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

Convergence of a spectral regularization of a time-reversed reaction-diffusion problem with high-order Sobolev-Gevrey smoothness

Vo Anh Khoa Department of Mathematics, Florida A&M University, Tallahassee, FL 32307, USA anhkhoa.vo@famu.edu, vakhoa.hcmus@gmail.com
Abstract.

The present paper analyzes a spectral regularization of a time-reversed reaction-diffusion problem with globally and locally Lipschitz nonlinearities. This type of inverse and ill-posed problems arises in a variety of real-world applications concerning heat conduction and tumour source localization. In accordance with the weak solvability result for the forward problem, we focus on the inverse problem with high-order Sobolev-Gevrey smoothness and with Sobolev measurements. As expected from the well-known results for the linear case, we prove that this nonlinear spectral regularization possesses a logarithmic rate of convergence in a high-order Sobolev norm. The proof can be done by the verification of variational source condition; this way validates such a fine strategy in the framework of inverse problems for nonlinear partial differential equations. Ultimately, we study a semi-discrete version of the regularization method for a class of reaction-diffusion problems with non-degenerate nonlinearity. The convergence of this iterative scheme is also investigated.

Key words and phrases:
Inverse reaction-diffusion problem, spectral regularization, variational source condition, error estimates, Sobolev-Gevrey smoothness, iterations.
2000 Mathematics Subject Classification:
35R30, 65J20, 45L05, 65N15, 35B65

1. Introduction

1.1. Statement of the inverse problem

Here we are interested in a time-reversed reaction-diffusion model, denoted by ()\left(\mathcal{B}\right), with nonlinear source terms. Let Ω=[0,]d\Omega=\left[0,\ell\right]^{d} be a cube of d\mathbb{R}^{d} for dd\in\mathbb{N}. In this context, we consider a population density u=u(x,t)u=u\left(x,t\right) where (x,t)QT:=Ω×(0,T)\left(x,t\right)\in Q_{T}:=\Omega\times\left(0,T\right) with T>0T>0, obeying the following evolution equation:

(1.1) ut+𝒜u=F(u)in QT,u_{t}+\mathcal{A}u=F\left(u\right)\quad\text{in }Q_{T},

associated with the periodic boundary conditions, i.e. u(x+ei,)=u(x,)u\left(x+e_{i}\ell,\cdot\right)=u\left(x,\cdot\right) for 1id1\leq i\leq d where eie_{i} denotes the standard basis vector for d\mathbb{R}^{d}.

Here, 𝒜:=IΔ\mathcal{A}:=I-\Delta involves the linear second-order differential operator and thus accounts for the anisotropic diffusion of the population. The nonlinearity FF indicates either the deterministic reaction rate or the proliferation rate for some mechanism processes. Eventually, we complete the time-reversed model by the final condition

(1.2) u(x,t=T)=gT(x)in Ω.u\left(x,t=T\right)=g_{T}\left(x\right)\quad\text{in }\Omega.

Together with the periodic boundary condition, (1.1) and (1.2) structure our time-reversed reaction-diffusion model. In principle, the problem ()\left(\mathcal{B}\right) is well known to be severely ill-posed and has been investigated in a wide range of real-world applications; see e.g. [17, 10, 5] and references cited therein for an overview of recent results and existing models. Taking into account tumour models (see e.g. [10]), the physical meaning behind this problem ()\left(\mathcal{B}\right) is locating the tumour source by recovering the initial density of the tumour cells. In other words, the inverse problem we want to investigate in this paper is seeking the initial value u(x,t=0)=g0(x)u\left(x,t=0\right)=g_{0}\left(x\right) in a regular tissue Ω\Omega, provided that (1.1) and (1.2) are satisfied.

Naturally, we impose here the standard measurement on the final data (1.2), which reads as

(1.3) gTgTεL2(Ω)ε,\left\|g_{T}-g_{T}^{\varepsilon}\right\|_{L^{2}\left(\Omega\right)}\leq\varepsilon,

where ε>0\varepsilon>0 represents the deterministic noise level. In practice, one attempts to get the initial function g0(x)g_{0}\left(x\right) from this measured data gTεg_{T}^{\varepsilon} using some potential regularization.

Remark 1.

As noteworthy examples for the nonlinearity F(u)F\left(u\right) described in Model ()\left(\mathcal{B}\right), we take into account sigmoidal laws in population dynamics of cancer. The modest and simplest tumor growth is the logistic law, which can be generalized by the von Bertalanffy law with f(u)=aubuN+1f\left(u\right)=au-bu^{N+1} where a,ba,b and NN are specific non-negative numbers depending on every model. We also know that if the growth rate NN decays exponentially, the logistic growth turns out to be the so-called Gompertz law F(u)=aubuloguF\left(u\right)=au-bu\log u whenever uu satisfies some additional information to avoid the singularity of the logarithmic form. In more complex scenarios (e.g. two-species models), one usually agrees with the de Pillis-Radunskaya law standing for the fractional kill rate of tumor-specific effector cells, which reads as F(u)=auN/(b+uN)F\left(u\right)=au^{N}/\left(b+u^{N}\right). This is, furthermore, analogous to the generalized Michaelis–Menten law in enzyme kinetics. Another example can also be the Frank-Kamenetskii model in combustion theory, governed by the Arrhenius law of the form F(u)=exp(au)F\left(u\right)=\text{exp}\left(au\right). Accordingly, we see that investigating Model ()\left(\mathcal{B}\right) can be helpful in many areas of science and engineering.

1.2. Organization of the paper

In the following, we denote by H#pH_{\#}^{p} for pp\in\mathbb{N} the Hilbert spaces equipped with the standard norms and with the periodic boundary conditions posed in the domain. For σ0\sigma\geq 0, we define the Gevrey classes Gσp/2=𝒟(𝒜p/2eσ𝒜1/2)G_{\sigma}^{p/2}=\mathcal{D}\left(\mathcal{A}^{p/2}e^{\sigma\mathcal{A}^{1/2}}\right) where the operator 𝒜\mathcal{A} is defined in (1.1). These are also Hilbert spaces with respect to the inner product

v,wGσp/2=jdvjw¯j(1+|j|2)pe2σ(1+|j|2)1/2,\left\langle v,w\right\rangle_{G_{\sigma}^{p/2}}=\sum_{j\in\mathbb{Z}^{d}}v_{j}\cdot\bar{w}_{j}\left(1+\left|j\right|^{2}\right)^{p}e^{2\sigma\left(1+\left|j\right|^{2}\right)^{1/2}},

and then with the corresponding norm

vGσp/2=(jd|vj|2(1+|j|2)pe2σ(1+|j|2)1/2)1/2.\left\|v\right\|_{G_{\sigma}^{p/2}}=\left(\sum_{j\in\mathbb{Z}^{d}}\left|v_{j}\right|^{2}\left(1+\left|j\right|^{2}\right)^{p}e^{2\sigma\left(1+\left|j\right|^{2}\right)^{1/2}}\right)^{1/2}.

With this setting, it is worth mentioning the weak solvability of the forward problem of ()\left(\mathcal{B}\right) where FF is real analytic. Note that cf. [2], the real analyticity of FF means that if it can be represented by F(u)=j=0ajujF\left(u\right)=\sum_{j=0}^{\infty}a_{j}u^{j} for aja_{j}\in\mathbb{R}, then the corresponding majorising series j=0|aj|uj\sum_{j=0}^{\infty}\left|a_{j}\right|u^{j} is convergent for any uu\in\mathbb{R}.

Theorem 2.

[2, Theorem 1] Assume the initial data g0H#p(Ω)g_{0}\in H_{\#}^{p}\left(\Omega\right) for p>d/2p>d/2 and the nonlinearity FF is real analytic. Then there exists a time T>0T^{*}>0 such that the forward model of Model ()\left(\mathcal{B}\right) has a unique regular solution uu in the sense that uC([0,T];H#p(Ω))L2(0,T;𝒟(𝒜))u\in C\left(\left[0,T^{*}\right];H_{\#}^{p}\left(\Omega\right)\right)\cap L^{2}\left(0,T^{*};\mathcal{D}\left(\mathcal{A}\right)\right) satisfying utL2(0,T;L2(Ω))u_{t}\in L^{2}\left(0,T^{*};L^{2}\left(\Omega\right)\right) and

ut,v(H1(Ω)),H1(Ω)+𝒜1/2u,𝒜1/2vL2(Ω)+F(u),vL2(Ω)=0,\left\langle u_{t},v\right\rangle_{\left(H^{1}(\Omega)\right)^{\prime},H^{1}(\Omega)}+\left\langle\mathcal{A}^{1/2}u,\mathcal{A}^{1/2}v\right\rangle_{L^{2}\left(\Omega\right)}+\left\langle F\left(u\right),v\right\rangle_{L^{2}\left(\Omega\right)}=0,

for all vH#1(Ω)v\in H_{\#}^{1}\left(\Omega\right) and for a.e. t(0,T)t\in\left(0,T^{*}\right). Furthermore, this regular solution satisfies u(,t)Gtp/2(Ω)u\left(\cdot,t\right)\in G_{t}^{p/2}\left(\Omega\right) for t[0,T)t\in\left[0,T^{*}\right).

As a by-product of Theorem 2, we can show that u(,t)GTp/2(Ω)u\left(\cdot,t\right)\in G_{T^{*}}^{p/2}\left(\Omega\right) for all tTt\geq T^{*} and then F(u)GTp/2(Ω)F\left(u\right)\in G_{T^{*}}^{p/2}\left(\Omega\right) by virtue of the Taylor series expansion of FF. Therefore, in this work we assume our final time of observation TT is such that T<TT<T^{*}. Since GTp/2(Ω)Hp(Ω)G_{T^{*}}^{p/2}\left(\Omega\right)\subset H^{p}\left(\Omega\right), it is reasonable to assume in ()\left(\mathcal{B}\right) that

(1.4) gTε,gTHp(Ω).g_{T}^{\varepsilon},g_{T}\in H^{p}\left(\Omega\right).

This paper is devoted to the convergence analysis of a modified cut-off regularization of the inverse problem ()\left(\mathcal{B}\right). In this regard, we comply with the weak solvability of the forward model to take into account the Gevrey-Sobolev source conditions for the inverse problem. In other words, together with Assumption (1.4) we make use of the following source condition:

(1.5) uC([0,T];H#p(Ω))L2(0,T;𝒟(𝒜))and u(,t)Gtp/2(Ω)for t[0,T),u\in C\left(\left[0,T\right];H_{\#}^{p}\left(\Omega\right)\right)\cap L^{2}\left(0,T;\mathcal{D}\left(\mathcal{A}\right)\right)\;\text{and }u\left(\cdot,t\right)\in G_{t}^{p/2}\left(\Omega\right)\;\text{for }t\in\left[0,T\right),

It is worth mentioning that in the forward process, we obtain the solution in C([0,T];H#p(Ω))C\left(\left[0,T^{*}\right];H_{\#}^{p}\left(\Omega\right)\right) with p>d/2p>d/2, which indicates the fact that the solution belongs to C([0,T];L(Ω))C\left(\left[0,T\right];L^{\infty}\left(\Omega\right)\right). Instead of working with the real analyticity of FF, we can further apply the mean value theorem to get

(1.6) |F(u)F(v)|(sup|w|M|Fw(w)|)|uv|,\left|F\left(u\right)-F\left(v\right)\right|\leq\left(\sup_{\left|w\right|\leq M}\left|\frac{\partial F}{\partial w}\left(w\right)\right|\right)\left|u-v\right|,

where for u,vC([0,T];L(Ω))u,v\in C\left(\left[0,T\right];L^{\infty}\left(\Omega\right)\right) we have denoted by

M=2max{uC([0,T];L(Ω)),vC([0,T];L(Ω))}>0.M=2\max\left\{\left\|u\right\|_{C\left(\left[0,T\right];L^{\infty}\left(\Omega\right)\right)},\left\|v\right\|_{C\left(\left[0,T\right];L^{\infty}\left(\Omega\right)\right)}\right\}>0.

