DoF | 163 | 579 | 2179 | 8451 | 33283 | 526339 |
avg(iter) | 4.2 | 4.0 | 4.0 | 3.8 | 4.2 | 4.3 |
We are interested in modelling the interaction of bacteria and certain nutrient concentration within a porous medium admitting viscous flow. The governing equations in primal-mixed form consist of an advection-reaction-diffusion system representing the bacteria-chemical mass exchange, coupled to the Brinkman problem written in terms of fluid vorticity, velocity and pressure, and describing the flow patterns driven by an external source depending on the local distribution of the chemical species. A priori stability bounds are derived for the uncoupled problems, and the solvability of the full system is analysed using a fixed-point approach. We introduce a primal-mixed finite element method to numerically solve the model equations, employing a primal scheme with piecewise linear approximation of the reaction-diffusion unknowns, while the discrete flow problem uses a mixed approach based on Raviart-Thomas elements for velocity, Nédélec elements for vorticity, and piecewise constant pressure approximations. In particular, this choice produces exactly divergence-free velocity approximations. We establish existence of discrete solutions and show their convergence to the weak solution of the continuous coupled problem. Finally, we report several numerical experiments illustrating the behaviour of the proposed scheme.
Citation: Verónica Anaya, Mostafa Bendahmane, David Mora, Ricardo Ruiz Baier. 2018: On a vorticity-based formulation for reaction-diffusion-Brinkman systems, Networks and Heterogeneous Media, 13(1): 69-94. doi: 10.3934/nhm.2018004
[1] | Verónica Anaya, Mostafa Bendahmane, David Mora, Ricardo Ruiz Baier . On a vorticity-based formulation for reaction-diffusion-Brinkman systems. Networks and Heterogeneous Media, 2018, 13(1): 69-94. doi: 10.3934/nhm.2018004 |
[2] | Javier A. Almonacid, Gabriel N. Gatica, Ricardo Oyarzúa, Ricardo Ruiz-Baier . A new mixed finite element method for the n-dimensional Boussinesq problem with temperature-dependent viscosity. Networks and Heterogeneous Media, 2020, 15(2): 215-245. doi: 10.3934/nhm.2020010 |
[3] | Tasnim Fatima, Ekeoma Ijioma, Toshiyuki Ogawa, Adrian Muntean . Homogenization and dimension reduction of filtration combustion in heterogeneous thin layers. Networks and Heterogeneous Media, 2014, 9(4): 709-737. doi: 10.3934/nhm.2014.9.709 |
[4] | Changli Yuan, Mojdeh Delshad, Mary F. Wheeler . Modeling multiphase non-Newtonian polymer flow in IPARS parallel framework. Networks and Heterogeneous Media, 2010, 5(3): 583-602. doi: 10.3934/nhm.2010.5.583 |
[5] | Thomas Geert de Jong, Georg Prokert, Alef Edou Sterk . Reaction–diffusion transport into core-shell geometry: Well-posedness and stability of stationary solutions. Networks and Heterogeneous Media, 2025, 20(1): 1-14. doi: 10.3934/nhm.2025001 |
[6] | Danielle Hilhorst, Hideki Murakawa . Singular limit analysis of a reaction-diffusion system with precipitation and dissolution in a porous medium. Networks and Heterogeneous Media, 2014, 9(4): 669-682. doi: 10.3934/nhm.2014.9.669 |
[7] | Alexander Mielke, Sina Reichelt, Marita Thomas . Two-scale homogenization of nonlinear reaction-diffusion systems with slow diffusion. Networks and Heterogeneous Media, 2014, 9(2): 353-382. doi: 10.3934/nhm.2014.9.353 |
[8] | James Nolen . A central limit theorem for pulled fronts in a random medium. Networks and Heterogeneous Media, 2011, 6(2): 167-194. doi: 10.3934/nhm.2011.6.167 |
[9] | Bedr'Eddine Ainseba, Mostafa Bendahmane, Yuan He . Stability of conductivities in an inverse problem in the reaction-diffusion system in electrocardiology. Networks and Heterogeneous Media, 2015, 10(2): 369-385. doi: 10.3934/nhm.2015.10.369 |
[10] | Andrea Picco, Lamberto Rondoni . Boltzmann maps for networks of chemical reactions and the multi-stability problem. Networks and Heterogeneous Media, 2009, 4(3): 501-526. doi: 10.3934/nhm.2009.4.501 |
We are interested in modelling the interaction of bacteria and certain nutrient concentration within a porous medium admitting viscous flow. The governing equations in primal-mixed form consist of an advection-reaction-diffusion system representing the bacteria-chemical mass exchange, coupled to the Brinkman problem written in terms of fluid vorticity, velocity and pressure, and describing the flow patterns driven by an external source depending on the local distribution of the chemical species. A priori stability bounds are derived for the uncoupled problems, and the solvability of the full system is analysed using a fixed-point approach. We introduce a primal-mixed finite element method to numerically solve the model equations, employing a primal scheme with piecewise linear approximation of the reaction-diffusion unknowns, while the discrete flow problem uses a mixed approach based on Raviart-Thomas elements for velocity, Nédélec elements for vorticity, and piecewise constant pressure approximations. In particular, this choice produces exactly divergence-free velocity approximations. We establish existence of discrete solutions and show their convergence to the weak solution of the continuous coupled problem. Finally, we report several numerical experiments illustrating the behaviour of the proposed scheme.
