_{1}-q_{2}](u^{k}_{1}-\varphi^{k}_{1})\}\Phi_{2}^{\star}\,dxdt=0

for all kk\in{\mathbb{N}}. As γ1γ2C0,α(n)\gamma_{1}-\gamma_{2}\in C^{0,\alpha}({\mathbb{R}}^{n}) the estimate (2.38) ensures that we can apply Lemma 3.2 under the condition b and so tΦ1=0\partial_{t}\Phi_{1}=0 shows that the first term in (3.17) goes to zero. So in the limit kk\to\infty what remains is

ΩT(q1q2)Φ2𝑑x𝑑t=0,\int_{\Omega_{T}}(q_{1}-q_{2})\Phi_{2}^{\star}\,dxdt=0,

where we used Φ1=1\Phi_{1}=1 on ω\omega. This ensures that q1=q2q_{1}=q_{2} on ω\omega. As the set ω\omega is arbitrary, we get q1=q2q_{1}=q_{2} in Ω\Omega. Now, the identity (3.15) reduces to

ΩT{[(γ1γ2)t](u1φ1)}(u2φ2)𝑑x𝑑t=0\int_{\Omega_{T}}\{[(\gamma_{1}-\gamma_{2})\partial_{t}](u_{1}-\varphi_{1})\}(u_{2}-\varphi_{2})^{\star}\,dxdt=0

for all φjCc((Wj)T)\varphi_{j}\in C_{c}^{\infty}((W_{j})_{T}). We choose ηCc(ΩT)\eta\in C_{c}^{\infty}(\Omega_{T}), define

Φ1(x,t)=0tη(x,τ)𝑑τCc(Ω×(0,T])\Phi_{1}(x,t)=\int_{0}^{t}\eta(x,\tau)\,d\tau\in C_{c}^{\infty}(\Omega\times(0,T])

and take Φ2Cc(ΩT)\Phi_{2}\in C_{c}^{\infty}(\Omega_{T}). Then using tΦ1=η\partial_{t}\Phi_{1}=\eta and arguing as above via a Runge approximation and Lemma 3.2, we get from (3.15) the identity

ΩT(γ1γ2)ηΦ2𝑑x𝑑t=0.\int_{\Omega_{T}}(\gamma_{1}-\gamma_{2})\eta\Phi_{2}^{\star}\,dxdt=0.

This again implies γ1=γ2\gamma_{1}=\gamma_{2} in Ω\Omega. ∎

3.4. Simultaneous determination of damping coefficient and nonlinearity

Before turning to the proof of our second main result, let us recall that the DN map related to the problem

(3.18) {Lγu+f(u)=0 in ΩT,u=φ on (Ωe)T,u(0)=0,tu(0)=0 on Ω\begin{cases}L_{\gamma}u+f(u)=0&\text{ in }\Omega_{T},\\ u=\varphi&\text{ on }(\Omega_{e})_{T},\\ u(0)=0,\,\partial_{t}u(0)=0&\text{ on }\Omega\end{cases}

is defined by

(3.19) Λγ,fφ,ψ:=Tn(Δ)s/2uφ(Δ)s/2ψ𝑑x𝑑t,\langle\Lambda_{\gamma,f}\varphi,\psi\rangle\vcentcolon=\int_{{\mathbb{R}}^{n}_{T}}(-\Delta)^{s/2}u_{\varphi}(-\Delta)^{s/2}\psi\,dxdt,

where φ,ψCc((Ωe)T)\varphi,\psi\in C_{c}^{\infty}((\Omega_{e})_{T}) and uφu_{\varphi} is the unique solution to (3.18) (see [LTZ24c, Proposition 3.7]).

Proof of Theorem 1.3.

Let ε>0\varepsilon>0 and denote by uε(j)u^{(j)}_{\varepsilon} the unique solutions to (3.18) with f=fjf=f_{j}, γ=γj\gamma=\gamma_{j} and φ=εη\varphi=\varepsilon\eta for some fixed ηCc((W1)T)\eta\in C_{c}^{\infty}((W_{1})_{T}). Let us observe that the UCP for the fractional Laplacian and the condition (1.15) imply that uε:=uε(1)=uε(2)u_{\varepsilon}\vcentcolon=u^{(1)}_{\varepsilon}=u^{(2)}_{\varepsilon}. Next, let us note that we can write

(3.20) uε=εvj+Rε(j)u_{\varepsilon}=\varepsilon v_{j}+R^{(j)}_{\varepsilon}

for j=1,2j=1,2, where vjv_{j} and Rε(j)R^{(j)}_{\varepsilon} are the unique solutions of