To this end, we will thus exploit Assumption (1.6) through the analysis of Section 2. We also assume that there exists a continuous function L(M)>0L\left(M\right)>0 such that

(1.7) sup|w|M|Fw(w)|L(M).\sup_{\left|w\right|\leq M}\left|\frac{\partial F}{\partial w}\left(w\right)\right|\leq L\left(M\right).

Naturally, observing the applications mentioned in Remark 1 we have

  • for the von Bertalanffy law: L(M)=|a|+|b|(N+1)MNL\left(M\right)=\left|a\right|+\left|b\right|\left(N+1\right)M^{N};

  • for the Gompertz law: L(M)=|a|+|b|ML\left(M\right)=\left|a\right|+\left|b\right|M

  • for the de Pillis-Radunskaya law: L(M)=|ab|NMN1L\left(M\right)=\left|\frac{a}{b}\right|NM^{N-1} for b0b\neq 0;

  • for the Arrhenius law: L(M)=|a|L\left(M\right)=\left|a\right| if a<0a<0 and L(M)=aeaML\left(M\right)=ae^{aM} if a0a\geq 0.

In this work, all the constants CC used here are independent of the measurement error ε\varepsilon. Nonetheless, their precise values may change from line to line and even within a single chain of estimates. On the other side, we use either the superscript or the subscript ε\varepsilon to accentuate the possible dependence of the present error on the constants.

In the first step, we derive the error bounds when L(M)=CL\left(M\right)=C is independent of MM. Our proof relies on the way we verify a variational source condition. This interestingly contributes to one of their first applications to the convergence analysis of regularization of inverse problems for nonlinear PDEs. In this approach, we prove the usual logarithmic-type rate of convergence of the nonlinear spectral regularization. We remark that during the time evolving backward in the open set (0,T)\left(0,T\right), the nonlinear scheme yields the asymptotic Hölder rate. When L(M)L\left(M\right) essentially depends on MM, the convergence is slower due to the phenomenal growth of the quantity L(M)L\left(M\right) involved in the Lipschitz property (1.6). Technically, the impediment of this growth, albeit its high-impact on the structure of the variational source condition, can be solved using a careful choice of a cut-off function for the nonlinearity FF. In this sense, the regularized solution is sought in a proper open set decided by the measurement parameter ε\varepsilon. Eventually, a slower logarithmic-type convergence is also expected, which extends the results in [8, 17, 15] and references cited therein. Typically, this extension illuminates that the Gevrey regularity conventionally restricted to the convergence analysis in [17, 15] is applicable when solving a class of time-reversed PDEs. We also add that this work partly completes the gap of verifying variational source conditions for a class of nonlinear PDEs, which is still questioned in [8].

In the second theme, our concentration moves to a derivation of an iteration-based version of the nonlinear spectral scheme with a stronger version of (1.6). We end up with the convergence of the approximate scheme by exploring the choice of the so-called stabilization constant, which depends not only on the number of iterations, but also on the measurement ε\varepsilon.

Prior to the setting of our approach and to closing this section, we introduce what the variational source condition concerns in principle as it is exploited in the skeleton of our proof.

1.3. Background of the variational source condition

Consider the ill-posed operator equation 𝒯(f)=g\mathcal{T}\left(f\right)=g where 𝒯\mathcal{T} maps from 𝒟(𝒯)𝕏\mathcal{D}\left(\mathcal{T}\right)\subset\mathbb{X} to 𝕐\mathbb{Y} with 𝕏\mathbb{X} and 𝕐\mathbb{Y} being Hilbert spaces. In this Hilbert setting, we denote by f𝒟(𝒯)f^{\dagger}\in\mathcal{D}\left(\mathcal{T}\right) the exact solution and by gε𝕐g^{\varepsilon}\in\mathbb{Y} the noisy data satisfying 𝒯(f)gε𝕐ε\left\|\mathcal{T}\left(f^{\dagger}\right)-g^{\varepsilon}\right\|_{\mathbb{Y}}\leq\varepsilon, where ε>0\varepsilon>0 represents the deterministic noise level (cf. e.g (1.3)). There are several stable approximations to regularize such equations. One of the most effective methods is Tikhonov regularization in which the exact solution is approximated by a solution of the minimization problem, viz.

(1.8) fαεargminf𝒟(𝒯)[𝒯(f)gε𝕐2+αf𝕏2]=:α(gε),f_{\alpha}^{\varepsilon}\in\underset{f\in\mathcal{D}\left(\mathcal{T}\right)}{\text{argmin}}\left[\left\|\mathcal{T}\left(f\right)-g^{\varepsilon}\right\|_{\mathbb{Y}}^{2}+\alpha\left\|f\right\|_{\mathbb{X}}^{2}\right]=:\mathcal{R}_{\alpha}\left(g^{\varepsilon}\right),

for some regularization parameter α:=α(ε)>0\alpha:=\alpha\left(\varepsilon\right)>0.

In regularization theory, one can prove that α(gε)α0+𝒯1\mathcal{R}_{\alpha}\left(g^{\varepsilon}\right)\stackrel{{\scriptstyle\alpha\searrow 0^{+}}}{{\longrightarrow}}\mathcal{T}^{-1} for suitable choices of α\alpha in some adequate topology. Furthermore, people usually want to specify rates of convergence, which turns out to be the question of finding the worst case error, since such rates are arbitrarily slow in general. Denote by ψ:[0,)[0,)\psi:\left[0,\infty\right)\to\left[0,\infty\right) an index function if it is continuous and monotonically increasing with ψ(0)=0\psi\left(0\right)=0. In this sense, one aims to:

Find (usually compact) subspaces 𝕂𝕏\mathbb{K}\subset\mathbb{X} such that there exists an index function ψ\psi satisfying for all f𝕂f\in\mathbb{K}, it holds sup{α(gε)f𝕏:𝒯(f)gε𝕐ε}ψ(ε).\sup\left\{\left\|\mathcal{R}_{\alpha}\left(g^{\varepsilon}\right)-f\right\|_{\mathbb{X}}:\left\|\mathcal{T}\left(f\right)-g^{\varepsilon}\right\|_{\mathbb{Y}}\leq\varepsilon\right\}\leq\psi\left(\varepsilon\right).

Nevertheless, we often need additional a priori assumptions on the exact solution ff^{\dagger} to gain the speed of convergence, which are called source conditions. In the literature, the former type of such conditions singles out that f=ψ(𝒯[f]𝒯[f])wf^{\dagger}=\psi\left(\mathcal{T}^{\prime}\left[f^{\dagger}\right]*\mathcal{T}^{\prime}\left[f^{\dagger}\right]\right)w for some w𝕏w\in\mathbb{X}, renowned as the spectral source condition in the community of inverse and ill-posed problems for several years; see e.g. [11] for a prevailing background of source conditions. Here ψ\psi could be of the Hölder and logarithmic types, depending on every situation. More recently, a novel formulation for source conditions has been derived in the sense of variational inequality, which reads as

(1.9) 2f,ff𝕏12ff𝕏2+ψ(𝒯(f)𝒯(f)𝕐2)for all f𝒟(𝒯)𝕏.2\left\langle f^{\dagger},f^{\dagger}-f\right\rangle_{\mathbb{X}}\leq\frac{1}{2}\left\|f-f^{\dagger}\right\|_{\mathbb{X}}^{2}+\psi\left(\left\|\mathcal{T}\left(f\right)-\mathcal{T}\left(f^{\dagger}\right)\right\|_{\mathbb{Y}}^{2}\right)\quad\text{for all }f\in\mathcal{D}\left(\mathcal{T}\right)\subset\mathbb{X}.

Cf. [6], it is in particular called as the variational source condition. Essentially, we can benefit from this source condition not only to simplify proofs for convergence rates in Hilbert frameworks, but also to extensively work in some Banach setting and in more general models with distinctive measurements (cf. e.g. [9, 4, 12]). Compared to the spectral source condition, it does not require the Fréchet derivative 𝒯\mathcal{T}^{\prime} and would thus be helpful in certain applications where the forward operator is not sufficiently smooth. Most importantly, variational source conditions have been shown to be well-adapted to bounded linear operators in Hilbert spaces; see [3]. On top of that, the variational source condition combined with some nonlinearity condition has been formulated in e.g. [16] to treat nonlinear operators.

Even though there have been enormous advantages over the spectral source conditions, variational source conditions could barely be verified in particular models in terms of partial differential equations. So far, with the aid of complex geometrical optics solutions the variational source condition holds true for an inverse medium scattering problem (cf. [7]) with ψ\psi in the logarithmic form when the solution belongs to some Sobolev ball. In this paper, we aim to show that it holds true for more complex and nonlinear models. It is worth mentioning that characterizations of the variational source conditions have been obtained in [8] where the degree of ill-posedness and the smoothness of solution are established for arbitrary families of subspaces. Denoting by (𝕏)\mathcal{L}\left(\mathbb{X}\right) the set of all bounded linear transformations mapping from 𝕏\mathbb{X} onto itself, the characterizations are provided cf. [8, Theorem 2.1] in the following.

Suppose there exists a family of orthogonal projections 𝒫r(𝕏)\mathcal{P}_{r}\in\mathcal{L}\left(\mathbb{X}\right) indexed by a parameter rr in some index set 𝒥\mathcal{J} such that for some functions κ,σ:𝒥(0,)\kappa,\sigma:\mathcal{J}\to\left(0,\infty\right) and some c0c\geq 0 the following conditions hold true for all r𝒥r\in\mathcal{J}: (C1)\left(\text{C}_{1}\right) f𝒫rf𝕏κ(r)\left\|f^{\dagger}-\mathcal{P}_{r}f^{\dagger}\right\|_{\mathbb{X}}\leq\kappa\left(r\right), (C2)\left(\text{C}_{2}\right) f,𝒫r(ff)𝕏σ(r)𝒯(f)𝒯(f)𝕐+cκ(r)ff𝕏\left\langle f^{\dagger},\mathcal{P}_{r}\left(f^{\dagger}-f\right)\right\rangle_{\mathbb{X}}\leq\sigma\left(r\right)\left\|\mathcal{T}\left(f^{\dagger}\right)-\mathcal{T}\left(f\right)\right\|_{\mathbb{Y}}+c\kappa\left(r\right)\left\|f^{\dagger}-f\right\|_{\mathbb{X}} for any f𝒟(𝒯)f\in\mathcal{D}\left(\mathcal{T}\right) with ff𝕏4f𝕏\left\|f-f^{\dagger}\right\|_{\mathbb{X}}\leq 4\left\|f^{\dagger}\right\|_{\mathbb{X}}. Then ff^{\dagger} satisfies the variational source condition (1.9) with ψ(δ):=2infr𝒥[(c+1)2|κ(r)|2+σ(r)δ],\psi\left(\delta\right):=2\inf_{r\in\mathcal{J}}\left[\left(c+1\right)^{2}\left|\kappa\left(r\right)\right|^{2}+\sigma\left(r\right)\sqrt{\delta}\right], and ψ\psi is a concave index function if infr𝒥κ(r)=0\inf_{r\in\mathcal{J}}\kappa\left(r\right)=0.

2. Spectral cut-off projection

2.1. Settings of the cut-off approach

In line with the characterizations of variational source conditions, in this section we verify and improve the cut-off projection that has been postulated in [15, 18], using Assumptions (1.3), (1.4), (1.5), (1.6), (1.7). The nature of this approach is that one solves the problem in a finite dimensional subspace, which turns out to be a well-posed problem effectively controlled by the measurement error.