Reaction-diffusion systems can explain many phenomena taking place in diverse disciplines such as industrial and environmental processes, biomedical applications, or population dynamics. For instance, non-equilibrium effects associated to mass exchange and local configuration modifications in the concentration of species are usually represented in terms of classical reaction-diffusion equations. These models allow to reproduce chaos, spatiotemporal patterns, rhythmic and oscillatory scenarios, and many other mechanisms. Nevertheless, in most of these applications the reactions do not occur in complete isolation. The species are rather immersed in a fluid, or they move within (and interact with) a fluid-solid continuum, and the motion of the fluid inevitably affects that of the species. In some circumstances, reciprocal effects might be substantially large, therefore leading to local changes in the observed flow patterns. More specifically, here it is assumed that the medium where the chemical reactions develop is a porous material saturated with an incompressible fluid. We remark that we are interested in viscous flows and consequently, the linear momentum and mass conservation for the fluid are governed by the Brinkman equations. The regime under study is relevant to fluids within highly porous structures composed by light fixed particles [7]. As indicated above, we also suppose that the local fluctuations of a species' concentration is important enough to affect the fluid flow. In turn, the reaction-diffusion equations include additional terms accounting for the advection of each species with the fluid velocity, therefore improving the mixing and interaction properties with respect to those observed under pure diffusion effects [13]. Diverse types of continuum-based models resulting in reaction-diffusion-momentum equations have been applied for e.g. enzyme reactions advected with a known Poiseuille velocity profile [16], population kinetics on moving domains [28,35], the control of drug release within soft living tissue [9,15], or biochemical interactions on growing surfaces [12].
We aim at developing numerical solutions to a class of similar systems, also addressing the solvability of the governing equations, the expected properties of the underlying solutions, and the convergence behaviour of suitable finite element schemes. More precisely, at both continuous and discrete levels, the Brinkman equations are set in a mixed form (that is, the associated formulation possesses a saddle-point structure involving the vorticity as additional unknown) whereas the formulation of the reaction-diffusion system is written exclusively in terms of the primal variables, in this case the species' concentration. Such a structure is motivated by the need of accurate vorticity rendering, obtained directly without postprocessing it from typically low-order discrete velocities (which usually leads to insufficiently reliable approximations). The nonlinear reaction-diffusion system will be placed in the framework of general semilinear parabolic equations and its well-posedness, for a fixed velocity, follows from the classical theory of Ladyženskaja [22]. On the other hand, for a fixed species' concentration, a number of recent Brinkman formulations based on vorticity, pressure and velocity are available from the literature [3,5,6,19] (see also [14]). Here we adopt the one introduced in [4], where the well-posedness of the flow equations is established thanks to the classical Babuška-Brezzi framework. In turn, the fully coupled problem is analysed using a Schauder fixed-point approach. Perhaps the closest contributions in the spirit of the present study are the error estimation for finite element formulations to doubly-diffusive Boussinesq flows presented in [2], the two-sidedly degenerate chemotaxis-fluid coupled system from [11] whose solvability relies on a regularisation step combined with compactness arguments; and the general analysis for degenerate parabolic equations idealising reactive solute transport in porous media recently carried out in [21].