(3.21) {Lγjv=0 in ΩT,v=η on (Ωe)T,v(0)=0,tv(0)=0 on Ω\begin{cases}L_{\gamma_{j}}v=0&\text{ in }\Omega_{T},\\ v=\eta&\text{ on }(\Omega_{e})_{T},\\ v(0)=0,\,\partial_{t}v(0)=0&\text{ on }\Omega\end{cases}

and

(3.22) {LγjR=fj(uε) in ΩT,R=0 on (Ωe)T,R(0)=0,tR(0)=0 on Ω,\begin{cases}L_{\gamma_{j}}R=-f_{j}(u_{\varepsilon})&\text{ in }\Omega_{T},\\ R=0&\text{ on }(\Omega_{e})_{T},\\ R(0)=0,\,\partial_{t}R(0)=0&\text{ on }\Omega,\end{cases}

respectively. This simply follows from the unique solvability of (3.18) and both functions uεu_{\varepsilon} and εvj+Rε(j)\varepsilon v_{j}+R^{(j)}_{\varepsilon} are solutions. Furthermore, we notice that the energy estimate of [LTZ24c, Theorem 3.1], [LTZ24c, eq. (3.18)] and the r+1r+1 homogeneity of fjf_{j} ensure that Rε(j)R^{(j)}_{\varepsilon} satisfies

(3.23) tRε(j)L(0,T;L2(Ω))+Rε(j)(t)L(0,T;Hs(n))fj(uε)L2(ΩT)uεL(0,T;Hs(n))r+1.\begin{split}\|\partial_{t}R^{(j)}_{\varepsilon}\|_{L^{\infty}(0,T;L^{2}(\Omega))}+\|R^{(j)}_{\varepsilon}(t)\|_{L^{\infty}(0,T;H^{s}({\mathbb{R}}^{n}))}&\lesssim\|f_{j}(u_{\varepsilon})\|_{L^{2}(\Omega_{T})}\\ &\lesssim\|u_{\varepsilon}\|^{r+1}_{L^{\infty}(0,T;H^{s}({\mathbb{R}}^{n}))}.\end{split}

Moreover, we may estimate

(3.24) tuεL(0,T;L2(n))+uεL(0,T;Hs(n))t(uεεη)L(0,T;L2(Ω))+uεεηL(0,T;Hs(n))+εηW1,(0,T;H2s(n))εηW1,(0,T;H2s(n)).\begin{split}&\|\partial_{t}u_{\varepsilon}\|_{L^{\infty}(0,T;L^{2}({\mathbb{R}}^{n}))}+\|u_{\varepsilon}\|_{L^{\infty}(0,T;H^{s}({\mathbb{R}}^{n}))}\\ &\lesssim\|\partial_{t}(u_{\varepsilon}-\varepsilon\eta)\|_{L^{\infty}(0,T;L^{2}(\Omega))}+\|u_{\varepsilon}-\varepsilon\eta\|_{L^{\infty}(0,T;H^{s}({\mathbb{R}}^{n}))}+\varepsilon\|\eta\|_{W^{1,\infty}(0,T;H^{2s}({\mathbb{R}}^{n}))}\\ &\lesssim\varepsilon\|\eta\|_{W^{1,\infty}(0,T;H^{2s}({\mathbb{R}}^{n}))}.\end{split}

This follows from the following observations. If uu solves (3.18) for a damping coefficient γC0,α(n)\gamma\in C^{0,\alpha}({\mathbb{R}}^{n}), a weak nonlinearity ff and φCc((Ωe)T)\varphi\in C_{c}^{\infty}((\Omega_{e})_{T}), then v=uφv=u-\varphi solves

(3.25) {Lγv+f(v)=(Δ)sφ in ΩT,v=0 on (Ωe)T,v(0)=0,tv(0)=0 on Ω.\begin{cases}L_{\gamma}v+f(v)=-(-\Delta)^{s}\varphi&\text{ in }\Omega_{T},\\ v=0&\text{ on }(\Omega_{e})_{T},\\ v(0)=0,\,\partial_{t}v(0)=0&\text{ on }\Omega.\end{cases}

