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

\tocauthor

Yikan Liu, and Masahiro Yamamoto 11institutetext: Research Center of Mathematics for Social Creativity, Research Institute for Electronic Science, Hokkaido University, N12W7, Kita-Ward, Sapporo 060-0812, Japan,
11email: [email protected]
22institutetext: Graduate School of Mathematical Sciences, The University of Tokyo, 3-8-1 Komaba, Meguro-ku, Tokyo 153-8914, Japan,
22email: [email protected]
33institutetext: Honorary Member of Academy of Romanian Scientists, Ilfov, nr. 3, Bucuresti, Romania 44institutetext: Correspondence member of Accademia Peloritana dei Pericolanti, Palazzo Università, Piazza S. Pugliatti 1 98122 Messina, Italy

Uniqueness of inverse source problems for
time-fractional diffusion equations with
singular functions in time

Yikan Liu 11    Masahiro Yamamoto 223344
Abstract

We consider a fractional diffusion equations of order α(0,1)\alpha\in(0,1) whose source term is singular in time:

(tα+A)u(𝒙,t)=μ(t)f(𝒙),(𝒙,t)Ω×(0,T),(\partial_{t}^{\alpha}+A)u(\bm{x},t)=\mu(t)f(\bm{x}),\quad(\bm{x},t)\in\Omega\times(0,T),

where μ\mu belongs to a Sobolev space of negative order. In inverse source problems of determining f|Ωf|_{\Omega} by the data u|ω×(0,T)u|_{\omega\times(0,T)} with a given subdomain ωΩ\omega\subset\Omega or μ|(0,T)\mu|_{(0,T)} by the data u|{𝒙0}×(0,T)u|_{\{\bm{x}_{0}\}\times(0,T)} with a given point 𝒙0Ω\bm{x}_{0}\in\Omega, we prove the uniqueness by reducing to the case μL2(0,T)\mu\in L^{2}(0,T). The key is a transformation of a solution to an initial-boundary value problem with a regular function in time.

keywords:
time-fractional diffusion equation, inverse source problem, uniqueness

1 Introduction

Let Ωd\Omega\subset\mathbb{R}^{d} (d:={1,2,}d\in\mathbb{N}:=\{1,2,\ldots\}) be a bounded domain with a smooth boundary Ω\partial\Omega, and let 𝝂=𝝂(𝒙)\bm{\nu}=\bm{\nu}(\bm{x}) be the unit outward normal vector of Ω\partial\Omega. We set

Av(𝒙)=i,j=1dj(aij(𝒙)iv(𝒙))+j=1dbj(𝒙)jv(𝒙)+c(𝒙)v(𝒙),𝒙Ω,Av(\bm{x})=-\sum_{i,j=1}^{d}\partial_{j}(a_{ij}(\bm{x})\partial_{i}v(\bm{x}))+\sum_{j=1}^{d}b_{j}(\bm{x})\partial_{j}v(\bm{x})+c(\bm{x})v(\bm{x}),\quad\bm{x}\in\Omega, (1)

where aij=ajiC1(Ω¯)a_{ij}=a_{ji}\in C^{1}(\overline{\Omega}), bjC(Ω¯)b_{j}\in C(\overline{\Omega}) (1i,jd1\leq i,j\leq d), cC(Ω¯)c\in C(\overline{\Omega}) and we assume that there exists a constant κ>0\kappa>0 such that

i,j=1daij(𝒙)ξiξjκj=1d|ξj|2,𝒙Ω,(ξ1,,ξd)d.\sum_{i,j=1}^{d}a_{ij}(\bm{x})\xi_{i}\xi_{j}\geq\kappa\sum_{j=1}^{d}|\xi_{j}|^{2},\quad\forall\,\bm{x}\in\Omega,\ \forall\,(\xi_{1},\ldots,\xi_{d})\in\mathbb{R}^{d}.

We consider an initial-boundary value problem for a time-fractional diffusion equation whose source term is described by μ(t)f(𝒙)\mu(t)f(\bm{x}), where ff is a spatial distribution of the source and μ\mu is a temporal change factor. We can describe a governing initial-boundary value problem as follows:

{(dtα+A)u(𝒙,t)=μ(t)f(𝒙),(𝒙,t)Ω×(0,T),u(𝒙,0)=0,𝒙Ω,u(𝒙,t)=0,(𝒙,t)Ω×(0,T).\begin{cases}(\mathrm{d}_{t}^{\alpha}+A)u(\bm{x},t)=\mu(t)f(\bm{x}),&(\bm{x},t)\in\Omega\times(0,T),\\ u(\bm{x},0)=0,&\bm{x}\in\Omega,\\ u(\bm{x},t)=0,&(\bm{x},t)\in\partial\Omega\times(0,T).\end{cases} (2)

Here, for 0<α<10<\alpha<1, we can formally define the pointwise Caputo derivative as

dtαv(t):=1Γ(1α)0t(ts)αv(s)ds,vW1,1(0,T).\mathrm{d}_{t}^{\alpha}v(t):=\frac{1}{\Gamma(1-\alpha)}\int^{t}_{0}(t-s)^{-\alpha}v^{\prime}(s)\,\mathrm{d}s,\quad v\in W^{1,1}(0,T).

Owing to their capability of describing memory effects, time-fractional partial differential equations such as (2) have gathered consistent popularity among multidisciplinary researchers as models for anomalous diffusion and viscoelasticity (e.g. [2, 3, 9]). Though the history of fractional calculus can be traced back to Leibniz, the main focus on fractional equations was biased to the construction of explicit and approximate solutions via special functions and transforms until the last decades due to the needs from applied science. It has only been started recently that problems like (2) are formulated in appropriate function spaces using modern mathematical tools, followed by rapidly increasing literature on their fundamental theories, numerical analysis and inverse problems. Here we do not intend to give a complete bibliography, but only refer to several milestone works [4, 8, 14, 22] and the references therein.

Especially, the source term in (2) takes the form of separated variables, where μ(t)\mu(t) and f(𝒙)f(\bm{x}) describe the time evolution and the spatial distribution of some contaminant source, respectively. Therefore, the determination of μ(t)\mu(t) or f(𝒙)f(\bm{x}) turns out to be important in the context of environmental issues, which motivates us to propose the following problem.

Problem 1.1 (inverse source problems).

Let uu satisfy (2), ωΩ\emptyset\neq\omega\subset\Omega be an arbitrary subdomain and 𝐱0Ω\bm{x}_{0}\in\Omega be an arbitrary point. Determine μ|(0,T)\mu|_{(0,T)} by u|{𝐱0}×(0,T)u|_{\{\bm{x}_{0}\}\times(0,T)} with given ff and f|Ωf|_{\Omega} by u|ω×(0,T)u|_{\omega\times(0,T)} with given μ,\mu, respectively.

Indeed, Problem 1.1 includes two inverse problems, namely, the determination of μ(t)\mu(t) by the single point observation and that of f(𝒙)f(\bm{x}) by the partial interior observation of uu. Both problems have been studied intensively in the last decade, especially among which the uniqueness was already known in literature. We can refer to many works, but here only to [18, 16, 19] for the inverse tt-source problem and [10, 13, 12, 15] for the inverse 𝒙\bm{x}-source problem. For a comprehensive survey on Problem 1.1 especially before 2019, we refer to Liu, Li and Yamamoto [17].

However, it reveals that the existing papers mainly discuss regular temporal components μ\mu, e.g. in C1[0,T]C^{1}[0,T] or L2(0,T)L^{2}(0,T). Such restrictions exclude a wide class of singular functions represented by the Dirac delta function, which corresponds with point sources in practice. This encourages us to reconsider Problem 1.1 in function spaces with lower regularity. More precisely, in this article we are mainly concerned with the case of μ=μ(t)\mu=\mu(t) in a Sobolev space of negative order, in particular μL2(0,T)\mu\not\in L^{2}(0,T).

For such less regular μ\mu, we cannot expect the differentiability of u(𝒙,)u(\bm{x},\,\cdot\,) in time and we must redefine the Caputo derivative dtα\mathrm{d}_{t}^{\alpha} in (2). In this case, the unique existence of a solution to the initial-boundary value problem is more delicate, and an adequate formulation is needed.

This article is composed of 6 sections and 1 appendix. Redefining dtα\mathrm{d}_{t}^{\alpha} and thus problem (2) in negative Sobolev spaces, in Sect. 2 we state the main result in this paper. Then Sect. 34 are devoted to the two key ingredients for treating the singular system, i.e., the transfer to another regular system in L2L^{2} and Duhamel’s principle in Hα(0,T){}_{\alpha}H(0,T)-space. Next, the proof of Theorem 2.5 is completed in Sect. 5. Finally, Sect. 6 provides concluding remarks, and the proof of a technical detail is postponed to Appendix A.