A further goal is to propose a numerical method based on piecewise linear and continuous polynomials for the species' concentrations, whereas the set of flow equations is discretised using a mixed approach based on Raviart-Thomas elements for the approximation of the velocity, Nédélec elements for vorticity, and piecewise constants for the pressure. The computational burden of this algorithm is comparable to classical low-cost approximations as the so-called MINI element for velocity-pressure formulations. On top of that, the proposed finite element method provides divergence-free velocities, thus preserving an essential constraint of the underlying physical system. Based on the present formulation, the stability properties of different staggered coupling techniques have been recently analysed in [24]. Other recent contributions regarding the numerical discretisation for related models for reactive solute transport in porous media (but using Darcy descriptions of flow), include the formulation and convergence studies for mixed and conformal methods as presented in [1, 8, 20, 27, 30, 31].
After this introductory section, the remainder of the paper is structured in the following manner. Section 2 states the model problem and introduces the concept of weak solutions, together with an auxiliary coupled problem. The solvability of these auxiliary equations is established in Section 3, using a fixed-point approach. The proof of uniqueness of weak solution is postponed to Section 4. Section 5 deals with the numerical approximation of the model problem, using mixed finite elements and a first-order backward Euler time advancing scheme. We provide in Section 6 a set of numerical tests to illustrate the properties of the numerical scheme and the features of the coupled model, and we close in Section 7 with a summary and discussion of possible extensions to this work.
Let
$∂tc+u⋅∇c−div(Dc(c)∇c)=Gc(c,s),∂ts+u⋅∇s−div(Ds(s)∇s)=Gs(c,s),K−1u+√μcurlω+∇p=sg+f,ω−√μcurlu=0,divu=0,$ | (2.1) |
where
Equations (2.1) are complemented with the following boundary and initial data:
$(cu−Dc(c)∇c)⋅n=(su−Ds(s)∇s)⋅n=0(x,t)∈∂Ω×(0,T],u⋅n=u∂, ω×n=ω∂(x,t)∈∂Ω×(0,T],c=c0, s=s0(x,t)∈Ω×{0},$ | (2.2) |
representing that the species cannot leave the medium and that a normal velocity
We suppose that the permeability tensor
$ \mathit{\boldsymbol{v}}^{\tt t}\mathbb{K}^{-1}(\mathit{\boldsymbol{x}})\mathit{\boldsymbol{v}} \geq C |\mathit{\boldsymbol{v}}|^2 \;\;\;\forall\;\mathit{\boldsymbol{v}}\in \mathbb{R}^3, \;\forall\; \mathit{\boldsymbol{x}}\in\Omega.$ | (2.3) |
In addition, the diffusivities are assumed always positive, coercive, and continuous: for
$ D_i:[0, 1]\mapsto \mathbb{R}^+\ \text{is continuous},\;\;\; 0 < D^{\min} \leq D_i(m)\leq D^{\max} < \infty,\;\;\; m\in \mathbb{R}. $ | (2.4) |
Regarding the reaction terms