We denote by {Eλ}\left\{E_{\lambda}\right\} the spectral family of the positive operator 𝒜\mathcal{A} defined in the backward problem ()\left(\mathcal{B}\right) and the function λEλv\lambda\mapsto\left\|E_{\lambda}v\right\| is called the spectral distribution function of vL2(Ω)v\in L^{2}\left(\Omega\right). Thereby, with Cε>0C_{\varepsilon}>0 being a cut-off parameter that will be chosen appropriately, we introduce the spectral cut-off projection Eλε:=𝟏(0,λε](𝒜)E_{\lambda_{\varepsilon}}:=\mathbf{1}_{\left(0,\lambda_{\varepsilon}\right]}\left(\mathcal{A}\right) where 𝟏(0,λε]\mathbf{1}_{\left(0,\lambda_{\varepsilon}\right]} denotes the characteristic function of the interval (0,λε]\left(0,\lambda_{\varepsilon}\right] with λε=Cε\left\lceil\lambda_{\varepsilon}\right\rceil=C_{\varepsilon}. To establish the stable approximate problem of ()\left(\mathcal{B}\right), we adapt the cut-off projection to both the nonlinearity FF and the final data gTεg_{T}^{\varepsilon}. When doing so, we characterize the abstract Gevrey class Gσp/2G_{\sigma}^{p/2} for σ0\sigma\geq 0 and pp\in\mathbb{N} as

Gσp/2={vL2(Ω):0λpe2σλdEλv2<},G_{\sigma}^{p/2}=\left\{v\in L^{2}\left(\Omega\right):\int_{0}^{\infty}\lambda^{p}e^{2\sigma\lambda}d\left\|E_{\lambda}v\right\|^{2}<\infty\right\},

equipped with the norm

vGσp/22=0λpe2σλdEλv2<.\left\|v\right\|_{G_{\sigma}^{p/2}}^{2}=\int_{0}^{\infty}\lambda^{p}e^{2\sigma\lambda}d\left\|E_{\lambda}v\right\|^{2}<\infty.

Henceforward, for each ε>0\varepsilon>0 we take into account the following approximate problem, denoted by (ε)\left(\mathcal{B}_{\varepsilon}\right),

{ut+𝒜u=0λεp/2𝑑EλεF(u)in QT,u(T)=0λεp/2𝑑EλεgTεin Ω.\begin{cases}u_{t}+\mathcal{A}u=\int_{0}^{\infty}\lambda_{\varepsilon}^{p/2}dE_{\lambda_{\varepsilon}}F\left(u\right)&\text{in }Q_{T},\\ u\left(T\right)=\int_{0}^{\infty}\lambda_{\varepsilon}^{p/2}dE_{\lambda_{\varepsilon}}g_{T}^{\varepsilon}&\text{in }\Omega.\end{cases}

We define a mildly weak solution of (ε)\left(\mathcal{B}_{\varepsilon}\right) (or a mild solution for short) to be a continuous mapping uε:[0,T]Hp(Ω)u_{\varepsilon}:\left[0,T\right]\to H^{p}\left(\Omega\right) obeying the integral equation

(2.1) uε(t)=0e(Tt)λελεp/2𝑑EλεgTεtT0e(st)λελεp/2𝑑EλεF(uε)(s)𝑑s.u_{\varepsilon}\left(t\right)=\int_{0}^{\infty}e^{\left(T-t\right)\lambda_{\varepsilon}}\lambda_{\varepsilon}^{p/2}dE_{\lambda_{\varepsilon}}g_{T}^{\varepsilon}-\int_{t}^{T}\int_{0}^{\infty}e^{\left(s-t\right)\lambda_{\varepsilon}}\lambda_{\varepsilon}^{p/2}dE_{\lambda_{\varepsilon}}F\left(u_{\varepsilon}\right)\left(s\right)ds.

Observe that in (2.1) we actually obtain an integral equation of the form uε(t)=𝒢(uε)(t)u_{\varepsilon}\left(t\right)=\mathcal{G}\left(u_{\varepsilon}\right)\left(t\right) where 𝒢\mathcal{G} is completely formulated by the right-hand side, mapping from C([0,T];Hp(Ω))C\left(\left[0,T\right];H^{p}\left(\Omega\right)\right) onto itself. Therefore, the existence and uniqueness results for (2.1) can be done by the standard fixed-point argument, requiring that the number of fixed-point iterations must be larger than the stability magnitude usually involved in the approximate problem. Since these results are standard, we, for simplicity, refer the reader to the concrete reference [17] for the detailed notion of proof.

Observe again from (2.1) that we can define the cut-off operator 𝒫ε(Hp(Ω))\mathcal{P}_{\varepsilon}\in\mathcal{L}\left(H^{p}\left(\Omega\right)\right) for the solution u(,t0)u\left(\cdot,t_{0}\right) by fixing t=t0t=t_{0}. In particular, it is an orthogonal projection given by

(2.2) 𝒫εu(,t0)=0e(Tt0)λελεp/2𝑑EλεgTεt0T0e(st0)λελεp/2𝑑EλεF(u)(s)𝑑s.\mathcal{P}_{\varepsilon}u\left(\cdot,t_{0}\right)=\int_{0}^{\infty}e^{\left(T-t_{0}\right)\lambda_{\varepsilon}}\lambda_{\varepsilon}^{p/2}dE_{\lambda_{\varepsilon}}g_{T}^{\varepsilon}-\int_{t_{0}}^{T}\int_{0}^{\infty}e^{\left(s-t_{0}\right)\lambda_{\varepsilon}}\lambda_{\varepsilon}^{p/2}dE_{\lambda_{\varepsilon}}F\left(u\right)\left(s\right)ds.

We also remark that in the same spirit of (2.1) the nonlinear ill-posed operator 𝒯\mathcal{T} (see Subsection 1.3) can be written in the following closed form:

(2.3) gT=𝒯g0:=0eTλλp/2𝑑Eλg0+0T0e(sT)λλp/2𝑑EλF(u)(s)𝑑s,g_{T}=\mathcal{T}g_{0}:=\int_{0}^{\infty}e^{-T\lambda}\lambda^{p/2}dE_{\lambda}g_{0}+\int_{0}^{T}\int_{0}^{\infty}e^{\left(s-T\right)\lambda}\lambda^{p/2}dE_{\lambda}F\left(u\right)\left(s\right)ds,

in which we define that 𝒯:𝒟(𝒯)H#p(Ω)\mathcal{T}:\mathcal{D}\left(\mathcal{T}\right)\to H_{\#}^{p}\left(\Omega\right) with 𝒟(𝒯)=H#p(Ω)\mathcal{D}\left(\mathcal{T}\right)=H_{\#}^{p}\left(\Omega\right).

2.2. Verification of the cut-off approach: high-order Sobolev-Gevrey smoothness

From now on, we aim to verify a modified version of the spectral cut-off regularization 𝒫ε\mathcal{P}_{\varepsilon} in a high-order smoothness setting. As delved into the nonlinear case in [15], if we impose the Gevrey smoothness on u(,t)u\left(\cdot,t\right) for all t[0,T]t\in\left[0,T\right], the verification of this approach is straightforward. We thereupon obtain the convergence for t0t\geq 0; compared to the standard cut-off method which yields the convergence only for t>0t>0 in the nonlinear case. This regularity assumption is very strong and it does not seem applicable as discussed in the analysis of the forward model (cf. Theorem 2), saying that the Gevrey smoothness can only be available for t>0t>0. It turns out that we need some modification of this cut-off projection in this paper. On the whole, we use the following scheme:

  • Since u(,t)u\left(\cdot,t\right) satisfies the Gevrey smoothness Gtp/2(Ω)G_{t}^{p/2}\left(\Omega\right) for t>0t>0, we keep solving the backward problem 𝒯u(t0)=gTε\mathcal{T}u\left(t_{0}\right)=g_{T}^{\varepsilon} for t0>0t_{0}>0 by the projection 𝒫ε\mathcal{P}_{\varepsilon} constructed from (2.1).

  • Assume that ut(,t)Hp(Ω)u_{t}\left(\cdot,t\right)\in H^{p}\left(\Omega\right). We find tε(0,T)t_{\varepsilon}\in\left(0,T\right) such that u(tε)u\left(t_{\varepsilon}\right) is an approximation of g0g_{0} in Hp(Ω)H^{p}\left(\Omega\right).

The central point of this modification is that this time we use the Sobolev smoothness on utu_{t} in the neighborhood of t=0t=0 to avoid the Gevrey smoothness on g0g_{0}. This assumption is consistent with the fact that utL2(0,T;L2(Ω))u_{t}\in L^{2}\left(0,T;L^{2}\left(\Omega\right)\right) obtained in Theorem 2 and thus attainable by the inclusion C(0,T;Hp(Ω))L2(0,T;L2(Ω))C\left(0,T;H^{p}\left(\Omega\right)\right)\subset L^{2}\left(0,T;L^{2}\left(\Omega\right)\right). Note that the Sobolev smoothness imposed on both g0g_{0} and u(,t)u\left(\cdot,t\right) in Hp(Ω)H^{p}\left(\Omega\right) with p>d/2p>d/2 is essential for the presence of the nonlinearity FF.

2.2.1. Case 1: L(M)=CL\left(M\right)=C independent of MM

Theorem 3.

Let p>d/2p>d/2. Assume that (1.3), (1.4), (1.5), (1.6), (1.7) hold. Then for t(0,T)t\in\left(0,T\right) the variational source condition (1.9) holds true for the operator (2.3) with

(2.4) ψt(δ):=C(δep+2(tT)L2u(,t)Gtp/22)tT+t+p2.\psi_{t}\left(\delta\right):=C\left(\frac{\delta}{e^{p+2\left(t-T\right)L^{2}}\left\|u\left(\cdot,t\right)\right\|_{G_{t}^{p/2}}^{2}}\right)^{\frac{t}{T+t+\frac{p}{2}}}.

Consequently, choosing the cut-off parameter

Cε=CT+t+p2logep2+(tT)L2u(,t)Gtp/2ε1/2,C_{\varepsilon}=\frac{C}{T+t+\frac{p}{2}}\log\frac{e^{\frac{p}{2}+\left(t-T\right)L^{2}}\left\|u\left(\cdot,t\right)\right\|_{G_{t}^{p/2}}}{\varepsilon^{1/2}},

the orthogonal projection 𝒫ε(Hp(Ω))\mathcal{P}_{\varepsilon}\in\mathcal{L}\left(H^{p}\left(\Omega\right)\right) defined in (2.2) is convergent with the rate

(2.5) uε(,t)u(,t)Hp(Ω)CεtT+t+p2.\left\|u_{\varepsilon}\left(\cdot,t\right)-u\left(\cdot,t\right)\right\|_{H^{p}\left(\Omega\right)}\leq C\varepsilon^{\frac{t}{T+t+\frac{p}{2}}}.
Proof.

Here we provide proof of the solution smoothness (C1)\left(\text{C}_{1}\right) (cf. Subsection 1.3) whenever u(,t)Gtp/2u\left(\cdot,t\right)\in G_{t}^{p/2} for p>d/2p>d/2 and t(0,T)t\in\left(0,T\right). In this sense, we have

u(,t)𝒫εu(,t)Hp(Ω)2\displaystyle\left\|u\left(\cdot,t\right)-\mathcal{P}_{\varepsilon}u\left(\cdot,t\right)\right\|_{H^{p}\left(\Omega\right)}^{2} =λελεpdEλu(,t)2\displaystyle=\int_{\lambda_{\varepsilon}}^{\infty}\lambda_{\varepsilon}^{p}d\left\|E_{\lambda}u\left(\cdot,t\right)\right\|^{2}
e2Cεtλελεpe2λεtdEλu(,t)2e2Cεtu(,t)Gtp/2(Ω)2,\displaystyle\leq e^{-2C_{\varepsilon}t}\int_{\lambda_{\varepsilon}}^{\infty}\lambda_{\varepsilon}^{p}e^{2\lambda_{\varepsilon}t}d\left\|E_{\lambda}u\left(\cdot,t\right)\right\|^{2}\leq e^{-2C_{\varepsilon}t}\left\|u\left(\cdot,t\right)\right\|_{G_{t}^{p/2}\left(\Omega\right)}^{2},

which implies that κt(Cε)\kappa_{t}\left(C_{\varepsilon}\right) decays with the rate eCεte^{-C_{\varepsilon}t}.