2 Preliminary and Statement of the Main Result

To start with, we first define a fractional derivative for vL2(0,T)v\in L^{2}(0,T) which extends the domain of dtα\mathrm{d}_{t}^{\alpha} and formulate the initial-boundary value problem. To this end, we introduce function spaces and operators. Set the forward and backward Riemann-Liouville integral operators as

(Jαv)(t)\displaystyle(J_{\alpha}v)(t) :=1Γ(α)0t(ts)α1v(s)ds,0<t<T,𝒟(Jα)=L2(0,T),\displaystyle:=\frac{1}{\Gamma(\alpha)}\int^{t}_{0}(t-s)^{\alpha-1}v(s)\,\mathrm{d}s,\quad 0<t<T,\quad\mathcal{D}(J_{\alpha})=L^{2}(0,T),
(Jαv)(t)\displaystyle(J^{\alpha}v)(t) :=1Γ(α)tT(st)α1v(s)ds,0<t<T,𝒟(Jα)=L2(0,T).\displaystyle:=\frac{1}{\Gamma(\alpha)}\int^{T}_{t}(s-t)^{\alpha-1}v(s)\,\mathrm{d}s,\quad 0<t<T,\quad\mathcal{D}(J_{\alpha})=L^{2}(0,T).

We define an operator τ:L2(0,T)L2(0,T)\tau:L^{2}(0,T)\longrightarrow L^{2}(0,T) by (τv)(t):=v(Tt)(\tau v)(t):=v(T-t) and obviously τ\tau is an isomorphism. We set

C10[0,T]\displaystyle{}_{0}C^{1}[0,T] :={vC1[0,T]v(0)=0},\displaystyle:=\{v\in C^{1}[0,T]\mid v(0)=0\},
C10[0,T]\displaystyle{}^{0}C^{1}[0,T] :={vC1[0,T]v(T)=0}=τ(C10[0,T]).\displaystyle:=\{v\in C^{1}[0,T]\mid v(T)=0\}=\tau({}_{0}C^{1}[0,T]).

By Hα(0,T)H^{\alpha}(0,T) we denote the Sobolev-Slobodecki space with the norm Hα(0,T)\|\cdot\|_{H^{\alpha}(0,T)} defined by

vHα(0,T):=(vL2(0,T)2+0T0T|v(t)v(s)|2|ts|1+2αdtds)1/2\|v\|_{H^{\alpha}(0,T)}:=\left(\|v\|^{2}_{L^{2}(0,T)}+\int^{T}_{0}\!\!\!\int^{T}_{0}\frac{|v(t)-v(s)|^{2}}{|t-s|^{1+2\alpha}}\,\mathrm{d}t\mathrm{d}s\right)^{1/2}

(e.g., Adams [1]). Then we further introduce the function spaces

Hα(0,T):=C10[0,T]¯Hα(0,T),Hα(0,T):=C10[0,T]¯Hα(0,T).{}_{\alpha}H(0,T):=\overline{{}_{0}C^{1}[0,T]}^{H^{\alpha}(0,T)},\quad{}^{\alpha}H(0,T):=\overline{{}^{0}{C^{1}}[0,T]}^{H^{\alpha}(0,T)}.

Regarding the operators Jα,JαJ_{\alpha},J^{\alpha} and the spaces Hα(0,T),Hα(0,T){}_{\alpha}H(0,T),{}^{\alpha}H(0,T), we have the following lemma.

Lemma 2.1.

Let 0<α<10<\alpha<1.

(i) Jα:L2(0,T)Hα(0,T)J_{\alpha}:L^{2}(0,T)\longrightarrow{}_{\alpha}H(0,T) is bijective and isomorphism.

(ii) Jα:L2(0,T)Hα(0,T)J^{\alpha}:L^{2}(0,T)\longrightarrow{}^{\alpha}H(0,T) is bijective and isomorphism.

As for the proof of Lemma 2.1(i), see Gorenflo, Luchko and Yamamoto [8], Kubica, Ryszewska and Yamamoto [14]. Lemma 2.1(ii) can be readily derived from Lemma 2.1(i), the identity Jαv=τ(Jα(τv))J^{\alpha}v=\tau(J_{\alpha}(\tau v)) and the fact that τ\tau is an isomorphism.

Based on the spaces Hα(0,T){}_{\alpha}H(0,T) and Hα(0,T){}^{\alpha}H(0,T), let

Hα(0,T)L2(0,T)(Hα(0,T))=:Hα(0,T),\displaystyle{}_{\alpha}H(0,T)\subset L^{2}(0,T)\subset({}_{\alpha}H(0,T))^{\prime}=:{}_{-\alpha}H(0,T),
Hα(0,T)L2(0,T)(Hα(0,T))=:Hα(0,T)\displaystyle{}^{\alpha}H(0,T)\subset L^{2}(0,T)\subset({}^{\alpha}H(0,T))^{\prime}=:{}^{-\alpha}H(0,T)

be the Gel’fand triples, where ()(\,\cdot\,)^{\prime} denotes the dual space. Let X,YX,Y be Hilbert spaces and K:XYK:X\longrightarrow Y be a bounded linear operator with its domain 𝒟(K)=X\mathcal{D}(K)=X. Then we recall that KK^{\prime} is the maximal operator among K^:YX\widehat{K}:Y^{\prime}\longrightarrow X^{\prime} with 𝒟(K^)Y\mathcal{D}(\widehat{K})\subset Y^{\prime} such that

K^y,xXX=y,KxYY,xX,y𝒟(K^)Y.{}_{X^{\prime}}\langle\widehat{K}y,x\rangle_{X}={}_{Y^{\prime}}\langle y,Kx\rangle_{Y},\quad\forall\,x\in X,\ \forall\,y\in\mathcal{D}(\widehat{K})\subset Y^{\prime}.

Then we can prove the following lemma.

Lemma 2.2 (Yamamoto [25, Proposition 9]).

Let 0<β<10<\beta<1.

(i) (Jβ):Hβ(0,T)L2(0,T)(J_{\beta})^{\prime}:{}_{-\beta}H(0,T)\longrightarrow L^{2}(0,T) is bijective and isomorphism.

(ii) (Jβ):Hβ(0,T)L2(0,T)(J^{\beta})^{\prime}:{}^{-\beta}H(0,T)\longrightarrow L^{2}(0,T) is bijective and isomorphism.

(iii) There holds Jβ(Jβ)J_{\beta}\subset(J^{\beta})^{\prime}. In particular, Jβv=(Jβ)vJ_{\beta}v=(J^{\beta})^{\prime}v for vL2(0,T)v\in L^{2}(0,T).

Now we can redefine the Caputo derivative for functions in L2(0,T)L^{2}(0,T).

Definition 2.3.

We define

tα:=((Jα))1,𝒟(tα)=L2(0,T).\partial_{t}^{\alpha}:=((J^{\alpha})^{\prime})^{-1},\quad\mathcal{D}(\partial_{t}^{\alpha})=L^{2}(0,T).
Remark 2.4.

Kubica, Ryszewska and Yamamoto [14] defined ~tα\widetilde{\partial}_{t}^{\alpha} by

~tα:=(Jα)1,𝒟(~tα)=Hα(0,T).\widetilde{\partial}_{t}^{\alpha}:=(J_{\alpha})^{-1},\quad\mathcal{D}(\widetilde{\partial}_{t}^{\alpha})={}_{\alpha}H(0,T).

By Lemma 2.2(iii), we see tα~tα\partial_{t}^{\alpha}\supset\widetilde{\partial}_{t}^{\alpha} and

~tαv=tαv=limndtαvnin L2(0,T),vHα(0,T),\widetilde{\partial}_{t}^{\alpha}v=\partial_{t}^{\alpha}v=\lim_{n\to\infty}\mathrm{d}_{t}^{\alpha}v_{n}\quad\mbox{in }L^{2}(0,T),\ \forall\,v\in{}_{\alpha}H(0,T), (3)

where vnC10[0,T]v_{n}\in{}_{0}C^{1}[0,T] and vnvv_{n}\longrightarrow v in Hα(0,T)H^{\alpha}(0,T). Thus tα\partial_{t}^{\alpha} is an extension of ~tα\widetilde{\partial}_{t}^{\alpha}.

Now we are ready to propose the initial-boundary value problem:

(tα+A)u=μ(t)f(𝒙)in βH(0,T;L2(Ω)),\displaystyle(\partial_{t}^{\alpha}+A)u=\mu(t)f(\bm{x})\quad\mbox{in }^{-\beta}H(0,T;L^{2}(\Omega)), (4)
uL2(0,T;L2(Ω))Hβ(0,T;H2(Ω)H01(Ω)),\displaystyle u\in L^{2}(0,T;L^{2}(\Omega))\cap{}^{-\beta}{H(0,T;H^{2}(\Omega)\cap H^{1}_{0}(\Omega))}, (5)

where

μHβ(0,T),fL2(Ω),αβ<1,β>1/2.\mu\in{}^{-\beta}H(0,T),\quad f\in L^{2}(\Omega),\quad\alpha\leq\beta<1,\quad\beta>1/2. (6)

Henceforth we assume the existence of such a solution uu, although it can be verified during the succeeding arguments.

Now we are well prepared to state the main result on the uniqueness for Problem 1.1 with singularity in time.