$ |Gc(c,s)|≤Cc(1+|c|+|s|),|Gs(c,s)|≤Cs(1+|c|+|s|)forallc,s≥0,|Gκ(c1,s1)|−Gκ(c2,s2)≤C(|c1−c2|+|s1−s2|),forallc1,c2,s1,s2≥0andκ=c,s,Gc(c,s)=ν1andGs(c,s)=ν2ifc≤0ors≤0, $ | (2.5) |
for some parameters
$ {c_0},{s_0} \ge 0,\;\;\;\;\;{c_0},{s_0} \in {{\rm{L}}^\infty }(\Omega ). $ | (2.6) |
Observe that for constant coefficients
According to (2.2), let us introduce the functional spaces
$H0(div;Ω)={v∈H(div;Ω):v⋅n=0 on ∂Ω},H0(curl;Ω)={z∈H(curl;Ω):z×n=0 on ∂Ω},$ |
where
Definition 2.1. In view of the properties outlined in Section 2.2, we shall say that the function
$s,c∈L∞(0,T;L2(Ω))∩L2(0,T;H1(Ω)),∂ts,∂tc∈L2(0,T;H1(Ω)′),u∈L2(0,T;H(div;Ω)),ω∈L2(0,T;H(curl;Ω)),p∈L2(0,T;L20(Ω)),$ |
and
$ \int_0^T \langle {\partial _t} c, m^c\rangle{\rm{d}}t + \iint_{\Omega_T} (D_c(c)\nabla c- c\mathit{\boldsymbol{u}})\cdot \nabla m^c {\rm{d}}\mathit{\boldsymbol{x}}\;{\rm{d}}t \\ \;\;\;\;\;\;\;\;\;\;= \iint_{\Omega_T} G_c(c, s) m^c {\rm{d}}\mathit{\boldsymbol{x}}\;{\rm{d}}t, \\ \int_0^T \langle {\partial _t} s, m^s\rangle{\rm{d}}t + \iint_{\Omega_T} (D_s(s)\nabla s- s\mathit{\boldsymbol{u}})\cdot \nabla m^s {\rm{d}}\mathit{\boldsymbol{x}}\;{\rm{d}}t \\ \;\;\;\;\;\;\;\;\;\; = \iint_{\Omega_T} G_s(c, s) m^s {\rm{d}}\mathit{\boldsymbol{x}}\;{\rm{d}}t, \\ \iint_{\Omega_T}\mathbb{K}^{-1}\mathit{\boldsymbol{u}}\cdot\mathit{\boldsymbol{v}}\;{\rm{d}}\mathit{\boldsymbol{x}}\;{\rm{d}}t + \sqrt{\mu}\iint_{\Omega_T}{\boldsymbol{\rm{curl}}}\mathit{\boldsymbol{\omega}}\cdot\mathit{\boldsymbol{v}}\;{\rm{d}}\mathit{\boldsymbol{x}}\;{\rm{d}}t - \iint_{\Omega_T}p\;{\rm{div}}\;\mathit{\boldsymbol{v}}\;{\rm{d}}\mathit{\boldsymbol{x}}\;{\rm{d}}t \\ \;\;\;\;\;\;\;\;\;\;= \iint_{\Omega_T} (s\mathit{\boldsymbol{g}}+\mathit{\boldsymbol{f}})\cdot \mathit{\boldsymbol{v}} \;{\rm{d}}\mathit{\boldsymbol{x}}\;{\rm{d}}t, \\ \sqrt{\mu}\iint_{\Omega_T}{\boldsymbol{\rm{curl}}}\mathit{\boldsymbol{z}}\cdot\mathit{\boldsymbol{u}}\;{\rm{d}}\mathit{\boldsymbol{x}}\;{\rm{d}}t - \iint_{\Omega_T}\mathit{\boldsymbol{\omega}}\cdot\mathit{\boldsymbol{z}}\;{\rm{d}}\mathit{\boldsymbol{x}}\;{\rm{d}}t = 0, \\ -\iint_{\Omega_T}\;q\;{\rm{div}}\;\mathit{\boldsymbol{u}}\;{\rm{d}}\mathit{\boldsymbol{x}}\;{\rm{d}}t = 0, $ |
for all
Our first main result states existence of weak solutions to the reaction-diffusion-Brinkman equations. The analysis of uniqueness will be postponed to Section 4.
Theorem 2.2. Assume that conditions (2.3), (2.4) and (2.5) hold. If
The proof of this result is based on an application of Schauder's fixed-point theorem (in an appropriate functional setting) with the derivation of a priori estimates and the compactness arguments, to the following uncoupled system
$∂tc+u⋅∇c−div(Dc(c)∇c)=Gc(c,s),∂ts+u⋅∇s−div(Ds(s)∇s)=Gs(c,s),K−1u+√μcurlω+∇p=ˆsg+f,ω−√μcurlu=0,divu=0,$ | (2.7) |
where
In this section we prove the existence of solutions to (2.7), looking first at the solvability of the uncoupled systems.