Concerning the degree of ill-posedness (C2)\left(\text{C}_{2}\right), we employ the equivalent formulation of the variational source condition (1.9):

𝒫r(ff)𝕏σ(r)𝒯(f)𝒯(f)𝕏for any f𝒟(𝒯)𝕏.\left\|\mathcal{P}_{r}\left(f^{\dagger}-f\right)\right\|_{\mathbb{X}}\leq\sigma\left(r\right)\left\|\mathcal{T}\left(f^{\dagger}\right)-\mathcal{T}\left(f\right)\right\|_{\mathbb{X}}\;\text{for any }f\in\mathcal{D}\left(\mathcal{T}\right)\subset\mathbb{X}.

Notice from (1.3) that the data measurement is not as smooth as the regularity assumption we imposed. Henceforward, for any given u¯C([0,T];Hp(Ω))\{0}\bar{u}\in C\left(\left[0,T\right];H^{p}\left(\Omega\right)\right)\backslash\left\{0\right\} it follows from (2.1) that

(2.6) 𝒫ε(u(,t)u¯(,t))Hp(Ω)2\displaystyle\left\|\mathcal{P}_{\varepsilon}\left(u\left(\cdot,t\right)-\bar{u}\left(\cdot,t\right)\right)\right\|_{H^{p}\left(\Omega\right)}^{2} 2Cεp0e2(Tt)CεdEλε(gTεg¯Tε)2\displaystyle\leq 2C_{\varepsilon}^{p}\int_{0}^{\infty}e^{2\left(T-t\right)C_{\varepsilon}}d\left\|E_{\lambda_{\varepsilon}}\left(g_{T}^{\varepsilon}-\bar{g}_{T}^{\varepsilon}\right)\right\|^{2}
+2tT0e2(st)CελεpdEλεF(uu¯)(s)2𝑑s,\displaystyle+2\int_{t}^{T}\int_{0}^{\infty}e^{2\left(s-t\right)C_{\varepsilon}}\lambda_{\varepsilon}^{p}d\left\|E_{\lambda_{\varepsilon}}F\left(u-\bar{u}\right)\left(s\right)\right\|^{2}ds,

where we have denoted by g¯Tε\bar{g}_{T}^{\varepsilon} the corresponding final data of u¯\bar{u}.

In (2.6), applying the globally Lipschitz continuity of FF and then multiplying the resulting estimate by the exponential weight e2tCεe^{2tC_{\varepsilon}} we find that

e2tCε𝒫ε(u(,t)u¯(,t))Hp(Ω)2\displaystyle e^{2tC_{\varepsilon}}\left\|\mathcal{P}_{\varepsilon}\left(u\left(\cdot,t\right)-\bar{u}\left(\cdot,t\right)\right)\right\|_{H^{p}\left(\Omega\right)}^{2} 2Cεpe2TCεgTεg¯TεL2(Ω)2\displaystyle\leq 2C_{\varepsilon}^{p}e^{2TC_{\varepsilon}}\left\|g_{T}^{\varepsilon}-\bar{g}_{T}^{\varepsilon}\right\|_{L^{2}\left(\Omega\right)}^{2}
+2L2tTe2sCε𝒫ε(u(,s)u¯(,s))Hp(Ω)2𝑑s.\displaystyle+2L^{2}\int_{t}^{T}e^{2sC_{\varepsilon}}\left\|\mathcal{P}_{\varepsilon}\left(u\left(\cdot,s\right)-\bar{u}\left(\cdot,s\right)\right)\right\|_{H^{p}\left(\Omega\right)}^{2}ds.

With the aid of the Gronwall inequality, we get

𝒫ε(u(,t)u¯(,t))Hp(Ω)22Cεpe2(Tt)(Cε+L2)gTεg¯TεL2(Ω)2,\left\|\mathcal{P}_{\varepsilon}\left(u\left(\cdot,t\right)-\bar{u}\left(\cdot,t\right)\right)\right\|_{H^{p}\left(\Omega\right)}^{2}\leq 2C_{\varepsilon}^{p}e^{2\left(T-t\right)\left(C_{\varepsilon}+L^{2}\right)}\left\|g_{T}^{\varepsilon}-\bar{g}_{T}^{\varepsilon}\right\|_{L^{2}\left(\Omega\right)}^{2},

yielding c=0c=0 and that σt(Cε)\sigma_{t}\left(C_{\varepsilon}\right) increases with the rate Cεpe(Tt)(Cε+L2)C_{\varepsilon}^{p}e^{\left(T-t\right)\left(C_{\varepsilon}+L^{2}\right)} in (C2)\left(\text{C}_{2}\right).

Therefore, we conclude that the solution u(,t)u\left(\cdot,t\right) satisfies the variational source condition (1.9) with

(2.7) ψt(δ):=2infCε>0[u(,t)Gtp/22e2Cεt+Cεp2e(Tt)(Cε+L2)δ],\psi_{t}\left(\delta\right):=2\inf_{C_{\varepsilon}>0}\left[\left\|u\left(\cdot,t\right)\right\|_{G_{t}^{p/2}}^{2}e^{-2C_{\varepsilon}t}+C_{\varepsilon}^{\frac{p}{2}}e^{\left(T-t\right)\left(C_{\varepsilon}+L^{2}\right)}\sqrt{\delta}\right],

then these two terms in the infimum are equal for δ=Cεpe2(T+t)Cεe2(tT)L2u(,t)Gtp/22\delta=C_{\varepsilon}^{-p}e^{-2\left(T+t\right)C_{\varepsilon}}e^{2\left(t-T\right)L^{2}}\left\|u\left(\cdot,t\right)\right\|_{G_{t}^{p/2}}^{2}. With this, we take the logarithm on both sides and use the elementary inequality that log(a)>1a1\log\left(a\right)>1-a^{-1} for any a>0a>0 to rule out the choice of CεC_{\varepsilon}. Thus, we can find some C>0C>0 independent of δ\delta such that

Cε=CT+t+p2logep2+(tT)L2u(,t)Gtp/2δ1/2,C_{\varepsilon}=\frac{C}{T+t+\frac{p}{2}}\log\frac{e^{\frac{p}{2}+\left(t-T\right)L^{2}}\left\|u\left(\cdot,t\right)\right\|_{G_{t}^{p/2}}}{\delta^{1/2}},

and plugging this into (2.7), we obtain the index function (2.4). On top of that, cf. [4] the error bound (2.5) itself holds for t(0,T)t\in\left(0,T\right). ∎

As a byproduct of Theorem 3 and in accordance with the argument in [4], the minimizers of the Tikhonov functional (1.8) satisfy the convergence rate (2.5) if α\alpha can be chosen such that 12α(ψ)(4ε2)-\frac{1}{2\alpha}\in\partial\left(-\psi\right)\left(4\varepsilon^{2}\right), where (ψ)\partial\left(-\psi\right) denotes the subdifferential of ψ-\psi. By this means, we compute that α=α(ε)=Cε2(T+p2)T+t+p2\alpha=\alpha\left(\varepsilon\right)=C\varepsilon^{\frac{2\left(T+\frac{p}{2}\right)}{T+t+\frac{p}{2}}}, indicating that α0+\alpha\searrow 0^{+} as the measurement error tends to 0. In addition, we obtain the following conditional stability estimate.

Corollary 4.

For t(0,T)t\in\left(0,T\right), let u~(,t)\tilde{u}\left(\cdot,t\right) and u¯(,t)\bar{u}\left(\cdot,t\right) be two solutions obtained from the integral equation (2.1). Under the assumptions of Theorem 3, the following estimate holds

u~(,t)u¯(,t)Hp(Ω)C(g~Tg¯TL2(Ω)ep+2(tT)L2u(,t)Gtp/22)tT+t+p2,\left\|\tilde{u}\left(\cdot,t\right)-\bar{u}\left(\cdot,t\right)\right\|_{H^{p}\left(\Omega\right)}\leq C\left(\frac{\left\|\tilde{g}_{T}-\bar{g}_{T}\right\|_{L^{2}\left(\Omega\right)}}{e^{p+2\left(t-T\right)L^{2}}\left\|u\left(\cdot,t\right)\right\|_{G_{t}^{p/2}}^{2}}\right)^{\frac{t}{T+t+\frac{p}{2}}},

where g~T\tilde{g}_{T} and g¯T\bar{g}_{T} are final conditions corresponding to u~(,t)\tilde{u}\left(\cdot,t\right) and u¯(,t)\bar{u}\left(\cdot,t\right), respectively.

Our modified scheme now brings into play its own feature: pointing out an approximation candidate of our solution g0g_{0}. In fact, it is clear to see that (2.5) is not convergent when t=0t=0. Our scheme to approximate the initial density g0g_{0} is very simple as we merely need to compute the “data” u(,tε)u\left(\cdot,t_{\varepsilon}\right) that have been solved through the standard spectral scheme 𝒫ε\mathcal{P}_{\varepsilon}. In the following, we not only show the existence of such tεt_{\varepsilon}, but also obtain a rigorous ε\varepsilon-dependent admissible set that contains it, directly proving the fact that tε0+t_{\varepsilon}\searrow 0^{+} as ε0+\varepsilon\searrow 0^{+}. Mathematically, one can establish an orthogonal projection and then mimic the way we gain Theorem 3 to deduce the rate of convergence. Since, again, this step is very trivial and somewhat self-contained by the help of the triangle inequality, our proof below follows the conventional way. Thus, the expression of the index function is clear through the derivation of the convergence rate.

Theorem 5.

Let p>d/2p>d/2. Suppose that ut(,t)Hp(Ω)u_{t}\left(\cdot,t\right)\in H^{p}\left(\Omega\right) for t(0,T)t\in\left(0,T\right). Let uε(,t)u_{\varepsilon}\left(\cdot,t\right) for t(0,T)t\in\left(0,T\right) be the unique solution obtained from the projection 𝒫ε(Hp(Ω))\mathcal{P}_{\varepsilon}\in\mathcal{L}\left(H^{p}\left(\Omega\right)\right) in (2.2). Then there always exists a sufficiently small ε\varepsilon-dependent time tε>0t_{\varepsilon}>0 such that

(2.8) uε(tε)u(0)Hp(Ω)C(T+p2)(T+p2+1)2+4(T+p2)log(ε1)+T+p21.\left\|u_{\varepsilon}\left(t_{\varepsilon}\right)-u\left(0\right)\right\|_{H^{p}\left(\Omega\right)}\leq\frac{C\left(T+\frac{p}{2}\right)}{\sqrt{\left(T+\frac{p}{2}+1\right)^{2}+4\left(T+\frac{p}{2}\right)\log\left(\varepsilon^{-1}\right)}+T+\frac{p}{2}-1}.
Proof.

Using (2.5) and the triangle inequality, for any tε(0,T)t_{\varepsilon}\in\left(0,T\right), we find the following estimate :

(2.9) uε(tε)u(0)Hp(Ω)\displaystyle\left\|u_{\varepsilon}\left(t_{\varepsilon}\right)-u\left(0\right)\right\|_{H^{p}\left(\Omega\right)} uε(tε)u(tε)Hp(Ω)+u(tε)u(0)Hp(Ω)\displaystyle\leq\left\|u_{\varepsilon}\left(t_{\varepsilon}\right)-u\left(t_{\varepsilon}\right)\right\|_{H^{p}\left(\Omega\right)}+\left\|u\left(t_{\varepsilon}\right)-u\left(0\right)\right\|_{H^{p}\left(\Omega\right)}
CεtεT+tε+p2+tεutHp(Ω).\displaystyle\leq C\varepsilon^{\frac{t_{\varepsilon}}{T+t_{\varepsilon}+\frac{p}{2}}}+t_{\varepsilon}\left\|u_{t}\right\|_{H^{p}\left(\Omega\right)}.

This upper bound can be done if we can find the infimum 12inftε>0(εtεT+tε+p2+tε)\frac{1}{2}\inf_{t_{\varepsilon}>0}\left(\varepsilon^{\frac{t_{\varepsilon}}{T+t_{\varepsilon}+\frac{p}{2}}}+t_{\varepsilon}\right) for some tε(0,T)t_{\varepsilon}\in\left(0,T\right). This indicates that we need to solve the following algebraic problem:

εtεT+tε+p2=tε,\varepsilon^{\frac{t_{\varepsilon}}{T+t_{\varepsilon}+\frac{p}{2}}}=t_{\varepsilon},

expecting that tε>0t_{\varepsilon}>0 is sufficiently small. Taking the logarithm on both sides of this equation and using the standard inequality log(a)>1a1\log\left(a\right)>1-a^{-1} for any a>0a>0, we have the following inequality:

tε2(logε1)(T+p21)tε+T+p2>0.t_{\varepsilon}^{2}\left(\log\varepsilon-1\right)-\left(T+\frac{p}{2}-1\right)t_{\varepsilon}+T+\frac{p}{2}>0.