Theorem 2.5.

Let uu satisfy (4)–(5) and assume (6). Let ωΩ\omega\subset\Omega and 𝐱0Ω\bm{x}_{0}\in\Omega be the same as that in Problem 1.1.

(i) Let μ0\mu\not\equiv 0 in (0,T)(0,T). If u=0u=0 in ω×(0,T),\omega\times(0,T), then f=0f=0 in L2(Ω)L^{2}(\Omega).

(ii) In (1) we assume that bj=0(j=1,,d)b_{j}=0\ (j=1,\ldots,d) and c0c\geq 0 in Ω\Omega. Let ff satisfy

f0 in Ωand(f0 or f0 in Ω),\displaystyle f\not\equiv 0\mbox{ in }\Omega\quad\mbox{and}\quad(f\geq 0\mbox{ or }f\leq 0\mbox{ in }\Omega), (7)
f{L2(Ω)if d=1,2,3,𝒟(Aθ)if d4,where θ>d/41.\displaystyle f\in\begin{cases}L^{2}(\Omega)&\mbox{if }d=1,2,3,\\ \mathcal{D}(A^{\theta})&\mbox{if }d\geq 4,\ \mbox{where }\theta>d/4-1.\end{cases} (8)

Then u=0u=0 at {𝐱0}×(0,T)\{\bm{x}_{0}\}\times(0,T) implies μ=0\mu=0 in Hβ(0,T){}^{-\beta}H(0,T).

In Theorem 2.5(ii), in terms of (8) we can verify

uHβ(0,T;C(Ω¯)).u\in{}^{-\beta}H(0,T;C(\overline{\Omega})). (9)

Therefore, the data u(𝒙0,)Hβ(0,T)u(\bm{x}_{0},\,\cdot\,)\in{}^{-\beta}H(0,T) makes sense. The proof of (9) is given in Appendix A.

In Theorem 2.5(ii), we can further study and prove the uniqueness in the case where we replace assumption (7) by more general conditions, e.g., bj0b_{j}\neq 0 and cc is not necessarily non-negative, but we omit details. Moreover, here we omit the treatments for the case 1<α<21<\alpha<2.

Since Hβ(0,T)C[0,T]{}^{\beta}H(0,T)\subset C[0,T] by β>1/2\beta>1/2, we see that the following μ\mu are in Hβ(0,T){}^{-\beta}H(0,T): μL1(0,T)\mu\in L^{1}(0,T) and μ(t)=k=1Nrkδak(t)\mu(t)=\sum_{k=1}^{N}r_{k}\delta_{a_{k}}(t), where rkr_{k}\in\mathbb{R}, 0<ak<T0<a_{k}<T and δak\delta_{a_{k}} is the Dirac delta function: δak,φ=φ(ak)\langle\delta_{a_{k}},\varphi\rangle=\varphi(a_{k}) for φC[0,T]\varphi\in C[0,T]. In particular, from Theorem 2.5(ii) we can directly derive

Corollary 2.6.

Under the same conditions as that in Theorem 2.5(ii), we set

μ(t)=k=1Nrkδak(t),=1,2,ak(0,T),rk{0},k=1,,N,\mu^{\ell}(t)=\sum_{k=1}^{N^{\ell}}r_{k}^{\ell}\delta_{a_{k}^{\ell}}(t),\quad\ell=1,2,\quad a_{k}^{\ell}\in(0,T),\quad r_{k}^{\ell}\in\mathbb{R}\setminus\{0\},\quad k=1,\ldots,N^{\ell},

where ak(k=1,2,,N)a_{k}^{\ell}\ (k=1,2,\ldots,N^{\ell}) are mutually distinct for =1,2\ell=1,2. Let uu^{\ell} be the solution to (4)–(5) with μ=μ(=1,2)\mu=\mu^{\ell}\ (\ell=1,2). Then u1(𝐱0,t)=u2(𝐱0,t)u^{1}(\bm{x}_{0},t)=u^{2}(\bm{x}_{0},t) for 0<t<T0<t<T implies N1=N2,N^{1}=N^{2}, rk1=rk2r_{k}^{1}=r_{k}^{2} and ak1=ak2(k=1,,N1)a_{k}^{1}=a_{k}^{2}\ (k=1,\ldots,N^{1}).

For the case of μL2(0,T)\mu\in L^{2}(0,T) or μL1(0,T)\mu\in L^{1}(0,T), there are rich references but we are here restricted to [10, 12, 13, 18]. In particular, [12] considers also the case 0<α20<\alpha\leq 2 and variable α(𝒙)\alpha(\bm{x}) for μL1(0,T)\mu\in L^{1}(0,T), and we can apply our method to their formulated inverse problems, as mentioned in Sect. 6.

3 Transfer to a Regular System

The main purpose of this section is the proof of

Proposition 3.1.

Let uu satisfy (4)–(5) with (6). Then

v:=(Jβ)uHβ(0,T;L2(Ω))L2(0,T;H2(Ω)H01(Ω))v:=(J^{\beta})^{\prime}u\in{}_{\beta}H(0,T;L^{2}(\Omega))\cap L^{2}(0,T;H^{2}(\Omega)\cap H^{1}_{0}(\Omega)) (10)

satisfies

(tα+A)v=(Jβ)μfin L2(0,T;L2(Ω)).(\partial_{t}^{\alpha}+A)v=(J^{\beta})^{\prime}\mu\,f\quad\mbox{in }L^{2}(0,T;L^{2}(\Omega)). (11)
Proof 3.2.

We show

Lemma 3.3.

Let α,γ>0\alpha,\gamma>0 and α+γ<1\alpha+\gamma<1. Then

(Jα+γ)v=(Jα)(Jγ)v=(Jγ)(Jα)vfor vHαγ(0,T).(J^{\alpha+\gamma})^{\prime}v=(J^{\alpha})^{\prime}(J^{\gamma})^{\prime}v=(J^{\gamma})^{\prime}(J^{\alpha})^{\prime}v\quad\mbox{for }v\in{}^{-\alpha-\gamma}H(0,T).
Proof 3.4.

We can directly verify

Jα+γ=JαJγ=JγJαin L2(0,T).J^{\alpha+\gamma}=J^{\alpha}J^{\gamma}=J^{\gamma}J^{\alpha}\quad\mbox{in }L^{2}(0,T).

Therefore, by e.g. Kato [11, Problem 5.26 (p.168)], we conclude the lemma.∎

We complete the proof of Proposition 3.1. Since uL2(0,T;L2(Ω))u\in L^{2}(0,T;L^{2}(\Omega)), by Lemma 2.2 and αβ\alpha\leq\beta, we see

(Jβ)u=JβuHβ(0,T;L2(Ω))Hα(0,T;L2(Ω)).(J^{\beta})^{\prime}u=J_{\beta}u\in{}_{\beta}H(0,T;L^{2}(\Omega))\subset{}_{\alpha}H(0,T;L^{2}(\Omega)).

Since uHα(0,T;H2(Ω)H01(Ω))u\in{}^{-\alpha}H(0,T;H^{2}(\Omega)\cap H_{0}^{1}(\Omega)), Lemma 2.2 yields (Jα)uL2(0,T;H2(Ω)H01(Ω))(J^{\alpha})^{\prime}u\in L^{2}(0,T;H^{2}(\Omega)\cap H_{0}^{1}(\Omega)). Then

(Jβ)u=(Jβα)(Jα)u=Jβα(Jα)u\displaystyle(J^{\beta})^{\prime}u=(J^{\beta-\alpha})^{\prime}(J^{\alpha})^{\prime}u=J_{\beta-\alpha}(J^{\alpha})^{\prime}u Hβα(0,T;H2(Ω)H01(Ω))\displaystyle\in{}_{\beta-\alpha}H(0,T;H^{2}(\Omega)\cap H_{0}^{1}(\Omega))
L2(0,T;H2(Ω)H01(Ω))\displaystyle\subset L^{2}(0,T;H^{2}(\Omega)\cap H_{0}^{1}(\Omega))

by Lemma 2.1(i), Lemma 2.2(iii) and (Jα)uL2(0,T;H2(Ω)H01(Ω))(J^{\alpha})^{\prime}u\in L^{2}(0,T;H^{2}(\Omega)\cap H_{0}^{1}(\Omega)). Therefore, we see vL2(0,T;H2(Ω)H01(Ω))v\in L^{2}(0,T;H^{2}(\Omega)\cap H_{0}^{1}(\Omega)).

Next we operate (Jβ)(J^{\beta})^{\prime} on both sides of (4) to deduce

(Jβ)((Jα))1u+A(Jβ)u=(Jβ)μf.(J^{\beta})^{\prime}((J^{\alpha})^{\prime})^{-1}u+A(J^{\beta})^{\prime}u=(J^{\beta})^{\prime}\mu\,f.