Let us recall the following abstract result (see e.g. [17,Theorem 1.3]):
Theorem 3.1. Let
$ \displaystyle\, \sup\limits_{\stackrel{\scriptstyle y\in {\mathcal X}}{y\ne0}}\frac{\mathcal A(x, y)}{\|y\|_{\mathcal X}}\, \ge\, \bar\beta\, \|x\|_{\mathcal X}\;\;\;\; \forall\, x\in \mathcal X. $ |
Then, there exists a unique
$ \mathcal A(x, y) = \mathcal G(y)\;\;\;\; \forall\, y\in \mathcal X. $ |
Moreover, there exists
$\|x\|_{\mathcal X}\le C\|\mathcal G\|_{\mathcal X'}.$ |
Let us also consider the kernel of the bilinear form
$X:={v∈H0(div;Ω):∫Ωqdivvdx=0,∀q∈L20(Ω)}={v∈H0(div;Ω):divv=0 a.e. in Ω}.$ |
Moreover, we endow the space
$\Vert \mathit{\boldsymbol{z}}\Vert_{{\boldsymbol{\rm{H}}}({\boldsymbol{\rm{curl}}};\Omega)}^2: = \Vert\mathit{\boldsymbol{z}}\Vert_{0, \Omega}^2 +\mu\Vert{\boldsymbol{\rm{curl}}}\mathit{\boldsymbol{z}}\Vert_{0, \Omega}^2, $ |
and recall the following inf-sup condition (cf. [17]): there exists
$ \sup\limits_{\stackrel{\scriptstyle\mathit{\boldsymbol{v}}\in{\boldsymbol{\rm{H}}}({\rm{div}};\Omega)}{\mathit{\boldsymbol{v}}\ne0}}\frac{{\Biggl | \displaystyle \int_{\Omega}}\;q\;{\rm{div}}\;\mathit{\boldsymbol{v}}\;{\rm{d}}\mathit{\boldsymbol{x}} {\Biggr |}}{\Vert\mathit{\boldsymbol{v}}\Vert_{{\boldsymbol{\rm{H}}}({\rm{div}};\Omega)}}\ge \beta_2\Vert q\Vert_{0, \Omega}\;\;\;\;\forall q\in{\rm{L}}_{0}^2(\Omega). $ | (3.1) |
In view of invoking Schauder's fixed-point theorem to establish solvability of (2.7), we introduce the following closed subset of the Banach space
$ {\mathcal{K}_\phi = \{\phi \in{\rm{L}}^2(\Omega_T) \: : \ 0\le \phi(\mathit{\boldsymbol{x}}, t)\le e^{(\beta-\lambda)t} k_m, \text{ for a.e. }(\mathit{\boldsymbol{x}}, t)\in \Omega_T \}, } $ | (3.2) |
for
Lemma 3.2. Assume that
$∫ΩK−1u⋅vdx+√μ∫Ωcurlω⋅vdx−∫Ωpdivvdx=∫Ω(ˆsg+f)⋅vdx,√μ∫Ωcurlz⋅udx−∫Ωω⋅zdx=0,−∫Ωqdivudx=0,$ | (3.3) |
admits a unique solution
$‖u‖H(div;Ω)+‖ω‖H(curl;Ω)+‖p‖0,Ω≤C(‖ˆs‖0,Ω‖g‖∞,Ω+‖f‖0,Ω+‖u∂‖−1/2,∂Ω+‖ω∂‖−1/2,∂Ω). $ | (3.4) |
Proof. First, we observe that, owing to the inf-sup condition (3.1), problem (3.3) is equivalent to: Find
$∫ΩK−1u⋅vdx+√μ∫Ωcurlω⋅vdx=∫Ω(ˆsg+f)⋅vdx,∀v∈X,√μ∫Ωcurlz⋅udx−∫Ωω⋅zdx=0,∀z∈H0(curl;Ω).$ |
Then, the desired result follows from Theorem 3.1, repeating the argument employed in the proofs of [4,Theorem 2.2 and Corollary 2.1].