Due to the fact that |T+p21|2+4(T+p2)(1logε)>0\left|T+\frac{p}{2}-1\right|^{2}+4\left(T+\frac{p}{2}\right)\left(1-\log\varepsilon\right)>0 and logε<1\log\varepsilon<1, we deduce that

tε(bb2+4(b+1)(1logε)2(logε1),b+b2+4(b+1)(1logε)2(logε1)),t_{\varepsilon}\in\left(\frac{-b-\sqrt{b^{2}+4\left(b+1\right)\left(1-\log\varepsilon\right)}}{2\left(\log\varepsilon-1\right)},\frac{-b+\sqrt{b^{2}+4\left(b+1\right)\left(1-\log\varepsilon\right)}}{2\left(\log\varepsilon-1\right)}\right),

where we have denoted by b:=T+p21b:=T+\frac{p}{2}-1. Notice that taking ε0+\varepsilon\searrow 0^{+}, the rationalizing technique gives the following limit:

limε0+(b+b2+4(b+1)(1logε)2(logε1))\displaystyle\lim_{\varepsilon\to 0^{+}}\left(-\frac{b+\sqrt{b^{2}+4\left(b+1\right)\left(1-\log\varepsilon\right)}}{2\left(\log\varepsilon-1\right)}\right)
=limε0+b2b24(b+1)(1+log(ε1))2(1+log(ε1))(bb2+4(b+1)(1+log(ε1)))\displaystyle=\lim_{\varepsilon\to 0^{+}}\frac{b^{2}-b^{2}-4\left(b+1\right)\left(1+\log\left(\varepsilon^{-1}\right)\right)}{2\left(1+\log\left(\varepsilon^{-1}\right)\right)\left(b-\sqrt{b^{2}+4\left(b+1\right)\left(1+\log\left(\varepsilon^{-1}\right)\right)}\right)}
=limε0+2(b+1)b2+4(b+1)(1+log(ε1))b=0,\displaystyle=\lim_{\varepsilon\to 0^{+}}\frac{2\left(b+1\right)}{\sqrt{b^{2}+4\left(b+1\right)\left(1+\log\left(\varepsilon^{-1}\right)\right)}-b}=0,

and similarly,

limε0+(bb2+4(b+1)log(ε1)2(logε1))\displaystyle\lim_{\varepsilon\to 0^{+}}\left(-\frac{b-\sqrt{b^{2}+4\left(b+1\right)\log\left(\varepsilon^{-1}\right)}}{2\left(\log\varepsilon-1\right)}\right) =limε0+2(b+1)b2+4(b+1)(1+log(ε1))+b=0.\displaystyle=\lim_{\varepsilon\to 0^{+}}\frac{2\left(b+1\right)}{\sqrt{b^{2}+4\left(b+1\right)\left(1+\log\left(\varepsilon^{-1}\right)\right)}+b}=0.

Hence, it follows from (2.9) that the upper bound we gain would be of the form in (2.8). ∎

2.2.2. Case 2: L(M)L\left(M\right) dependent of MM

When L(M)L\left(M\right) depends on MM, our nonlinear spectral regularization 𝒫ε\mathcal{P}_{\varepsilon} may be no longer convergent as the boundedness of the regularized solution is not well-controlled; cf. (2.4). This means that the quantity MM now has to be dependent of ε\varepsilon, saying that M=MεM=M_{\varepsilon}, when getting involved in the scheme 𝒫ε\mathcal{P}_{\varepsilon}. This ε\varepsilon dependence impacts on the structure of the index function and further on the whole rate of convergence in Theorem 3.

From now on, for some constant >0\ell>0 and for ww\in\mathbb{R} we introduce the cut-off function of FF, denoted by FF_{\ell}, as follows:

(2.10) F(w):={F()if w,F(w)if |w|,F()if w.F_{\ell}\left(w\right):=\begin{cases}F\left(\ell\right)&\text{if }w\geq\ell,\\ F\left(w\right)&\text{if }\left|w\right|\leq\ell,\\ F\left(-\ell\right)&\text{if }w\leq-\ell.\end{cases}

Following this way, we modify the regularization scheme (2.2) by

(2.11) 𝒫¯εu(,t0)=0eTλελεp/2𝑑EλεgTεt0T0esλελεp/2𝑑EλεFMε(u)(s)𝑑s.\bar{\mathcal{P}}_{\varepsilon}u\left(\cdot,t_{0}\right)=\int_{0}^{\infty}e^{T\lambda_{\varepsilon}}\lambda_{\varepsilon}^{p/2}dE_{\lambda_{\varepsilon}}g_{T}^{\varepsilon}-\int_{t_{0}}^{T}\int_{0}^{\infty}e^{s\lambda_{\varepsilon}}\lambda_{\varepsilon}^{p/2}dE_{\lambda_{\varepsilon}}F_{M_{\varepsilon}}\left(u\right)\left(s\right)ds.

and therefore, it enables us to derive the convergence rate under a suitable choice of MεM_{\varepsilon}. Note that our cut-off function FMεF_{M_{\varepsilon}} (2.10) possesses the similar property in (1.6), i.e.

|FMε(u)FMε(v)|L(Mε)|uv|.\left|F_{M_{\varepsilon}}\left(u\right)-F_{M_{\varepsilon}}\left(v\right)\right|\leq L\left(M_{\varepsilon}\right)\left|u-v\right|.
Theorem 6.

Under the assumptions of Theorem 3, we choose Mε>0M_{\varepsilon}>0 such that for t(0,T)t\in\left(0,T\right)

(2.12) L2(Mε)12(tT)log(εβ)for β(0,12).L^{2}\left(M_{\varepsilon}\right)\leq\frac{1}{2\left(t-T\right)}\log\left(\varepsilon^{\beta}\right)\quad\text{for }\beta\in\left(0,\frac{1}{2}\right).

Then the variational source condition (1.9) holds true for the operator (2.3) with

(2.13) ψt(δ):=C(δ12βepu(,t)Gtp/22)tT+t+p2.\psi_{t}\left(\delta\right):=C\left(\frac{\delta^{1-2\beta}}{e^{p}\left\|u\left(\cdot,t\right)\right\|_{G_{t}^{p/2}}^{2}}\right)^{\frac{t}{T+t+\frac{p}{2}}}.

Consequently, choosing the cut-off parameter

Cε=CT+t+p2logep2u(,t)Gtp/2ε1/2β,C_{\varepsilon}=\frac{C}{T+t+\frac{p}{2}}\log\frac{e^{\frac{p}{2}}\left\|u\left(\cdot,t\right)\right\|_{G_{t}^{p/2}}}{\varepsilon^{1/2-\beta}},

the orthogonal projection 𝒫¯ε(Hp(Ω))\mathcal{\bar{P}}_{\varepsilon}\in\mathcal{L}\left(H^{p}\left(\Omega\right)\right) defined in (2.11) is convergent with the rate

(2.14) uε(,t)u(,t)Hp(Ω)Cεt(12β)T+t+p2.\left\|u_{\varepsilon}\left(\cdot,t\right)-u\left(\cdot,t\right)\right\|_{H^{p}\left(\Omega\right)}\leq C\varepsilon^{\frac{t\left(1-2\beta\right)}{T+t+\frac{p}{2}}}.
Proof.

Proof of this theorem is straightforward. In fact, starting from the argument (2.7) we know in this case that δ=Cεpe2(T+t)Cεe2(tT)L2(Mε)u(,t)Gtp/22\delta=C_{\varepsilon}^{-p}e^{-2\left(T+t\right)C_{\varepsilon}}e^{2\left(t-T\right)L^{2}\left(M_{\varepsilon}\right)}\left\|u\left(\cdot,t\right)\right\|_{G_{t}^{p/2}}^{2}. Thus, the choice of MεM_{\varepsilon} in (2.12) is apparent. This then results in the form (2.13) of the index function. Hence, the proof is complete. ∎

Remark 7.

Based on the result obtained in Theorem 6, we remark the following:

  • As is well-known in continuous population models for single species, we suppose the nonlinearity in the form of

    F(u)=u(1u)u21+u2,F\left(u\right)=u\left(1-u\right)-\frac{u^{2}}{1+u^{2}},

    as a prominent example of the growth-and-predation rate for the spruce budworm which critically defoliated the balsam fir in Canada (cf. [14]). In this circumstance, it is easy to get L(M)=1+4M>0L\left(M\right)=1+4M>0 and hence, we can choose that

    Mε14(12(tT)log(εβ)1),M_{\varepsilon}\leq\frac{1}{4}\left(\sqrt{\frac{1}{2\left(t-T\right)}\log\left(\varepsilon^{\beta}\right)}-1\right),

    working with the concrete assumption ε<e2(tT)β\varepsilon<e^{\frac{2\left(t-T\right)}{\beta}}.

  • This result can be further extended to multiple-species cases. Indeed, we consider a time evolution of the concentration of LL\in\mathbb{N} chemical components or constituents (molecules, radicals, ions) in a reaction network (cf. e.g. [1]) which reads as

    l=1Lα(l,r)Xll=1Lη(l,r)Xlfor r=1,R¯,\sum_{l=1}^{L}\alpha\left(l,r\right)X_{l}\to\sum_{l=1}^{L}\eta\left(l,r\right)X_{l}\quad\text{for }r=\overline{1,R},

    where α(l,r)0\alpha\left(l,r\right)\geq 0 and η(l,r)0\eta\left(l,r\right)\geq 0 are the stoichiometric coefficients or molecularities. In a constructive manner, the mass action kinetic deterministic model of this reaction is governed by the following system of PDEs:

    tcl=Δcl+r=1R(η(l,r)α(l,r))l=1Lclα(l,r)for l=1,L¯,\partial_{t}c_{l}=\Delta c_{l}+\sum_{r=1}^{R}\left(\eta\left(l,r\right)-\alpha\left(l,r\right)\right)\prod_{l=1}^{L}c_{l}^{\alpha\left(l,r\right)}\quad\text{for }l=\overline{1,L},

    where clc_{l} is viewed as the concentration of the llth component at time t[0,T]t\in\left[0,T\right]. In this regard, we obtain the vector-valued reaction-diffusion equation (1.1) in the form of 𝐮t+𝒜𝐮=(𝐮)\mathbf{u}_{t}+\mathcal{A}\mathbf{u}=\mathcal{F}\left(\mathbf{u}\right) by denoting the vector of concentrations 𝐮(t)=(c1(t),,cL(t))L\mathbf{u}\left(t\right)=\left(c_{1}\left(t\right),\ldots,c_{L}\left(t\right)\right)\in\mathbb{R}^{L} and

    𝐮α=(𝐮α(,1),,𝐮α(,R))for 𝐮α(,r)=l=1Lclα(l,r),\mathbf{u}^{\alpha}=\left(\mathbf{u}^{\alpha\left(\cdot,1\right)},\ldots,\mathbf{u}^{\alpha\left(\cdot,R\right)}\right)\quad\text{for }\mathbf{u}^{\alpha\left(\cdot,r\right)}=\prod_{l=1}^{L}c_{l}^{\alpha\left(l,r\right)},
    diag(𝐳)=|z1000z20zL1000zL|for 𝐳L,\text{diag}\left(\mathbf{z}\right)=\begin{vmatrix}z_{1}&0&\cdots&\cdots&0\\ 0&z_{2}&\ddots&&\vdots\\ \vdots&\ddots&\ddots&\ddots&\vdots\\ \vdots&&0&z_{L-1}&0\\ 0&\cdots&\cdots&0&z_{L}\end{vmatrix}\quad\text{for }\mathbf{z}\in\mathbb{R}^{L},

    with (𝐮)=(ηα)diag(𝐮α)\mathcal{F}\left(\mathbf{u}\right)=\left(\eta-\alpha\right)\text{diag}\left(\mathbf{u}^{\alpha}\right) where η\eta and α\alpha are the vectors of the stoichiometric coefficients. By this way Theorem 6 can be applied.