We will prove

(Jβ)((Jα))1u=((Jα))1(Jβ)u,(J^{\beta})^{\prime}((J^{\alpha})^{\prime})^{-1}u=((J^{\alpha})^{\prime})^{-1}(J^{\beta})^{\prime}u,

that is,

(Jα)(Jβ)((Jα))1u=(Jβ)u.(J^{\alpha})^{\prime}(J^{\beta})^{\prime}((J^{\alpha})^{\prime})^{-1}u=(J^{\beta})^{\prime}u.

In fact, by Lemma 3.3, we see (Jβ)=(Jβα)(Jα)=(Jα)(Jβα)=(Jβ)(J^{\beta})^{\prime}=(J^{\beta-\alpha})^{\prime}(J^{\alpha})^{\prime}=(J^{\alpha})^{\prime}(J^{\beta-\alpha})^{\prime}=(J^{\beta})^{\prime} and then

(Jα)(Jβ)((Jα))1u\displaystyle(J^{\alpha})^{\prime}(J^{\beta})^{\prime}((J^{\alpha})^{\prime})^{-1}u =(Jα)(Jβα)(Jα)((Jα))1u=(Jα)(Jβα)u\displaystyle=(J^{\alpha})^{\prime}(J^{\beta-\alpha})^{\prime}(J^{\alpha})^{\prime}((J^{\alpha})^{\prime})^{-1}u=(J^{\alpha})^{\prime}(J^{\beta-\alpha})^{\prime}u
=(Jβ)u.\displaystyle=(J^{\beta})^{\prime}u.

Therefore, it follows that

(Jβ)((Jα))1u=((Jα))1(Jβ)u=((Jα))1v=tαv.(J^{\beta})^{\prime}((J^{\alpha})^{\prime})^{-1}u=((J^{\alpha})^{\prime})^{-1}(J^{\beta})^{\prime}u=((J^{\alpha})^{\prime})^{-1}v=\partial_{t}^{\alpha}v.

Hence, we conclude tαv+Av=(Jβ)μf\partial_{t}^{\alpha}v+Av=(J^{\beta})^{\prime}\mu\,f.∎

In terms of Proposition 3.1, the unique existence of the solution uu to (4)–(5) can be clarified via v:=(Jβ)uv:=(J^{\beta})^{\prime}u in (10).

4 Duhamel’s Principle

In this section, we establish Duhamel’s principle which transforms a solution to an initial-boundary value problem without the inhomogeneous term to a solution to (11). Such a principle is known and see e.g. [18] and we can refer also to the survey [17] and Umarov [24]. Here we reformulate the formula in the space Hα(0,T){}_{\alpha}H(0,T) in order to apply within our framework.

Lemma 4.1 (Duhamel’s principle in Hα(0,T){}_{\alpha}H(0,T)).

Let gL2(0,T)g\in L^{2}(0,T) and fL2(Ω)f\in L^{2}(\Omega). Let zz satisfy

{tα(zf)+Az=0,zfHα(0,T;L2(Ω)),zL2(0,T;H2(Ω)H01(Ω)).\begin{cases}\partial_{t}^{\alpha}(z-f)+Az=0,\\ z-f\in{}_{\alpha}H(0,T;L^{2}(\Omega)),\quad z\in L^{2}(0,T;H^{2}(\Omega)\cap H^{1}_{0}(\Omega)).\end{cases} (12)

Then

w(,t)=0tg(s)z(,ts)ds,0<t<Tw(\,\cdot\,,t)=\int^{t}_{0}g(s)z(\,\cdot\,,t-s)\,\mathrm{d}s,\quad 0<t<T (13)

satisfies

{(tα+A)w=J1αgfin L2(0,T;L2(Ω)),wHα(0,T;L2(Ω))L2(0,T;H2(Ω)H01(Ω)).\begin{cases}(\partial_{t}^{\alpha}+A)w=J_{1-\alpha}g\,f\quad\mbox{in }L^{2}(0,T;L^{2}(\Omega)),\\ w\in{}_{\alpha}H(0,T;L^{2}(\Omega))\cap L^{2}(0,T;H^{2}(\Omega)\cap H^{1}_{0}(\Omega)).\end{cases} (14)
Proof 4.2.

Step 1. We prove

Lemma 4.3.

Let gL2(0,T)g\in L^{2}(0,T) and vHα(0,T)v\in{}_{\alpha}H(0,T). Then

tα(0tg(s)v(ts)ds)=0tg(s)tαv(ts)dsin L2(0,T).\partial_{t}^{\alpha}\left(\int^{t}_{0}g(s)v(t-s)\,\mathrm{d}s\right)=\int^{t}_{0}g(s)\partial_{t}^{\alpha}v(t-s)\,\mathrm{d}s\quad\mbox{in }L^{2}(0,T). (15)
Proof 4.4.

First we assume gC01[0,T]:={hC1[0,T]h(0)=h(T)=0}g\in C_{0}^{1}[0,T]:=\{h\in C^{1}[0,T]\mid h(0)=h(T)=0\} and vC10[0,T]v\in{}_{0}C^{1}[0,T]. Then by v(0)=0v(0)=0, we see

(0tg(s)v(ts)ds)=0tg(s)v(ts)dsL1(0,T)\left(\int^{t}_{0}g(s)v(t-s)\,\mathrm{d}s\right)^{\prime}=\int^{t}_{0}g(s)v^{\prime}(t-s)\,\mathrm{d}s\in L^{1}(0,T)

and thus

tα(0tg(s)v(ts)ds)\displaystyle\partial_{t}^{\alpha}\left(\int^{t}_{0}g(s)v(t-s)\,\mathrm{d}s\right) =dtα(0tg(s)v(ts)ds)\displaystyle=\mathrm{d}_{t}^{\alpha}\left(\int^{t}_{0}g(s)v(t-s)\,\mathrm{d}s\right)
=1Γ(1α)0t(ts)α(0sg(r)v(sr)dr)ds\displaystyle=\frac{1}{\Gamma(1-\alpha)}\int^{t}_{0}(t-s)^{-\alpha}\left(\int^{s}_{0}g(r)v^{\prime}(s-r)\,\mathrm{d}r\right)\mathrm{d}s
=0tg(r)(1Γ(1α)rt(ts)αv(sr)ds)dr\displaystyle=\int^{t}_{0}g(r)\left(\frac{1}{\Gamma(1-\alpha)}\int^{t}_{r}(t-s)^{-\alpha}v^{\prime}(s-r)\,\mathrm{d}s\right)\mathrm{d}r
=0tg(r)dtαv(tr)dr=0tg(r)tαv(tr)dr,\displaystyle=\int^{t}_{0}g(r)\,\mathrm{d}_{t}^{\alpha}v(t-r)\,\mathrm{d}r=\int^{t}_{0}g(r)\partial_{t}^{\alpha}v(t-r)\,\mathrm{d}r,

where we exchanged the orders of the integrals with respect to ss and rr. Therefore, (15) holds for each gC01[0,T]g\in C_{0}^{1}[0,T] and vC10[0,T]v\in{}_{0}C^{1}[0,T].

Next, let gL2(0,T)g\in L^{2}(0,T) and vHα(0,T)v\in{}_{\alpha}H(0,T). Since Hα(0,T)=C10[0,T]¯Hα(0,T){}_{\alpha}H(0,T)=\overline{{}_{0}C^{1}[0,T]}^{H^{\alpha}(0,T)} (e.g., [14]) and L2(0,T)=C01[0,T]¯L2(0,T)L^{2}(0,T)=\overline{C_{0}^{1}[0,T]}^{L^{2}(0,T)}, we can choose sequences {gn}C01[0,T]\{g_{n}\}\subset C_{0}^{1}[0,T] and {vn}C10[0,T]\{v_{n}\}\subset{}_{0}C^{1}[0,T] such that gngg_{n}\longrightarrow g in L2(0,T)L^{2}(0,T) and vnvv_{n}\longrightarrow v in Hα(0,T){}_{\alpha}H(0,T) as nn\to\infty. Henceforth we write (gv)(t)=0tg(s)v(ts)ds(g*v)(t)=\int^{t}_{0}g(s)v(t-s)\,\mathrm{d}s for 0<t<T0<t<T, and we regard tα\partial_{t}^{\alpha} as (Jα)1(J_{\alpha})^{-1}, that is, 𝒟(tα)=Hα(0,T)\mathcal{D}(\partial_{t}^{\alpha})={}_{\alpha}H(0,T) (see also (3)). As is directly proved, we have gnvnC10[0,T]g_{n}*v_{n}\in{}_{0}C^{1}[0,T] and so gnvnHα(0,T)=𝒟(tα)g_{n}*v_{n}\in{}_{\alpha}H(0,T)=\mathcal{D}(\partial_{t}^{\alpha}). Therefore, we see

tα(gnvn)(t)=(gntαvn)(t),0<t<T.\partial_{t}^{\alpha}(g_{n}*v_{n})(t)=(g_{n}*\partial_{t}^{\alpha}v_{n})(t),\quad 0<t<T.