Now, we may invoke [LTZ24c, eq. (3.15)] to find that there holds

tv(t)L2(Ω)2+v(t)Hs(n)20t|γtv,tvL2(Ω)|𝑑τ+0t|(Δ)sφ,tvL2(Ω)|𝑑τ(Δ)sφL2(0,t;L2(Ω))2+0ttvL2(Ω)2𝑑τ.\begin{split}&\|\partial_{t}v(t)\|_{L^{2}(\Omega)}^{2}+\|v(t)\|_{H^{s}({\mathbb{R}}^{n})}^{2}\\ &\lesssim\int_{0}^{t}|\langle\gamma\partial_{t}v,\partial_{t}v\rangle_{L^{2}(\Omega)}|\,d\tau+\int_{0}^{t}|\langle(-\Delta)^{s}\varphi,\partial_{t}v\rangle_{L^{2}(\Omega)}|\,d\tau\\ &\lesssim\|(-\Delta)^{s}\varphi\|_{L^{2}(0,t;L^{2}(\Omega))}^{2}+\int_{0}^{t}\|\partial_{t}v\|_{L^{2}(\Omega)}^{2}\,d\tau.\end{split}

Thus, Gronwall’s inequality gives

tv(t)L2(Ω)+v(t)Hs(n)(Δ)sφL2(0,t;L2(Ω)).\|\partial_{t}v(t)\|_{L^{2}(\Omega)}+\|v(t)\|_{H^{s}({\mathbb{R}}^{n})}\lesssim\|(-\Delta)^{s}\varphi\|_{L^{2}(0,t;L^{2}(\Omega))}.

This ensures the validity of the second estimate in (3.24). Next, observe that by subtracting the PDEs for uε(1)u^{(1)}_{\varepsilon} and uε(2)u^{(2)}_{\varepsilon}, we deduce that

(3.26) (γ1γ2)tuε=f2(uε)f1(uε) in ΩT.(\gamma_{1}-\gamma_{2})\partial_{t}u_{\varepsilon}=f_{2}(u_{\varepsilon})-f_{1}(u_{\varepsilon})\text{ in }\Omega_{T}.

By (3.20), we may write

(3.27) (γ1γ2)(εtv1+tRε(1))=f2(uε)f1(uε) in ΩT.(\gamma_{1}-\gamma_{2})(\varepsilon\partial_{t}v_{1}+\partial_{t}R^{(1)}_{\varepsilon})=f_{2}(u_{\varepsilon})-f_{1}(u_{\varepsilon})\text{ in }\Omega_{T}.

Combining (3.23) and (3.24), we see that

(3.28) tRε(j)L(0,T;L2(Ω))+Rε(j)(t)L(0,T;Hs(n))εr+1.\|\partial_{t}R^{(j)}_{\varepsilon}\|_{L^{\infty}(0,T;L^{2}(\Omega))}+\|R^{(j)}_{\varepsilon}(t)\|_{L^{\infty}(0,T;H^{s}({\mathbb{R}}^{n}))}\lesssim\varepsilon^{r+1}.

Multiplying by ε1\varepsilon^{-1} gives

(3.29) (γ1γ2)(tv1+ε1tRε(1))=f2(ε1/(r+1)uε)f1(ε1/(r+1)uε) in ΩT.\begin{split}&(\gamma_{1}-\gamma_{2})(\partial_{t}v_{1}+\varepsilon^{-1}\partial_{t}R^{(1)}_{\varepsilon})=f_{2}(\varepsilon^{-1/(r+1)}u_{\varepsilon})-f_{1}(\varepsilon^{-1/(r+1)}u_{\varepsilon})\text{ in }\Omega_{T}.\end{split}

Next, let us focus one the case 2s<n2s<n as the other one can be treated similarly. As r>0r>0 we deduce from (3.24) that ε1/(1+r)uε0\varepsilon^{-1/(1+r)}u_{\varepsilon}\to 0 in L(0,T;Hs(n))L^{\infty}(0,T;H^{s}({\mathbb{R}}^{n})) and so by Sobolev’s embedding in Lq(0,T;L2s(Ω))L^{q}(0,T;L^{2_{s}^{*}}(\Omega)) for all 1q1\leq q\leq\infty and 2s=2nn2s2_{s}^{*}=\frac{2n}{n-2s}. Hence, by our assumptions on fjf_{j} and [Zim24, Lemma 3.6], we get

(3.30) fj(ε1/(r+1)uε)0 in Lq/(r+1)(0,T;L2s/(r+1)(Ω))f_{j}(\varepsilon^{-1/(r+1)}u_{\varepsilon})\to 0\text{ in }L^{q/(r+1)}(0,T;L^{2_{s}^{*}/(r+1)}(\Omega))

for all qr+1q\geq r+1 as ε0\varepsilon\to 0. Additionally, using (3.28) we know that

(3.31) ε1tRε(j)0 in L(0,T;L2(Ω)).\varepsilon^{-1}\partial_{t}R^{(j)}_{\varepsilon}\to 0\text{ in }L^{\infty}(0,T;L^{2}(\Omega)).