Last but not least, we state the convergence rate at t=0t=0 by combining the results of Theorems 5 and 6.

Theorem 8.

Under the assumptions of Theorem 5, let uε(,t)u_{\varepsilon}\left(\cdot,t\right) for t(0,T)t\in\left(0,T\right) be the unique solution obtained from the projection 𝒫¯ε(Hp(Ω))\bar{\mathcal{P}}_{\varepsilon}\in\mathcal{L}\left(H^{p}\left(\Omega\right)\right) in (2.11). Then there always exists a sufficiently small ε\varepsilon-dependent time tε>0t_{\varepsilon}>0 such that

(2.15) uε(tε)u(0)Hp(Ω)C(T+p2)(T+p2+1)2+4(T+p2)(12β)log(ε1)+T+p21.\left\|u_{\varepsilon}\left(t_{\varepsilon}\right)-u\left(0\right)\right\|_{H^{p}\left(\Omega\right)}\leq\frac{C\left(T+\frac{p}{2}\right)}{\sqrt{\left(T+\frac{p}{2}+1\right)^{2}+4\left(T+\frac{p}{2}\right)\left(1-2\beta\right)\log\left(\varepsilon^{-1}\right)}+T+\frac{p}{2}-1}.
Proof.

In the same manner as in proof of Theorem 5, we prove the target estimate (2.15) by seeking the infimum 12inftε>0(εtε(12β)T+tε+p2+tε)\frac{1}{2}\inf_{t_{\varepsilon}>0}\left(\varepsilon^{\frac{t_{\varepsilon}\left(1-2\beta\right)}{T+t_{\varepsilon}+\frac{p}{2}}}+t_{\varepsilon}\right). Solving the algebraic problem

εtε(12β)T+tε+p2=tε,\varepsilon^{\frac{t_{\varepsilon}\left(1-2\beta\right)}{T+t_{\varepsilon}+\frac{p}{2}}}=t_{\varepsilon},

we are led to the following inequality

tε2((12β)logε1)(T+p21)tε+T+p2>0.t_{\varepsilon}^{2}\left(\left(1-2\beta\right)\log\varepsilon-1\right)-\left(T+\frac{p}{2}-1\right)t_{\varepsilon}+T+\frac{p}{2}>0.

Hereby, we find the admissible interval for tεt_{\varepsilon}, as follows:

tε(bb2+4(b+1)(1logε)2((12β)logε1),b+b2+4(b+1)(1logε)2((12β)logε1)),t_{\varepsilon}\in\left(\frac{-b-\sqrt{b^{2}+4\left(b+1\right)\left(1-\log\varepsilon\right)}}{2\left(\left(1-2\beta\right)\log\varepsilon-1\right)},\frac{-b+\sqrt{b^{2}+4\left(b+1\right)\left(1-\log\varepsilon\right)}}{2\left(\left(1-2\beta\right)\log\varepsilon-1\right)}\right),

where b=T+p21b=T+\frac{p}{2}-1 is recalled.

Therefore, it is evident to obtain the rate (2.15). We complete the proof of the theorem. ∎

3. Convergence of an iterative scheme

In this section, we reduce ourselves to the case that the nonlinearity FF does not degenerate, i.e. F(0)=0F\left(0\right)=0 and there exist positive constants L0L_{0} and L1L_{1} such that

(3.1) 0<L0sup|w|MFw(w)L1.0<L_{0}\leq\sup_{\left|w\right|\leq M}\frac{\partial F}{\partial w}\left(w\right)\leq L_{1}.

Note that we now focus on solving the regularized solution in the open set (0,T)\left(0,T\right) since at t=0t=0 we only need to compute the approximation at t=tεt=t_{\varepsilon}. Let 1N1\leq N\in\mathbb{N} and take ω=T/N\omega=T/N. In this regard, we put

tn=Tnωfor n=1,N¯.t_{n}=T-n\omega\quad\text{for }n=\overline{1,N}.

This setting allows us to seek a numerical solution uεr,n(x)uε(x,tn)u_{\varepsilon}^{r,n}\left(x\right)\approx u_{\varepsilon}\left(x,t_{n}\right) for rr\in\mathbb{N} in the equivalent mesh-width in tt. The function uε(x,tn)u_{\varepsilon}\left(x,t_{n}\right) is the semi-discrete solution of the nonlinear scheme (2.11) under scrutiny in the previous part. Starting from the projection 𝒫¯ε\bar{\mathcal{P}}_{\varepsilon} with the cut-off function FMεF_{M_{\varepsilon}} used in (2.11), the iterative scheme is designed by

(3.2) (K+1)uεr+1,n=Kuεr,n+0e(Ttn)λελεp/2𝑑EλεgTε0γε(tn)λεp/2𝑑EλεFMε(uεr,n),\left(K+1\right)u_{\varepsilon}^{r+1,n}=Ku_{\varepsilon}^{r,n}+\int_{0}^{\infty}e^{\left(T-t_{n}\right)\lambda_{\varepsilon}}\lambda_{\varepsilon}^{p/2}dE_{\lambda_{\varepsilon}}g_{T}^{\varepsilon}-\int_{0}^{\infty}\gamma_{\varepsilon}\left(t_{n}\right)\lambda_{\varepsilon}^{p/2}dE_{\lambda_{\varepsilon}}F_{M_{\varepsilon}}\left(u_{\varepsilon}^{r,n}\right),

with the initial guesses uεr,0gTεu_{\varepsilon}^{r,0}\equiv g_{T}^{\varepsilon} and uε0,n0u_{\varepsilon}^{0,n}\equiv 0. In addition, we have denoted by γε(t):=tTe(st)λεdsCε1(e(Tt)Cε1)=:γ¯ε(t)\gamma_{\varepsilon}\left(t\right):=\int_{t}^{T}e^{\left(s-t\right)\lambda_{\varepsilon}}ds\leq C_{\varepsilon}^{-1}\left(e^{\left(T-t\right)C_{\varepsilon}}-1\right)=:\bar{\gamma}_{\varepsilon}\left(t\right) for t(0,T)t\in\left(0,T\right).

The presence of the so-called stabilization constant K>0K>0 is to guarantee the convergence of the scheme, somehow hindered by the Lipschitz nonlinearity FF. In fact, we wish to designate an unconditional numerical scheme for the projection 𝒫¯ε\bar{\mathcal{P}}_{\varepsilon} in the sense that the number of discretizations NN becomes free-to-choose by a suitable choice of KK. It is worth mentioning that the sequence {uεr,n}r\left\{u_{\varepsilon}^{r,n}\right\}_{r\in\mathbb{N}} is well-defined in the Sobolev space Hp(Ω)H^{p}\left(\Omega\right) for each ε>0\varepsilon>0 and thus the existence and uniqueness are self-contained by virtue of the linearity of the scheme. Note that the function γε\gamma_{\varepsilon} is decreasing in the time argument because of γε(tn+1)γε(tn)\gamma_{\varepsilon}\left(t_{n+1}\right)\geq\gamma_{\varepsilon}\left(t_{n}\right) for any 0tn+1tnT0\leq t_{n+1}\leq t_{n}\leq T, whilst it is transparently increasing in the argument λε\lambda_{\varepsilon}.

Theorem 9.

Under the assumptions of Theorem 5, let {uεr,n}r\left\{u_{\varepsilon}^{r,n}\right\}_{r\in\mathbb{N}} be the solution of the scheme (3.2). Then by choosing

(3.3) K:=K(n,ε)=max{γ¯ε(tn)L1,eTCε}>0for n,K:=K\left(n,\varepsilon\right)=\max\left\{\bar{\gamma}_{\varepsilon}\left(t_{n}\right)L_{1},e^{TC_{\varepsilon}}\right\}>0\;\text{for }n\in\mathbb{N},

this sequence is uniformly bounded in Hp(Ω)H^{p}\left(\Omega\right).

Proof.

In view of the decomposition

uεr,n=Cελp/2𝑑Eλuεr,n+0λεp/2𝑑Eλεuεr,n,u_{\varepsilon}^{r,n}=\int_{C_{\varepsilon}}^{\infty}\lambda^{p/2}dE_{\lambda}u_{\varepsilon}^{r,n}+\int_{0}^{\infty}\lambda_{\varepsilon}^{p/2}dE_{\lambda_{\varepsilon}}u_{\varepsilon}^{r,n},

we can find the upper bound in HpH^{p}-norm of uεr+1,nu_{\varepsilon}^{r+1,n} as follows:

(K+1)uεr+1,nHp(Ω)\displaystyle\left(K+1\right)\left\|u_{\varepsilon}^{r+1,n}\right\|_{H^{p}\left(\Omega\right)} e(Ttn)CεgTεHp(Ω)+KCελp/2𝑑Eλuεr,n\displaystyle\leq e^{\left(T-t_{n}\right)C_{\varepsilon}}\left\|g_{T}^{\varepsilon}\right\|_{H^{p}\left(\Omega\right)}+K\int_{C_{\varepsilon}}^{\infty}\lambda^{p/2}dE_{\lambda}u_{\varepsilon}^{r,n}
+0λεp/2|hMε|(λε)𝑑Eλεuεr,n,\displaystyle+\int_{0}^{\infty}\lambda_{\varepsilon}^{p/2}\left|h_{M_{\varepsilon}}\right|\left(\lambda_{\varepsilon}\right)dE_{\lambda_{\varepsilon}}u_{\varepsilon}^{r,n},

where we have denoted by hMε[yj]:=Kyjγε(tj)FMε[yj]h_{M_{\varepsilon}}\left[y^{j}\right]:=Ky^{j}-\gamma_{\varepsilon}\left(t_{j}\right)F_{M_{\varepsilon}}\left[y^{j}\right] for y=(yj)0jNy=\left(y^{j}\right)_{0\leq j\leq N}.

At this stage, we remark that h[yj]=Kγε(tj)FMε(yj)h^{\prime}\left[y^{j}\right]=K-\gamma_{\varepsilon}\left(t_{j}\right)F_{M_{\varepsilon}}^{\prime}\left(y^{j}\right) and therefore, it holds |h|Kγε(tj)L0\left|h^{\prime}\right|\leq K-\gamma_{\varepsilon}\left(t_{j}\right)L_{0} for a.e. yjy^{j}\in\mathbb{R}. By the mean value theorem together with the fact that F(0)=0F\left(0\right)=0, we estimate that

(K+1)uεr+1,nHp(Ω)\displaystyle\left(K+1\right)\left\|u_{\varepsilon}^{r+1,n}\right\|_{H^{p}\left(\Omega\right)} e(Ttn)CεgTεHp(Ω)+KCελp/2𝑑Eλuεr,n\displaystyle\leq e^{\left(T-t_{n}\right)C_{\varepsilon}}\left\|g_{T}^{\varepsilon}\right\|_{H^{p}\left(\Omega\right)}+K\int_{C_{\varepsilon}}^{\infty}\lambda^{p/2}dE_{\lambda}u_{\varepsilon}^{r,n}
+0λεp/2(Kγε(tn)L0)𝑑Eλεuεr,n\displaystyle+\int_{0}^{\infty}\lambda_{\varepsilon}^{p/2}\left(K-\gamma_{\varepsilon}\left(t_{n}\right)L_{0}\right)dE_{\lambda_{\varepsilon}}u_{\varepsilon}^{r,n}
e(Ttn)CεgTεHp(Ω)+(Kγε(tn)L0)uεr,nHp(Ω)\displaystyle\leq e^{\left(T-t_{n}\right)C_{\varepsilon}}\left\|g_{T}^{\varepsilon}\right\|_{H^{p}\left(\Omega\right)}+\left(K-\gamma_{\varepsilon}\left(t_{n}\right)L_{0}\right)\left\|u_{\varepsilon}^{r,n}\right\|_{H^{p}\left(\Omega\right)}
+γ¯ε(tn)L0Cελp/2𝑑Eλuεr,n.\displaystyle+\bar{\gamma}_{\varepsilon}\left(t_{n}\right)L_{0}\int_{C_{\varepsilon}}^{\infty}\lambda^{p/2}dE_{\lambda}u_{\varepsilon}^{r,n}.