Since vnvv_{n}\longrightarrow v in Hα(0,T){}_{\alpha}H(0,T), it follows that tαvntαv\partial_{t}^{\alpha}v_{n}\longrightarrow\partial_{t}^{\alpha}v in L2(0,T)L^{2}(0,T). Then Young’s convolution inequality yields gntαvngtαvg_{n}*\partial_{t}^{\alpha}v_{n}\longrightarrow g*\partial_{t}^{\alpha}v in L2(0,T)L^{2}(0,T). Therefore, tα(gnvn)\partial_{t}^{\alpha}(g_{n}*v_{n}) converges in L2(0,T)L^{2}(0,T) and tα(gv)=gtαv\partial_{t}^{\alpha}(g*v)=g*\partial_{t}^{\alpha}v in L2(0,T)L^{2}(0,T). The proof of Lemma 4.3 is complete.∎

Step 2. First we have

w(,t)=0tg(s)(z(,ts)f)ds+f0tg(s)ds,0<t<T.w(\,\cdot\,,t)=\int^{t}_{0}g(s)(z(\,\cdot\,,t-s)-f)\,\mathrm{d}s+f\int^{t}_{0}g(s)\,\mathrm{d}s,\quad 0<t<T.

Since 0tg(s)ds=(J1g)(t)\int^{t}_{0}g(s)\,\mathrm{d}s=(J_{1}g)(t) for 0<t<T0<t<T and (tαJ1g)(t)=(Jα1J1g)(t)=(J1αg)(t)(\partial_{t}^{\alpha}J_{1}g)(t)=(J_{\alpha}^{-1}J_{1}g)(t)=(J_{1-\alpha}g)(t), by Lemma 4.3 and zfHα(0,T;L2(Ω))z-f\in{}_{\alpha}H(0,T;L^{2}(\Omega)) we see

tαw(,t)=0tg(s)tα(zf)(,ts)ds+(J1αg)(t)f,0<t<T.\partial_{t}^{\alpha}w(\,\cdot\,,t)=\int^{t}_{0}g(s)\partial_{t}^{\alpha}(z-f)(\,\cdot\,,t-s)\,\mathrm{d}s+(J_{1-\alpha}g)(t)f,\quad 0<t<T.

Moreover, we have (Aw)(,t)=0tg(s)Az(,ts)ds(Aw)(\,\cdot\,,t)=\int^{t}_{0}g(s)Az(\,\cdot\,,t-s)\,\mathrm{d}s. Hence, we arrive at

(tα+A)w(,t)\displaystyle(\partial_{t}^{\alpha}+A)w(\,\cdot\,,t) =0tg(s)(tα(zf)+Az)(,ts)ds+(J1αg)(t)f\displaystyle=\int^{t}_{0}g(s)(\partial_{t}^{\alpha}(z-f)+Az)(\,\cdot\,,t-s)\,\mathrm{d}s+(J_{1-\alpha}g)(t)f
=(J1αg)(t)f,0<t<T.\displaystyle=(J_{1-\alpha}g)(t)f,\quad 0<t<T.

The regularity of ww described in (14) can follow directly from (13). Thus, by the uniqueness of solution to (14), the proof of Lemma 4.1 is complete.∎

5 Completion of the Proof of Theorem 2.5

We first show the following key lemma.

Lemma 5.1.

We assume

{(tα+A)v=gfin L2(0,T;L2(Ω)),vHα(0,T;L2(Ω))L2(0,T;H2(Ω)H01(Ω)),\begin{cases}(\partial_{t}^{\alpha}+A)v=g\,f\quad\mbox{in }L^{2}(0,T;L^{2}(\Omega)),\\ v\in{}_{\alpha}H(0,T;L^{2}(\Omega))\cap L^{2}(0,T;H^{2}(\Omega)\cap H^{1}_{0}(\Omega)),\end{cases}

where gL2(0,T)g\in L^{2}(0,T) and fL2(Ω)f\in L^{2}(\Omega). Let ωΩ\omega\subset\Omega and 𝐱0Ω\bm{x}_{0}\in\Omega be the same as that in Problem 1.1.

(i) Let g0g\not\equiv 0 in (0,T)(0,T). If v=0v=0 in ω×(0,T),\omega\times(0,T), then f0f\equiv 0 in Ω\Omega.

(ii) Let ff satisfy (7)–(8). If v=0v=0 at {𝐱0}×(0,T),\{\bm{x}_{0}\}\times(0,T), then g0g\equiv 0 in (0,T)(0,T).

Let Lemma 5.1 be proved. Then we can complete the proof of Theorem 2.5 as follows. Let uu be the solution to (4)–(5) and satisfy

u=0in ω×(0,T)oru=0at {𝒙0}×(0,T).u=0\quad\mbox{in }\omega\times(0,T)\qquad\mbox{or}\qquad u=0\quad\mbox{at }\{\bm{x}_{0}\}\times(0,T). (16)

Setting v:=(Jβ)uv:=(J^{\beta})^{\prime}u and g:=(Jβ)μL2(0,T)g:=(J^{\beta})^{\prime}\mu\in L^{2}(0,T), we have (10)–(11) and vHα(0,T;L2(Ω))v\in{}_{\alpha}H(0,T;L^{2}(\Omega)) according to Proposition 3.1. Moreover, (16) yields v=0v=0 in ω×(0,T)\omega\times(0,T) or v=0v=0 at {𝒙0}×(0,T)\{\bm{x}_{0}\}\times(0,T). Thus, since (Jβ)μ=0(J^{\beta})^{\prime}\mu=0 in L2(0,T)L^{2}(0,T) implies μ=0\mu=0 in Hβ(0,T){}^{-\beta}H(0,T), the application of Lemma 5.1 completes the proof of Theorem 2.5. Thus it suffices to prove Lemma 5.1.

Proof 5.2 (Proof of Lemma 5.1).

We set w:=J1αvw:=J_{1-\alpha}v. We note that by the injectivity of J1αJ_{1-\alpha}, it follows that v=0v=0 in ω×(0,T)\omega\times(0,T) and v=0v=0 at {𝐱0}×(0,T)\{\bm{x}_{0}\}\times(0,T) are equivalent to w=0w=0 in ω×(0,T)\omega\times(0,T) and w=0w=0 at {𝐱0}×(0,T)\{\bm{x}_{0}\}\times(0,T), respectively. Hence, we can reduce the proof to the following:

Let ww satisfy (14). Then

(i) Let g0g\not\equiv 0 in (0,T)(0,T). If w=0w=0 in ω×(0,T),\omega\times(0,T), then f0f\equiv 0 in Ω\Omega.

(ii) Let fL2(Ω)f\in L^{2}(\Omega) satisfy (7)–(8). If w=0w=0 at {𝐱0}×(0,T),\{\bm{x}_{0}\}\times(0,T), then g0g\equiv 0 in (0,T)(0,T).

Proof of (i). The proof is similar to that of Jiang, Li, Liu and Yamamoto [10, Theorem 2.6] and we describe the essence.

In terms of Lemma 4.1, we have

w(,t)=0tg(s)z(,ts)ds=0in ω, 0<t<T.w(\,\cdot\,,t)=\int^{t}_{0}g(s)z(\,\cdot\,,t-s)\,\mathrm{d}s=0\quad\mbox{in }\omega,\ 0<t<T.

The Titchmarsh convolution theorem (e.g., [23]) yields that there exists t[0,T]t_{*}\in[0,T] such that g=0g=0 in (0,Tt)(0,T-t_{*}) and z=0z=0 in ω×(0,t)\omega\times(0,t_{*}). Since g0g\not\equiv 0, we see that t>0t_{*}>0, indicating that z(,t)=0z(\,\cdot\,,t)=0 in ω\omega holds for tt in some open interval in \mathbb{R}. We apply a uniqueness result (e.g., [10] for non-symmetric AA) to obtain z0z\equiv 0 in Ω×(0,)\Omega\times(0,\infty). Consequently (12) implies tα(zf)0\partial_{t}^{\alpha}(z-f)\equiv 0 and thus zf=Jαtα(zf)0z-f=J_{\alpha}\partial_{t}^{\alpha}(z-f)\equiv 0 in Ω×(0,T)\Omega\times(0,T). By z0z\equiv 0 in Ω×(0,T)\Omega\times(0,T), we reach f0f\equiv 0 in Ω\Omega. This completes the proof of Lemma 5.1(i).

Proof of (ii). Henceforth, we set (f,g):=Ωf(𝐱)g(𝐱)d𝐱(f,g):=\int_{\Omega}f(\bm{x})g(\bm{x})\,\mathrm{d}\bm{x}.

We define AA by (1) with bj=0b_{j}=0 (j=1,,dj=1,\ldots,d), c0c\geq 0 in Ω\Omega and 𝒟(A)=H2(Ω)H01(Ω)\mathcal{D}(A)=H^{2}(\Omega)\cap H^{1}_{0}(\Omega). Then we number all of its eigenvalues with their multiplicities as

0<λ1λ2.0<\lambda_{1}\leq\lambda_{2}\leq\cdots\longrightarrow\infty.