Therefore, from (3.29), (3.30) and (3.31), we infer

(γ1γ2)tv1=0 in ΩT.(\gamma_{1}-\gamma_{2})\partial_{t}v_{1}=0\text{ in }\Omega_{T}.

In particular, this ensures that there holds

ΩT(γ1γ2)t(v1η)(w2ψ)dxdt=0\int_{\Omega_{T}}(\gamma_{1}-\gamma_{2})\partial_{t}(v_{1}-\eta)(w_{2}-\psi)^{\star}\,dxdt=0

for any ψCc((W2)T)\psi\in C_{c}^{\infty}((W_{2})_{T}), where w2w_{2} is the unique solution of

(3.32) {Lγ2w=0 in ΩT,w=ψ on (Ωe)T,w(0)=0,tw(0)=0 on Ω.\begin{cases}L_{\gamma_{2}}w=0&\text{ in }\Omega_{T},\\ w=\psi&\text{ on }(\Omega_{e})_{T},\\ w(0)=0,\,\partial_{t}w(0)=0&\text{ on }\Omega.\end{cases}

Now, arguing as in the previous section, we get γ1=γ2\gamma_{1}=\gamma_{2} in Ω\Omega. Hence, (3.26) reduces to

(3.33) f1(uε)=f2(uε) in ΩT.f_{1}(u_{\varepsilon})=f_{2}(u_{\varepsilon})\text{ in }\Omega_{T}.

Multiplying this identity by ε(r+1)\varepsilon^{-(r+1)} and arguing as before, we deduce that

f1(v)=f2(v) in ΩT,f_{1}(v)=f_{2}(v)\text{ in }\Omega_{T},

where v:=v1=v2v\vcentcolon=v_{1}=v_{2} as γ1=γ2\gamma_{1}=\gamma_{2}. One can now show f1(x,τ)=f2(x,τ)f_{1}(x,\tau)=f_{2}(x,\tau) for all xΩx\in\Omega and τ\tau\in{\mathbb{R}} exactly as described in [LTZ24a, p. 29]. Hence, we can conclude the proof. ∎

Acknowledgments. P. Zimmermann was supported by the Swiss National Science Foundation (SNSF), under grant number 214500.

Statements and Declarations

Data availability statement

No datasets were generated or analyzed during the current study.

Conflict of Interests

Hereby we declare there are no conflict of interests.