By the choice of CεC_{\varepsilon}, we gain

limε0+Cελp/2𝑑Eλuεr,n=limCεCελp/2𝑑Eλuεr,n=0.\lim_{\varepsilon\to 0^{+}}\int_{C_{\varepsilon}}^{\infty}\lambda^{p/2}dE_{\lambda}u_{\varepsilon}^{r,n}=\lim_{C_{\varepsilon}\to\infty}\int_{C_{\varepsilon}}^{\infty}\lambda^{p/2}dE_{\lambda}u_{\varepsilon}^{r,n}=0.

Thus, we now enjoy the choice of KK in (3.3) to rule out that there exists μ(0,1)\mu\in\left(0,1\right) independent of r,nr,n and ε\varepsilon such that

uεr+1,nHp(Ω)μ(gTεHp(Ω)+Cελp/2𝑑Eλuεr,n+uεr,nHp(Ω)).\left\|u_{\varepsilon}^{r+1,n}\right\|_{H^{p}\left(\Omega\right)}\leq\mu\left(\left\|g_{T}^{\varepsilon}\right\|_{H^{p}\left(\Omega\right)}+\int_{C_{\varepsilon}}^{\infty}\lambda^{p/2}dE_{\lambda}u_{\varepsilon}^{r,n}+\left\|u_{\varepsilon}^{r,n}\right\|_{H^{p}\left(\Omega\right)}\right).

By induction, we obtain

(3.4) uεr,nHp(Ω)Cj=1rμjgTεHp(Ω),\left\|u_{\varepsilon}^{r,n}\right\|_{H^{p}\left(\Omega\right)}\leq C\sum_{j=1}^{r}\mu^{j}\left\|g_{T}^{\varepsilon}\right\|_{H^{p}\left(\Omega\right)},

which enables us to state that the scheme (3.2) is bounded in Hp(Ω)H^{p}\left(\Omega\right) for any r,nr,n and ε\varepsilon. Hence, we complete the proof of the theorem. ∎

Using the Banach-Alaoglu theorem, the uniform bound deduced in the proof of Theorem 9 indicates that we can extract a further subsequence (which we relabel with the same indexes if necessary) such that uεr,nuεnu_{\varepsilon}^{r,n}\to u_{\varepsilon}^{n} weakly in Hp(Ω)H^{p}\left(\Omega\right) as rr\to\infty. Furthermore, thanks to the Banach-Saks theorem we know that this subsequence also admits another subsequence such that the so-called Cesàro mean is strongly convergent to uεnu_{\varepsilon}^{n} in Hp(Ω)H^{p}\left(\Omega\right). In this sense, we can write

(3.5) 1Rr=0Ruεr,nuεnHp(Ω)0as R.\left\|\frac{1}{R}\sum_{r=0}^{R}u_{\varepsilon}^{r,n}-u_{\varepsilon}^{n}\right\|_{H^{p}\left(\Omega\right)}\to 0\quad\text{as }R\to\infty.

Define wεR,n:=1Rr=0Ruεr,nw_{\varepsilon}^{R,n}:=\frac{1}{R}\sum_{r=0}^{R}u_{\varepsilon}^{r,n} for RR\in\mathbb{N}. Our next step is to find the rate of convergence of the sequence {wεR,n}R\left\{w_{\varepsilon}^{R,n}\right\}_{R\in\mathbb{N}} acquired by (3.5). By this way, we not only prove that uεr,nuεnu_{\varepsilon}^{r,n}\to u_{\varepsilon}^{n} strongly in Hp(Ω)H^{p}\left(\Omega\right), but also show that uεnu_{\varepsilon}^{n} is identically the semi-discrete solution of the nonlinear scheme (2.11).

Theorem 10.

Under the assumptions of Theorem 5, let uε(,t)u_{\varepsilon}\left(\cdot,t\right) for t(0,T)t\in\left(0,T\right) be the unique solution obtained from the projection 𝒫¯ε(Hp(Ω))\bar{\mathcal{P}}_{\varepsilon}\in\mathcal{L}\left(H^{p}\left(\Omega\right)\right) in (2.11). Let {uεr,n}r\left\{u_{\varepsilon}^{r,n}\right\}_{r\in\mathbb{N}} be the solution of the iterative scheme (3.2) and let {wεR,n}R\left\{w_{\varepsilon}^{R,n}\right\}_{R\in\mathbb{N}} be the Cesàro mean of uεr,nu_{\varepsilon}^{r,n} with respect to rr. Then for r,nr,n\in\mathbb{N} and ε>0\varepsilon>0 there exists μ¯(0,1)\bar{\mu}\in\left(0,1\right) independent of r,nr,n and ε\varepsilon such that the following error bound holds

uεr,nuε(,tn)Hp(Ω)Cμ¯r,\left\|u_{\varepsilon}^{r,n}-u_{\varepsilon}\left(\cdot,t_{n}\right)\right\|_{H^{p}\left(\Omega\right)}\leq C\bar{\mu}^{r},

for rr sufficiently large. Moreover, for RR\in\mathbb{N} it holds

wεR,nuε(,tn)Hp(Ω)C(μ¯R)R.\left\|w_{\varepsilon}^{R,n}-u_{\varepsilon}\left(\cdot,t_{n}\right)\right\|_{H^{p}\left(\Omega\right)}\leq C\left(\frac{\bar{\mu}}{R}\right)^{R}.
Proof.

Define vεR+1,n:=wεR+1,nwεR,nHp(Ω)v_{\varepsilon}^{R+1,n}:=w_{\varepsilon}^{R+1,n}-w_{\varepsilon}^{R,n}\in H^{p}\left(\Omega\right). To gain the convergence rate, we compute the difference equation:

(3.6) (R+1)vεR+1,n\displaystyle\left(R+1\right)v_{\varepsilon}^{R+1,n} =r=0R+1uεr,nr=0Ruεr,n+(R+1)(1R+11R)r=0Ruεr,n\displaystyle=\sum_{r=0}^{R+1}u_{\varepsilon}^{r,n}-\sum_{r=0}^{R}u_{\varepsilon}^{r,n}+\left(R+1\right)\left(\frac{1}{R+1}-\frac{1}{R}\right)\sum_{r=0}^{R}u_{\varepsilon}^{r,n}
=uεR+1,n1Rr=0Ruεr,n=1Rr=0R(uεR+1,nuεr,n).\displaystyle=u_{\varepsilon}^{R+1,n}-\frac{1}{R}\sum_{r=0}^{R}u_{\varepsilon}^{r,n}=\frac{1}{R}\sum_{r=0}^{R}\left(u_{\varepsilon}^{R+1,n}-u_{\varepsilon}^{r,n}\right).

From now onward, we define Uεr+1,n:=uεr+1,nuεr,nHp(Ω)U_{\varepsilon}^{r+1,n}:=u_{\varepsilon}^{r+1,n}-u_{\varepsilon}^{r,n}\in H^{p}\left(\Omega\right). Following the same way we have done in the proof of Theorem 9, the function Uεr+1,nU_{\varepsilon}^{r+1,n} is expressed as

(K+1)Uεr+1,n\displaystyle\left(K+1\right)U_{\varepsilon}^{r+1,n} =KUεr,n0γε(tn)λεp/2𝑑EλεFMε(uεr,n)\displaystyle=KU_{\varepsilon}^{r,n}-\int_{0}^{\infty}\gamma_{\varepsilon}\left(t_{n}\right)\lambda_{\varepsilon}^{p/2}dE_{\lambda_{\varepsilon}}F_{M_{\varepsilon}}\left(u_{\varepsilon}^{r,n}\right)
+0γε(tn)λεp/2𝑑EλεFMε(uεr1,n).\displaystyle+\int_{0}^{\infty}\gamma_{\varepsilon}\left(t_{n}\right)\lambda_{\varepsilon}^{p/2}dE_{\lambda_{\varepsilon}}F_{M_{\varepsilon}}\left(u_{\varepsilon}^{r-1,n}\right).

With the aid of the decomposition

Uεr,n=Cελp/2𝑑EλUεr,n+0λεp/2𝑑EλεUεr,n,U_{\varepsilon}^{r,n}=\int_{C_{\varepsilon}}^{\infty}\lambda^{p/2}dE_{\lambda}U_{\varepsilon}^{r,n}+\int_{0}^{\infty}\lambda_{\varepsilon}^{p/2}dE_{\lambda_{\varepsilon}}U_{\varepsilon}^{r,n},

the function Uεn+1U_{\varepsilon}^{n+1} can be bounded from above in the HpH^{p}-norm by

(K+1)Uεr+1,nHp(Ω)\displaystyle\left(K+1\right)\left\|U_{\varepsilon}^{r+1,n}\right\|_{H^{p}\left(\Omega\right)} KCελp/2𝑑EλUεr,n\displaystyle\leq K\int_{C_{\varepsilon}}^{\infty}\lambda^{p/2}dE_{\lambda}U_{\varepsilon}^{r,n}
+|0λεp/2hMε(λε)𝑑Eλεuεr,n0λεp/2hMε(λε)𝑑Eλεuεr1,n|,\displaystyle+\left|\int_{0}^{\infty}\lambda_{\varepsilon}^{p/2}h_{M_{\varepsilon}}\left(\lambda_{\varepsilon}\right)dE_{\lambda_{\varepsilon}}u_{\varepsilon}^{r,n}-\int_{0}^{\infty}\lambda_{\varepsilon}^{p/2}h_{M_{\varepsilon}}\left(\lambda_{\varepsilon}\right)dE_{\lambda_{\varepsilon}}u_{\varepsilon}^{r-1,n}\right|,

where we have recalled that hMε[yj]:=Kyjγε(tj)FMε[yj]h_{M_{\varepsilon}}\left[y^{j}\right]:=Ky^{j}-\gamma_{\varepsilon}\left(t_{j}\right)F_{M_{\varepsilon}}\left[y^{j}\right] for y=(yj)0jNy=\left(y^{j}\right)_{0\leq j\leq N}.

Henceforward, we arrive at

(K+1)Uεr+1,nHp(Ω)\displaystyle\left(K+1\right)\left\|U_{\varepsilon}^{r+1,n}\right\|_{H^{p}\left(\Omega\right)} (Kγε(tn)L0)Uεr,nHp(Ω)+γε(tn)L0Cελp/2𝑑EλUεr,n\displaystyle\leq\left(K-\gamma_{\varepsilon}\left(t_{n}\right)L_{0}\right)\left\|U_{\varepsilon}^{r,n}\right\|_{H^{p}\left(\Omega\right)}+\gamma_{\varepsilon}\left(t_{n}\right)L_{0}\int_{C_{\varepsilon}}^{\infty}\lambda^{p/2}dE_{\lambda}U_{\varepsilon}^{r,n}
KUεr,nHp(Ω),\displaystyle\leq K\left\|U_{\varepsilon}^{r,n}\right\|_{H^{p}\left(\Omega\right)},

by virtue of the fact already known that |h|Kγε(tj)L0\left|h^{\prime}\right|\leq K-\gamma_{\varepsilon}\left(t_{j}\right)L_{0} for a.e. yjy^{j}\in\mathbb{R} under the choice of KK in (3.3). Hereby, choosing μ¯=K(K+1)1(0,1)\bar{\mu}=K\left(K+1\right)^{-1}\in\left(0,1\right) independent of r,nr,n and ε\varepsilon we can conclude that

(3.7) Uεr+1,nHp(Ω)μ¯Uεr,nHp(Ω),\left\|U_{\varepsilon}^{r+1,n}\right\|_{H^{p}\left(\Omega\right)}\leq\bar{\mu}\left\|U_{\varepsilon}^{r,n}\right\|_{H^{p}\left(\Omega\right)},

which, by mathematical induction, leads to

Uεr,nHp(Ω)Cμ¯rgTεHp(Ω).\left\|U_{\varepsilon}^{r,n}\right\|_{H^{p}\left(\Omega\right)}\leq C\bar{\mu}^{r}\left\|g_{T}^{\varepsilon}\right\|_{H^{p}\left(\Omega\right)}.