Let φn\varphi_{n} be an eigenfunction for λn\lambda_{n}: Aφn=λnφnA\varphi_{n}=\lambda_{n}\varphi_{n} such that {φn}n\{\varphi_{n}\}_{n\in\mathbb{N}} forms an orthonormal basis in L2(Ω)L^{2}(\Omega).

Then, the fractional power AγA^{\gamma} is defined with γ>0\gamma>0, and

𝒟(Aγ)H2γ(Ω),fH2γ(Ω)C(n=1λn2γ|(f,φn)|2)1/2\mathcal{D}(A^{\gamma})\subset H^{2\gamma}(\Omega),\quad\|f\|_{H^{2\gamma}(\Omega)}\leq C\left(\sum_{n=1}^{\infty}\lambda_{n}^{2\gamma}|(f,\varphi_{n})|^{2}\right)^{1/2}

for γ0\gamma\geq 0 (e.g., Fujiwara [5], Pazy [20]). Moreover, we define the Mittag-Leffler functions Eα,β(z)E_{\alpha,\beta}(z) with α,β>0\alpha,\beta>0 by

Eα,β(z)=k=0zkΓ(αk+β),z,E_{\alpha,\beta}(z)=\sum_{k=0}^{\infty}\frac{z^{k}}{\Gamma(\alpha k+\beta)},\quad z\in\mathbb{C},

where the power series is uniformly and absolutely convergent in any compact set in \mathbb{C} (e.g., Gorenflo, Kilbas, Mainardi and Rogosin [7], Podlubny [21]). Then we can represent

z(,t)=n=1Eα,1(λntα)(f,φn)φnin L2(0,T;L2(Ω)).z(\,\cdot\,,t)=\sum_{n=1}^{\infty}E_{\alpha,1}(-\lambda_{n}t^{\alpha})(f,\varphi_{n})\varphi_{n}\quad\mbox{in }L^{2}(0,T;L^{2}(\Omega)). (17)

Then by (8) we can prove that for any fixed δ>0\delta>0 and T1>0T_{1}>0,

the series (17) is absolutely convergent in L(δ,T1;C(Ω¯))\mbox{the series \eqref{eq4.6} is absolutely convergent in }L^{\infty}(\delta,T_{1};C(\overline{\Omega})) (18)

and

z(𝒙0,)L1(0,T1).z(\bm{x}_{0},\,\cdot\,)\in L^{1}(0,T_{1}). (19)

Verification of (18) and (19). First let d=1,2,3d=1,2,3. Since

|Eα,1(λntα)|C1+λntα,n,t>0|E_{\alpha,1}(-\lambda_{n}t^{\alpha})|\leq\frac{C}{1+\lambda_{n}t^{\alpha}},\quad n\in\mathbb{N},\ t>0

(e.g., [21]), using (17), we have

Az(,t)L2(Ω)2\displaystyle\|Az(\,\cdot\,,t)\|^{2}_{L^{2}(\Omega)} =n=1λn2|Eα,1(λntα)|2|(f,φn)|2Cn=1|λn1+λntα|2|(f,φn)|2\displaystyle=\sum_{n=1}^{\infty}\lambda_{n}^{2}|E_{\alpha,1}(-\lambda_{n}t^{\alpha})|^{2}|(f,\varphi_{n})|^{2}\leq C\sum_{n=1}^{\infty}\left|\frac{\lambda_{n}}{1+\lambda_{n}t^{\alpha}}\right|^{2}|(f,\varphi_{n})|^{2}
Ct2αn=1|λntα1+λntα|2|(f,φn)|2Ct2αn=1|(f,φn)|2\displaystyle\leq C\,t^{-2\alpha}\sum_{n=1}^{\infty}\left|\frac{\lambda_{n}t^{\alpha}}{1+\lambda_{n}t^{\alpha}}\right|^{2}|(f,\varphi_{n})|^{2}\leq C\,t^{-2\alpha}\sum_{n=1}^{\infty}|(f,\varphi_{n})|^{2}
=Ct2αfL2(Ω)2,\displaystyle=C\,t^{-2\alpha}\|f\|^{2}_{L^{2}(\Omega)},

that is, z(,t)C(Ω¯)CtαfL2(Ω)\|z(\,\cdot\,,t)\|_{C(\overline{\Omega})}\leq C\,t^{-\alpha}\|f\|_{L^{2}(\Omega)} because

z(,t)C(Ω¯)Cz(,t)H2(Ω)CAz(,t)L2(Ω)\|z(\,\cdot\,,t)\|_{C(\overline{\Omega})}\leq C\|z(\,\cdot\,,t)\|_{H^{2}(\Omega)}\leq C^{\prime}\|Az(\,\cdot\,,t)\|_{L^{2}(\Omega)}

which is seen by d3d\leq 3 and the Sobolev embedding. Therefore, (18) and (19) are seen for d=1,2,3d=1,2,3.

Next let d4d\geq 4. Then, since θ>d/41\theta>d/4-1 by (8), the Sobolev embedding yields 𝒟(A1+θ)H2+2θ(Ω)C(Ω¯)\mathcal{D}(A^{1+\theta})\subset H^{2+2\theta}(\Omega)\subset C(\overline{\Omega}). Therefore,

|z(𝒙0,t)|\displaystyle|z(\bm{x}_{0},t)| Cz(,t)C(Ω¯)Cz(,t)H2+2θ(Ω)CA1+θz(,t)L2(Ω)\displaystyle\leq C\|z(\,\cdot\,,t)\|_{C(\overline{\Omega})}\leq C\|z(\,\cdot\,,t)\|_{H^{2+2\theta}(\Omega)}\leq C\|A^{1+\theta}z(\,\cdot\,,t)\|_{L^{2}(\Omega)}
=Cn=1λnEα,1(λntα)(Aθf,φn)φnL2(Ω)\displaystyle=C\left\|\sum_{n=1}^{\infty}\lambda_{n}E_{\alpha,1}(-\lambda_{n}t^{\alpha})(A^{\theta}f,\varphi_{n})\varphi_{n}\right\|_{L^{2}(\Omega)}
=C(n=1λn2|Eα,1(λntα)|2|(Aθf,φn)|2)1/2\displaystyle=C\left(\sum_{n=1}^{\infty}\lambda_{n}^{2}|E_{\alpha,1}(-\lambda_{n}t^{\alpha})|^{2}|(A^{\theta}f,\varphi_{n})|^{2}\right)^{1/2}
Ctα(n=1|λntα1+λntα|2|(Aθf,φn)|2)1/2CtαAθfL2(Ω),\displaystyle\leq C\,t^{-\alpha}\left(\sum_{n=1}^{\infty}\left|\frac{\lambda_{n}t^{\alpha}}{1+\lambda_{n}t^{\alpha}}\right|^{2}|(A^{\theta}f,\varphi_{n})|^{2}\right)^{1/2}\leq C\,t^{-\alpha}\|A^{\theta}f\|_{L^{2}(\Omega)},

so that the verification of (18) and (19) is complete.

Similarly to the proof of (i), if g0g\not\equiv 0 in (0,T)(0,T), then z(𝐱0,)0z(\bm{x}_{0},\,\cdot\,)\equiv 0 in (0,t)(0,t_{*}) with some constant t>0t_{*}>0. In terms of (18), we apply the tt-analyticity of z(𝐱0,t)z(\bm{x}_{0},t) (e.g., Sakamoto and Yamamoto [22]) to reach z(𝐱0,t)=0z(\bm{x}_{0},t)=0 for all t>0t>0. Therefore, we obtain

n=1Eα,1(λntα)(f,φn)φn(𝒙0)=0,t>δ,\sum_{n=1}^{\infty}E_{\alpha,1}(-\lambda_{n}t^{\alpha})(f,\varphi_{n})\varphi_{n}(\bm{x}_{0})=0,\quad t>\delta,

where the series is absolutely convergent in L(δ,T1)L^{\infty}(\delta,T_{1}) with arbitrary T1>0T_{1}>0.

Not counting the multiplicities, we rearrange all the eigenvalues of AA as

0<ρ1<ρ2<0<\rho_{1}<\rho_{2}<\cdots\longrightarrow\infty

and by {φnj}1jdn\{\varphi_{nj}\}_{1\leq j\leq d_{n}} we denote an orthonormal basis of ker(ρnA)\ker(\rho_{n}-A). In other words, {ρn}n\{\rho_{n}\}_{n\in\mathbb{N}} is the set of all distinct eigenvalues of AA. We set Pnf:=j=1dn(f,φnj)φnjP_{n}f:=\sum_{j=1}^{d_{n}}(f,\varphi_{nj})\varphi_{nj}. Hence we can write

n=1Eα,1(ρntα)(Pnf)(𝒙0)=0,t>δ.\sum_{n=1}^{\infty}E_{\alpha,1}(-\rho_{n}t^{\alpha})(P_{n}f)(\bm{x}_{0})=0,\quad t>\delta. (20)

On the other hand, we know

Eα,1(ρntα)=1Γ(1α)ρntα+O(1ρn2t2α)as tE_{\alpha,1}(-\rho_{n}t^{\alpha})=\frac{1}{\Gamma(1-\alpha)\rho_{n}t^{\alpha}}+O\left(\frac{1}{\rho_{n}^{2}t^{2\alpha}}\right)\quad\mbox{as }t\to\infty (21)

(e.g., [21, Theorem 1.4 (pp.33–34)]).