References

  • [CGRU23] Giovanni Covi, Tuhin Ghosh, Angkana Rüland, and Gunther Uhlmann. A reduction of the fractional Calderón problem to the local Calderón problem by means of the Caffarelli-Silvestre extension. arXiv preprint arXiv:2305.04227, 2023.
  • [CLL19] Xinlin Cao, Yi-Hsuan Lin, and Hongyu Liu. Simultaneously recovering potentials and embedded obstacles for anisotropic fractional Schrödinger operators. Inverse Probl. Imaging, 13(1):197–210, 2019.
  • [CLR20] Mihajlo Cekic, Yi-Hsuan Lin, and Angkana Rüland. The Calderón problem for the fractional Schrödinger equation with drift. Cal. Var. Partial Differential Equations, 59(91), 2020.
  • [CRTZ22] Giovanni Covi, Jesse Railo, Teemu Tyni, and Philipp Zimmermann. Stability estimates for the inverse fractional conductivity problem, 2022.
  • [CRTZ24] Giovanni Covi, Jesse Railo, Teemu Tyni, and Philipp Zimmermann. Stability estimates for the inverse fractional conductivity problem. SIAM Journal on Mathematical Analysis, 56(2):2456–2487, 2024.
  • [DL92] Robert Dautray and Jacques-Louis Lions. Mathematical analysis and numerical methods for science and technology. Vol. 5. Springer-Verlag, Berlin, 1992. Evolution problems. I, With the collaboration of Michel Artola, Michel Cessenat and Hélène Lanchon, Translated from the French by Alan Craig.
  • [Fei21] Ali Feizmohammadi. Fractional Calderón’ problem on a closed Riemannian manifold. arXiv preprint arXiv:2110.07500, 2021.
  • [FGKU21] Ali Feizmohammadi, Tuhin Ghosh, Katya Krupchyk, and Gunther Uhlmann. Fractional anisotropic Calderón problem on closed Riemannian manifolds. arXiv:2112.03480, 2021.
  • [FKU24] Ali Feizmohammadi, Katya Krupchyk, and Gunther Uhlmann. Calderón problem for fractional Schrödinger operators on closed Riemannian manifolds. arXiv preprint arXiv:2407.16866, 2024.
  • [FY24] Song-Ren Fu and Yongyi Yu. Well-posedness and inverse problems for the nonlocal third-order acoustic equation with time-dependent nonlinearity, 2024.
  • [GLX17] Tuhin Ghosh, Yi-Hsuan Lin, and Jingni Xiao. The Calderón problem for variable coefficients nonlocal elliptic operators. Comm. Partial Differential Equations, 42(12):1923–1961, 2017.
  • [GSU20] Tuhin Ghosh, Mikko Salo, and Gunther Uhlmann. The Calderón problem for the fractional Schrödinger equation. Anal. PDE, 13(2):455–475, 2020.
  • [KLW22] Pu-Zhao Kow, Yi-Hsuan Lin, and Jenn-Nan Wang. The Calderón problem for the fractional wave equation: uniqueness and optimal stability. SIAM J. Math. Anal., 54(3):3379–3419, 2022.
  • [KLZ24] Manas Kar, Yi-Hsuan Lin, and Philipp Zimmermann. Determining coefficients for a fractional pp-laplace equation from exterior measurements. J. Differential Equations, accepted for publication, 2024.
  • [LL22] Ru-Yu Lai and Yi-Hsuan Lin. Inverse problems for fractional semilinear elliptic equations. Nonlinear Anal., 216:Paper No. 112699, 21, 2022.
  • [LL23] Yi-Hsuan Lin and Hongyu Liu. Inverse problems for fractional equations with a minimal number of measurements. Communications and Computational Analysis, 1:72–93, 2023.
  • [LLU23] Ching-Lung Lin, Yi-Hsuan Lin, and Gunther Uhlmann. The Calderón problem for nonlocal parabolic operators: A new reduction from the nonlocal to the local. arXiv preprint arXiv:2308.09654, 2023.
  • [LNZ24] Yi-Hsuan Lin, Gen Nakamura, and Philipp Zimmermann. The calderón problem for the schrödinger equation in transversally anisotropic geometries with partial data, 2024.
  • [LRZ22] Yi-Hsuan Lin, Jesse Railo, and Philipp Zimmermann. The Calderón problem for a nonlocal diffusion equation with time-dependent coefficients. arXiv preprint arXiv:2211.07781, 2022.
  • [LTZ24a] Yi-Hsuan Lin, Teemu Tyni, and Philipp Zimmermann. Optimal runge approximation for nonlocal wave equations and unique determination of polyhomogeneous nonlinearities, 2024.
  • [LTZ24b] Yi-Hsuan Lin, Teemu Tyni, and Philipp Zimmermann. Well-posedness and inverse problems for semilinear nonlocal wave equations. Nonlinear Analysis, 247:113601, 2024.
  • [LTZ24c] Yi-Hsuan Lin, Teemu Tyni, and Philipp Zimmermann. Well-posedness and inverse problems for semilinear nonlocal wave equations. Nonlinear Analysis, 247:113601, 2024.
  • [LZ23] Yi-Hsuan Lin and Philipp Zimmermann. Unique determination of coefficients and kernel in nonlocal porous medium equations with absorption term. arXiv preprint arXiv:2305.16282, 2023.
  • [LZ24] Yi-Hsuan Lin and Philipp Zimmermann. Approximation and uniqueness results for the nonlocal diffuse optical tomography problem. arXiv preprint arXiv:2406.06226, 2024.
  • [RS20] Angkana Rüland and Mikko Salo. The fractional Calderón problem: low regularity and stability. Nonlinear Anal., 193:111529, 56, 2020.
  • [RZ23] Jesse Railo and Philipp Zimmermann. Fractional Calderón problems and Poincaré inequalities on unbounded domains. J. Spectr. Theory, 13(1):63–131, 2023.
  • [RZ24] Jesse Railo and Philipp Zimmermann. Low regularity theory for the inverse fractional conductivity problem. Nonlinear Analysis, 239:113418, 2024.
  • [SU87] John Sylvester and Gunther Uhlmann. A global uniqueness theorem for an inverse boundary value problem. Ann. of Math. (2), 125(1):153–169, 1987.
  • [Zim24] Philipp Zimmermann. Calderón problem for nonlocal viscous wave equations: Unique determination of linear and nonlinear perturbations, 2024.