Eventually, by the back-substitution of the function Uεr,nU_{\varepsilon}^{r,n} this means that

(3.8) uεr+l,nuεr,nHp(Ω)\displaystyle\left\|u_{\varepsilon}^{r+l,n}-u_{\varepsilon}^{r,n}\right\|_{H^{p}\left(\Omega\right)} uεr+l,nuεr+l1,nHp(Ω)++uεr+1,nuεr,nHp(Ω)\displaystyle\leq\left\|u_{\varepsilon}^{r+l,n}-u_{\varepsilon}^{r+l-1,n}\right\|_{H^{p}\left(\Omega\right)}+\ldots+\left\|u_{\varepsilon}^{r+1,n}-u_{\varepsilon}^{r,n}\right\|_{H^{p}\left(\Omega\right)}
μ¯r+l1uε1,nuε0,nHp(Ω)++μ¯ruε1,nuε0,nHp(Ω)\displaystyle\leq\bar{\mu}^{r+l-1}\left\|u_{\varepsilon}^{1,n}-u_{\varepsilon}^{0,n}\right\|_{H^{p}\left(\Omega\right)}+\ldots+\bar{\mu}^{r}\left\|u_{\varepsilon}^{1,n}-u_{\varepsilon}^{0,n}\right\|_{H^{p}\left(\Omega\right)}
μ¯r(1μ¯l)1μ¯uε1,nHp(Ω),\displaystyle\leq\frac{\bar{\mu}^{r}\left(1-\bar{\mu}^{l}\right)}{1-\bar{\mu}}\left\|u_{\varepsilon}^{1,n}\right\|_{H^{p}\left(\Omega\right)},

which proves the fact that the sequence {uεr,n}r\left\{u_{\varepsilon}^{r,n}\right\}_{r\in\mathbb{N}} is Cauchy in Hp(Ω)H^{p}\left(\Omega\right). Consequently, there exists u¯εnHp(Ω)\bar{u}_{\varepsilon}^{n}\in H^{p}\left(\Omega\right) to which uεr,nu_{\varepsilon}^{r,n} is strongly convergent as rr\to\infty. In addition, it follows from (3.8) that when ll\to\infty,

(3.9) uεr,nu¯εnHp(Ω)Cμ¯r,\left\|u_{\varepsilon}^{r,n}-\bar{u}_{\varepsilon}^{n}\right\|_{H^{p}\left(\Omega\right)}\leq C\bar{\mu}^{r},

and thus, for rr sufficiently large we obtain the convergence of the source term, i.e. F(uεr,n)F(u¯εn)F\left(u_{\varepsilon}^{r,n}\right)\to F\left(\bar{u}_{\varepsilon}^{n}\right) strongly in Hp(Ω)H^{p}\left(\Omega\right) as rr\to\infty, whenever L1L_{1} is dependent of MM or not; see again the choice (2.12) to decide how big the iteration step rr needs to be.

Collectively, we have proved that u¯εn\bar{u}_{\varepsilon}^{n} is identically the semi-discrete solution of the nonlinear regularization (2.11) and further, it coincides the function uεnu_{\varepsilon}^{n} derived from the weak convergence above. Now, it suffices to close the proof of the theorem by combining (3.6) and (3.7). Essentially, we have

vεR+1,nHp(Ω)1R(R+1)r=0RuεR+1,nuεr,nHp(Ω)Cμ¯R+1.\left\|v_{\varepsilon}^{R+1,n}\right\|_{H^{p}\left(\Omega\right)}\leq\frac{1}{R\left(R+1\right)}\sum_{r=0}^{R}\left\|u_{\varepsilon}^{R+1,n}-u_{\varepsilon}^{r,n}\right\|_{H^{p}\left(\Omega\right)}\leq\frac{C\bar{\mu}}{R+1}.

Similar to (3.9), the Cesàro mean is strongly convergent to uεnu_{\varepsilon}^{n} with the rate

wεR,nuεnHp(Ω)C(μ¯R)R.\left\|w_{\varepsilon}^{R,n}-u_{\varepsilon}^{n}\right\|_{H^{p}\left(\Omega\right)}\leq C\left(\frac{\bar{\mu}}{R}\right)^{R}.

Hence, we complete the proof of the theorem. ∎

Corollary 11.

Under the assumptions of Theorems 6 and 10, one has the following error estimate:

uεr,nu(,tn)Hp(Ω)C(μ¯r+εtn(12β)T+tn+p2)for n=1,N1¯.\left\|u_{\varepsilon}^{r,n}-u\left(\cdot,t_{n}\right)\right\|_{H^{p}\left(\Omega\right)}\leq C\left(\bar{\mu}^{r}+\varepsilon^{\frac{t_{n}\left(1-2\beta\right)}{T+t_{n}+\frac{p}{2}}}\right)\quad\text{for }n=\overline{1,N-1}.
Remark 12.

When L1L_{1} is independent of MM, the assumption (3.1) can also be found in some examples, which also aids the applicability of the global Lipschitz case in Subsection 2.2. Observe what have been enlisted in Subsection 1.2. It is immediate to see that the Michaelis–Menten law (N=1N=1 in the de Pillis-Radunskaya law) with F(u)=au/(b+u)F\left(u\right)=au/\left(b+u\right) for a,b>0a,b>0 gives

sup|w|MFw(w)=sup|w|Mab(b+w)2[ab(b+M)2,ab].\sup_{\left|w\right|\leq M}\frac{\partial F}{\partial w}\left(w\right)=\sup_{\left|w\right|\leq M}\frac{ab}{\left(b+w\right)^{2}}\in\left[\frac{ab}{\left(b+M\right)^{2}},\frac{a}{b}\right].

4. Discussions

We have studied a nonlinear spectral regularization to solve a semi-linear backward parabolic equation. The scheme significantly modifies the cut-off method developed in [15, 18, 8] so that it not only fits the nonlinear context under consideration, but also handles certain smoothness of the solution to the forward model in an appropriate manner. In this fashion, our proposed method is convergent in a Hölder-type rate for t(0,T)t\in\left(0,T\right), decreasing backwards in time, and in a logarithmic-type rate for t=0t=0. It is worth mentioning that the strong convergence obtained in HpH^{p} may allow us to gain the convergence of the regularized solution on the boundary with the same rates by the standard trace theorem.

We have also studied the convergence of an iterative method for this nonlinear scheme. To gain the strong convergence, this approximation works with the non-degeneracy of FF, which is a certainly stronger condition than those met in the analysis of the nonlinear scheme. Essentially, we see that the property of the nonlinearity FF plays a pretty much important role in deciding the convergence of the numerical scheme, as postulated in Theorems 9 and 10. This also points out the most difficult issue in solving inverse problems for nonlinear PDEs; compared to the linear cases investigated so far. One may think that the presence of Theorem 9 seems unnecessary (and so is the largeness of the stabilization constant KK taken in (3.3)) since it is clear that the strong convergence of the numerical scheme has already been obtained in Theorem 10. Nevertheless, it can be understood that we have depicted a general procedure to verify the convergence of numerical regularization schemes in the future topics. In fact, such boundedness (as the stability anlysis) obtained in Theorem 9 orientates towards the strong convergence of the Cesàro mean. In some sense, this unravels the possibility that the strong convergence of the numerical sequence is not obtainable, generally hindered by the property of FF. The choice of KK can also be very helpful because the discretization as well as the number of iterations become more effectively economical. In the near future, we wish to understand deeper numerical issues caught in particularly complex networks as presented in Remark 7.

The results of this paper can initiate the convergence analysis of the other classes of nonlinear backward PDEs using the strategy of verifying variational source conditions. One can also attempt to achieve the strong convergence result in the Besov spaces for regularization of the present backward model ()\left(\mathcal{B}\right) in the unbounded domain as it is in agreement with the well-posedness of the forward problem (see e.g. [13]). Observe that although the variational source condition theory does not require the fact that the ill-posed operator 𝒯\mathcal{T} admits the Fréchet derivative, it is self-contained in this framework. Thus, this time we are allowed to not only derive the variational source condition from the spectral source condition, but also apply the iteration-based regularized Gauß-Newton method. The convergence analysis of this method for the nonlinear backward PDEs should also be considered in the forthcoming works.

Acknowledgments

The author thanks Prof. Dr. Mohammad Kazemi (Charlotte, USA) for his support of the author’s research career.

References

  • [1] P. Érdi and J. Tóth. Mathematical Models of Chemical Reactions: Theory and Applications of Deterministic and Stochastic Models. Nonlinear Science. Princeton University Press, 1989.
  • [2] A. B. Ferrari and E. S. Titi. Gevrey regularity for nonlinear analytic parabolic equations. Communications in Partial Differential Equations, 23(1–2):1–16, 1998.
  • [3] J. Flemming, B. Hofmann, and P. Mathé. Sharp converse results for the regularization error using distance functions. Inverse Problems, 27(2):025006, 2011.
  • [4] M. Grasmair. Generalized Bregman distances and convergence rates for non-convex regularization methods. Inverse Problems, 26(11):115014, 2010.
  • [5] D. N. Hào, N. V. Duc, and N. V. Thang. Backward semi-linear parabolic equations with time-dependent coefficients and local Lipschitz source. Inverse Problems, 34(5):055010, 2018.
  • [6] B. Hofmann, B. Kaltenbacher, C. Pöschl, and O. Scherzer. A convergence rates result for Tikhonov regularization in Banach spaces with non-smooth operators. Inverse Problems, 23(3):987–1010, 2007.
  • [7] T. Hohage and F. Weidling. Verification of a variational source condition for acoustic inverse medium scattering problems. Inverse Problems, 31(7):075006, 2015.
  • [8] T. Hohage and F. Weidling. Characterizations of variational source conditions, converse results, and maxisets of spectral regularization methods. SIAM Journal on Numerical Analysis, 55(2):598–620, 2017.
  • [9] T. Hohage and F. Werner. Convergence rates for inverse problems with impulsive noise. SIAM Journal on Numerical Analysis, 52(3):1203–1221, 2014.
  • [10] R. Jaroudi, G. Baravdish, B. T. Johansson, and F. Åström. Numerical reconstruction of brain tumours. Inverse Problems in Science and Engineering, pages 1–21, 2018.
  • [11] B. Kaltenbacher, A. Neubauer, and O. Scherzer. Iterative Regularization Methods for Nonlinear Ill-Posed Problems. Radon Series on Computational and Applied Mathematics. Berlin, Boston: De Gruyter, 2008.
  • [12] C. König, F. Werner, and T. Hohage. Convergence rates for exponentially ill-posed inverse problems with impulsive noise. SIAM Journal on Numerical Analysis, 54(1):341–360, 2016.
  • [13] C. Miao and B. Zhang. The Cauchy problem for semilinear parabolic equations in Besov spaces. Houston Journal of Mathematics, 30(3):829–878, 2004.
  • [14] J. D. Murray. Continuous Population Models for Single Species. In Interdisciplinary Applied Mathematics, volume 17, pages 1–43. Springer, New York, 1993.
  • [15] P. T. Nam. An approximate solution for nonlinear backward parabolic equations. Journal of Mathematical Analysis and Applications, 367(2):337–349, 2010.
  • [16] O. Scherzer, M. Grasmair, H. Grossauer, M. Haltmeier, and F. Lenzen. Variational Methods in Imaging. Applied Mathematical Sciences. Springer, New York, 2008.
  • [17] N. H. Tuan, V. V. Au, V. A. Khoa, and D. Lesnic. Identification of the population density of a species model with nonlocal diffusion and nonlinear reaction. Inverse Problems, 33(5):055019, 2017.
  • [18] N. H. Tuan, L. D. Thang, V. A. Khoa, and T. Tran. On an inverse boundary value problem of a nonlinear elliptic equation in three dimensions. Journal of Mathematical Analysis and Applications, 426(2):1232–1261, 2015.