Since the series in (20) converges in L(δ,T1)L^{\infty}(\delta,T_{1}), extracting a subsequence of partial sums for the limit, we see that the subsequence of the partial sums is convergent for almost all t(δ,T1)t\in(\delta,T_{1}). Hence, by (21) we obtain

1Γ(1α)1tαn=1(Pnf)(𝒙0)ρn+1t2αn=1cn(Pnf)(𝒙0)=0\frac{1}{\Gamma(1-\alpha)}\frac{1}{t^{\alpha}}\sum_{n=1}^{\infty}\frac{(P_{n}f)(\bm{x}_{0})}{\rho_{n}}+\frac{1}{t^{2\alpha}}\sum_{n=1}^{\infty}c_{n}(P_{n}f)(\bm{x}_{0})=0

for almost all t(δ,)t\in(\delta,\infty), where |cn|O(ρn2)|c_{n}|\leq O(\rho_{n}^{-2}). Multiplying by tαt^{\alpha} and choosing a sequence {tm}\{t_{m}\} tending to \infty, we reach

n=1(Pnf)(𝒙0)ρn=0.\sum_{n=1}^{\infty}\frac{(P_{n}f)(\bm{x}_{0})}{\rho_{n}}=0.

Since it is assumed in Theorem 2.5(ii) that c0c\geq 0 in Ω\Omega, we see that A1A^{-1} exists and is a bounded operator from L2(Ω)L^{2}(\Omega) to itself and

A1f=n=1Pnfρnin L2(Ω).A^{-1}f=\sum_{n=1}^{\infty}\frac{P_{n}f}{\rho_{n}}\quad\mbox{in }L^{2}(\Omega).

Therefore, we conclude (A1f)(𝐱0)=0(A^{-1}f)(\bm{x}_{0})=0.

Set ψ:=A1fH2(Ω)H01(Ω)\psi:=A^{-1}f\in H^{2}(\Omega)\cap H^{1}_{0}(\Omega). Then Aψ=fA\psi=f in Ω\Omega and

ψ(𝒙0)=0.\psi(\bm{x}_{0})=0. (22)

Since c0c\geq 0 in (1) and f0f\geq 0 or f0f\leq 0 in Ω\Omega, using (22) and applying the strong maximum principle for AA (e.g., Gilbarg and Trudinger [6]), we conclude ψ0\psi\equiv 0 in Ω\Omega, that is, f0f\equiv 0 in Ω\Omega. This contradicts the assumption f0f\not\equiv 0 in Ω\Omega. Therefore, t>0t_{*}>0 is impossible. Hence t=0t_{*}=0 and so Titchmarsh convolution theorem yields g0g\equiv 0 in (0,T)(0,T). This completes the proof of Lemma 5.1(ii).∎

6 Concluding remarks

In this article, we consider an initial-boundary value problem for

(tα+A)u=μ(t)f(𝒙),(\partial_{t}^{\alpha}+A)u=\mu(t)f(\bm{x}), (23)

where μ=μ(t)\mu=\mu(t) is in a Sobolev space of negative order. The main machinery is to operate the extended Riemann-Liouville fractional integral operator (Jβ)(J^{\beta})^{\prime} (see Lemma 2.2 in Sect. 2) to reduce (23) to

(tα+A)v=((Jβ)μ)(t)f(𝒙),(\partial_{t}^{\alpha}+A)v=((J^{\beta})^{\prime}\mu)(t)f(\bm{x}), (24)

where v:=(Jβ)uv:=(J^{\beta})^{\prime}u and (Jβ)μL2(0,T)(J^{\beta})^{\prime}\mu\in L^{2}(0,T). Thus for the inverse source problems for (23), we can assume that μL2(0,T)\mu\in L^{2}(0,T) by replacing (23) by (24).

In this article, we limit the range of β\beta to (0,1)(0,1) for simplicity, but we can choose arbitrary β>0\beta>0. Therefore, for inverse source problems for (23), it is even sufficient to assume that μ\mu is smooth or μL(0,T)\mu\in L^{\infty}(0,T).

The same transformation for inverse source problems is valid for general time-fractional differential equations including fractional derivatives of variable orders αk(𝒙)\alpha_{k}(\bm{x}):

k=1Npk(𝒙)tαk(𝒙)u+Au=μ(t)f(𝒙)\sum_{k=1}^{N}p_{k}(\bm{x})\partial_{t}^{\alpha_{k}(\bm{x})}u+Au=\mu(t)f(\bm{x}) (25)

with suitable conditions on pk,αkp_{k},\alpha_{k}. In particular, also for (25), we can similarly discuss the determination of μHβ(0,T)\mu\in{}^{-\beta}H(0,T) with β>0\beta>0 by transforming μ\mu to a smooth function.

Appendix A Proof of (9)

Since g:=(Jβ)μL2(0,T)g:=(J^{\beta})^{\prime}\mu\in L^{2}(0,T) by μHβ(0,T)\mu\in{}^{-\beta}H(0,T), by Lemma 2.2(ii), it suffices to prove the solution v:=(Jβ)uL2(0,T;C(Ω¯))v:=(J^{\beta})^{\prime}u\in L^{2}(0,T;C(\overline{\Omega})) to (11) if ff satisfies (8). By [22] for example, we have

v(,t)=n=1(0tsα1Eα,α(λnsα)g(ts)ds)(f,φn)φn,0<t<T.v(\,\cdot\,,t)=\sum_{n=1}^{\infty}\left(\int^{t}_{0}s^{\alpha-1}E_{\alpha,\alpha}(-\lambda_{n}s^{\alpha})g(t-s)\,\mathrm{d}s\right)(f,\varphi_{n})\varphi_{n},\quad 0<t<T.

Moreover, we know

λntα1Eα,α(λntα)=ddtEα,1(λntα),t>0\lambda_{n}t^{\alpha-1}E_{\alpha,\alpha}(-\lambda_{n}t^{\alpha})=-\frac{\mathrm{d}}{\mathrm{d}t}E_{\alpha,1}(-\lambda_{n}t^{\alpha}),\quad t>0 (26)

and

tα1Eα,α(λntα)0,t>0.t^{\alpha-1}E_{\alpha,\alpha}(-\lambda_{n}t^{\alpha})\geq 0,\quad t>0. (27)

We can directly verify (26) by the termwise differentiation because Eα,1(λntα)E_{\alpha,1}(-\lambda_{n}t^{\alpha}) is an entire function, while (27) follows from the complete monotonicity of Eα,1(λntα)E_{\alpha,1}(-\lambda_{n}t^{\alpha}) (e.g., Gorenflo, Kilbas, Mainardi and Rogosin [7]).

Let d=1,2,3d=1,2,3. Then

Av(,t)L2(Ω)2=n=1|0tλnsα1Eα,α(λnsα)g(ts)ds|2|(f,φn)|2.\|Av(\,\cdot\,,t)\|_{L^{2}(\Omega)}^{2}=\sum_{n=1}^{\infty}\left|\int^{t}_{0}\lambda_{n}s^{\alpha-1}E_{\alpha,\alpha}(-\lambda_{n}s^{\alpha})g(t-s)\,\mathrm{d}s\right|^{2}|(f,\varphi_{n})|^{2}.

Hence, Young’s convolution inequality implies

0tλnsα1Eα,α(λnsα)g(ts)dsL2(0,T)\displaystyle\quad\,\left\|\int^{t}_{0}\lambda_{n}s^{\alpha-1}E_{\alpha,\alpha}(-\lambda_{n}s^{\alpha})g(t-s)\,\mathrm{d}s\right\|_{L^{2}(0,T)}
λntα1Eα,α(λntα)L1(0,T)gL2(0,T),\displaystyle\leq\left\|\lambda_{n}t^{\alpha-1}E_{\alpha,\alpha}(-\lambda_{n}t^{\alpha})\right\|_{L^{1}(0,T)}\|g\|_{L^{2}(0,T)}, (28)

and (26) and (27) yield

λntα1Eα,α(λntα)L1(0,T)\displaystyle\left\|\lambda_{n}t^{\alpha-1}E_{\alpha,\alpha}(-\lambda_{n}t^{\alpha})\right\|_{L^{1}(0,T)} =0Tλntα1Eα,α(λntα)dt\displaystyle=\int^{T}_{0}\lambda_{n}t^{\alpha-1}E_{\alpha,\alpha}(-\lambda_{n}t^{\alpha})\,\mathrm{d}t
=0TddtEα,1(λntα)dt\displaystyle=-\int^{T}_{0}\frac{\mathrm{d}}{\mathrm{d}t}E_{\alpha,1}(-\lambda_{n}t^{\alpha})\,\mathrm{d}t
=1Eα,1(λnTα)1.\displaystyle=1-E_{\alpha,1}(-\lambda_{n}T^{\alpha})\leq 1. (29)

Therefore, by (28) we see

0TAv(,t)L2(Ω)2dt\displaystyle\quad\,\int^{T}_{0}\|Av(\,\cdot\,,t)\|^{2}_{L^{2}(\Omega)}\,\mathrm{d}t
=n=10tλnsα1Eα,α(λnsα)g(ts)dsL2(0,T)2|(f,φn)|2\displaystyle=\sum_{n=1}^{\infty}\left\|\int^{t}_{0}\lambda_{n}s^{\alpha-1}E_{\alpha,\alpha}(-\lambda_{n}s^{\alpha})g(t-s)\,\mathrm{d}s\right\|^{2}_{L^{2}(0,T)}|(f,\varphi_{n})|^{2}
gL2(0,T)2n=1|(f,φn)|2gL2(0,T)2fL2(Ω)2.\displaystyle\leq\|g\|^{2}_{L^{2}(0,T)}\sum_{n=1}^{\infty}|(f,\varphi_{n})|^{2}\leq\|g\|^{2}_{L^{2}(0,T)}\|f\|^{2}_{L^{2}(\Omega)}.

Therefore, the Sobolev embedding H2(Ω)C(Ω¯)H^{2}(\Omega)\subset C(\overline{\Omega}) by d=1,2,3d=1,2,3, yields

vL2(0,T;C(Ω¯))2CgL2(0,T)2fL2(Ω)2,\|v\|^{2}_{L^{2}(0,T;C(\overline{\Omega}))}\leq C\|g\|^{2}_{L^{2}(0,T)}\|f\|^{2}_{L^{2}(\Omega)},

which means (9) for d=1,2,3d=1,2,3.

Let d4d\geq 4. Then we assume f𝒟(Aθ)f\in\mathcal{D}(A^{\theta}) with θ>d/41\theta>d/4-1. Since λnθφn=Aθφn\lambda_{n}^{\theta}\varphi_{n}=A^{\theta}\varphi_{n}, applying (28) and (29), we have

A1+θv(,t)L2(Ω)2=n=1|0tλnsα1Eα,α(λnsα)g(ts)ds|2|(Aθf,φn)|2,\|A^{1+\theta}v(\,\cdot\,,t)\|^{2}_{L^{2}(\Omega)}=\sum_{n=1}^{\infty}\left|\int^{t}_{0}\lambda_{n}s^{\alpha-1}E_{\alpha,\alpha}(-\lambda_{n}s^{\alpha})g(t-s)\,\mathrm{d}s\right|^{2}|(A^{\theta}f,\varphi_{n})|^{2},

and so

vL2(0,T;𝒟(A1+θ))2Cn=1gL2(0,T)2|(Aθf,φn)|2CgL2(0,T)2AθfL2(Ω)2.\|v\|_{L^{2}(0,T;\mathcal{D}(A^{1+\theta}))}^{2}\leq C\sum_{n=1}^{\infty}\|g\|^{2}_{L^{2}(0,T)}|(A^{\theta}f,\varphi_{n})|^{2}\leq C\|g\|^{2}_{L^{2}(0,T)}\|A^{\theta}f\|^{2}_{L^{2}(\Omega)}.

By the Sobolev embedding 𝒟(A1+θ)H2+2θ(Ω)C(Ω¯)\mathcal{D}(A^{1+\theta})\subset H^{2+2\theta}(\Omega)\subset C(\overline{\Omega}), we complete the proof of (9).

Acknowledgement  Y.​ Liu is supported by Grant-in-Aid for Early Career Scientists 20K14355 and 22K13954, JSPS. M.​ Yamamoto is supported by Grant-in-Aid for Scientific Research (A) 20H00117 and Grant-in-Aid for Challenging Research (Pioneering) 21K18142, JSPS.

References

  • [1] Adams, R.A.: Sobolev Spaces. Academic, New York (1975)
  • [2] Barlow, M.T., Perkins, E.A.: Brownian motion on the Sierpiński gasket. Probab. Theory Related Fields 79, 543–623 (1988). doi:10.1007/BF00318785
  • [3] Brown, T.S., Du, S., Eruslu, H., Sayas, F.J.: Analysis of models for viscoelastic wave propagation. Appl. Math. Nonlinear Sci. 3, 55–96 (2018). doi:10.21042/AMNS.2018.1.00006
  • [4] Eidelman, S.D., Kochubei, A.N.: Cauchy problem for fractional diffusion equations. J. Differential Equations 199, 211–255 (2004). doi:10.1016/j.jde.2003.12.002
  • [5] Fujiwara, D.: Concrete characterization of the domains of fractional powers of some elliptic differential operators of the second order. Proc. Japan Acad. 43, 82–86 (1967). doi:10.3792/pja/1195521686
  • [6] Gilbarg, D., Trudinger, N.S.: Elliptic Partial Differential Equations of Second Order. Springer, Berlin (2001)
  • [7] Gorenflo, R., Kilbas, A.A., Mainardi, F., Rogosin, S.V.: Mittag-Leffler Functions, Related Topics and Applications. Springer, Berlin (2014)
  • [8] Gorenflo, R., Luchko, Y., Yamamoto, M.: Time-fractional diffusion equation in the fractional Sobolev spaces. Fract. Calc. Appl. Anal. 18, 799–820 (2015). doi:10.1515/fca-2015-0048
  • [9] Hatano, Y., Hatano, N.: Dispersive transport of ions in column experiments: an explanation of long-tailed profiles. Water Resour. Res. 34, 1027–1033 (1998). doi:10.1029/98WR00214
  • [10] Jiang, D., Li, Z., Liu, Y., Yamamoto, M.: Weak unique continuation property and a related inverse source problem for time-fractional diffusion-advection equations. Inverse Problems 33, 055013 (2017). doi:10.1088/1361-6420/aa58d1
  • [11] Kato, T.: Perturbation Theory for Linear Operators. Springer, Berlin (1976)
  • [12] Kian, Y., Liu, Y., Yamamoto, M.: Uniqueness of inverse source problems for general evolution equations. Commun. Contemporary Math. (accepted). doi:10.1142/S0219199722500092
  • [13] Kian, Y., Soccorsi, É., Xue, Q., Yamamoto, M.: Identification of time-varying source term in time-fractional diffusion equations. Commun. Math. Sci. 20, 53–84 (2022). doi:10.4310/CMS.2022.v20.n1.a2
  • [14] Kubica, A., Ryszewska, K., Yamamoto, M.: Theory of Time-fractional Differential Equations: an Introduction. Springer, Tokyo (2020)
  • [15] Li, Z., Liu, Y., Yamamoto, M.: Inverse source problem for a one-dimensional time-fractional diffusion equation and unique continuation for weak solutions. Inverse Probl. Imaging (accepted). doi:10.3934/ipi.2022027
  • [16] Liu, Y.: Strong maximum principle for multi-term time-fractional diffusion equations and its application to an inverse source problem. Comput. Math. Appl. 73, 96–108 (2017). doi:10.1016/j.camwa.2016.10.021
  • [17] Liu, Y., Li, Z., Yamamoto, M.: Inverse problems of determining sources of the fractional partial differential equations. In: Kochubei, A., Luchko, Y. (eds.) Handbook of Fractional Calculus with Applications Volume 2: Fractional Differential Equations, pp. 411–430. De Gruyter, Berlin (2019). doi:10.1515/9783110571660-018
  • [18] Liu, Y., Rundell, W., Yamamoto, M.: Strong maximum principle for fractional diffusion equations and an application to an inverse source problem. Fract. Calc. Appl. Anal. 19, 888–906 (2016). doi:10.1515/fca-2016-0048
  • [19] Liu, Y., Zhang, Z.: Reconstruction of the temporal component in the source term of a (time-fractional) diffusion equation. J. Phys. A 50, 305203 (2017). doi:10.1088/1751-8121/aa763a
  • [20] Pazy, A.: Semigroups of Linear Operators and Applications to Partial Differential Equations. Springer, Berlin (1983).
  • [21] Podlubny, I.: Fractional Differential Equations. Academic Press, San Diego (1999).
  • [22] Sakamoto, K., Yamamoto, M.: Initial value/boundary value problems for fractional diffusion-wave equations and applications to some inverse problems. J. Math. Anal. Appl. 382, 426–447 (2011). doi:10.1016/j.jmaa.2011.04.058
  • [23] Titchmarsh, E.C.: The zeros of certain integral functions. Proc. Lond. Math. Soc. 25, 283–302 (1926). doi:10.1112/plms/s2-25.1.283
  • [24] Umarov, S.: Fractional Duhamel principle. In: Kochubei, A., Luchko, Y. (eds.) Handbook of Fractional Calculus with Applications Volume 2: Fractional Differential Equations, pp. 383–410. De Gruyter, Berlin (2019). doi:10.1515/9783110571660-017
  • [25] Yamamoto, M.: Fractional calculus and time-fractional differential equations: Revisit and construction of a theory. Mathematics 10, 698 (2022). doi:10.3390/math10050698