On the other hand, given a fixed velocity
Lemma 3.3. For any
$∫T0⟨∂tc,mc⟩dt+∬ΩT(Dc(c)∇c−cu)⋅∇mcdxdt=∬ΩTGc(c,s)mcdxdt,∫T0⟨∂ts,ms⟩dt+∬ΩT(Ds(s)∇s−su)⋅∇msdxdt=∬ΩTGs(c,s)msdxdt,$ | (3.5) |
is uniquely solvable and there exists
$ \|c\|_{{\rm{L}}^2(0, T;{\rm{H}}^1(\Omega))}+\|s\|_{{\rm{L}}^2(0, T;{\rm{H}}^1(\Omega))} \leq C.$ |
Proof. It suffices to combine assumptions (2.4), (2.6) with the general result for quasilinear parabolic problems given in [22,Section 5].
As the next step, we consider a constant
$ c = e^{\lambda t}\phi_c\;\;{\rm{and}}\;\;s = e^{\lambda t}\phi_s. $ | (3.6) |
Then
$\left\{Dc(c):=Dc(eλtc) and Ds(s):=Ds(eλts),Gc(c,s):=−λc+e−λtGc(eλtc,eλts),Gs(c,s):=−λs+e−λtGs(eλtc,eλts).\right.$ |
We now introduce a map
Lemma 3.4. Let
Proof.
$ \partial_t c_n- {\rm{div}}(D_c(c_n)\nabla c_n)+\mathit{\boldsymbol{u}}\cdot \nabla c_n = -\lambda c_n+ e^{-\lambda t}G_{c}(e^{\lambda t}c_n, e^{\lambda t}s_n), \;\;\;\;\mbox{ in }\Omega_T. $ | (3.7) |
Multiplying (3.7) by
$12ddt∫Ω|c−n|2dx+Dmin $ | (3.8) |
Now, we use that
$ -\int_\Omega c_n^- \mathit{\boldsymbol{u}}\cdot \nabla c_n{\rm{d}}\mathit{\boldsymbol{x}} = \frac{1}{2}\int_\Omega \mathit{\boldsymbol{u}}\cdot \nabla (c_n^-)^2{\rm{d}}\mathit{\boldsymbol{x}} = 0, $ |
and
$\begin{cases} \displaystyle -\int_\Omega e^{-\lambda t}G_{c}(e^{\lambda t}c_n, e^{\lambda t}s_n)c_n^- {\rm{d}}\mathit{\boldsymbol{x}} \leq 0&\text{ if }c_n\leq 0, \\ \displaystyle -\int_\Omega e^{-\lambda t}G_{c}(e^{\lambda t}c_n, e^{\lambda t}s_n)c_n^- {\rm{d}}\mathit{\boldsymbol{x}} = 0& \text{ if }c_n > 0. \end{cases}$ |
According to the positivity of the second and the fourth terms in the left-hand side of (3.8), we get
$ \frac{1}{2}\frac{d}{{\rm{d}}t}\int_\Omega {\left| {{c_n^-}} \right|}^2{\rm{d}}\mathit{\boldsymbol{x}}\le 0. $ |
Since
Next, we let
${\xi_c: = (c_n-e^{(\beta-\lambda) t}k)^+}, \;\;{\rm{with}}\;\beta \geq \lambda,\;\;{\rm{and\;with}}\;\;k = \sup\{k_c, k_s\}, $ |
and to integrate over
$ \begin{align} &\frac{1}{2}\frac{d}{{\rm{d}}t}\int_\Omega \xi_c^2 {\rm{d}}\mathit{\boldsymbol{x}} + (\beta-\lambda)\int_\Omega e^{(\beta-\lambda) t}k\, \xi_c {\rm{d}}\mathit{\boldsymbol{x}} +D^{\min}\!\!\int_\Omega \!\left| {\nabla \xi_c} \right|^2\! {\rm{d}}\mathit{\boldsymbol{x}} \nonumber \\&\;\;\;\;+ \int_\Omega\! \xi_c\, \mathit{\boldsymbol{u}}\cdot \nabla c_n{\rm{d}}\mathit{\boldsymbol{x}} +\lambda \int_\Omega\! c_n\, \xi_c\!{\rm{d}}\mathit{\boldsymbol{x}} \nonumber\\ &\;\;\;\; = \frac{1}{2}\frac{d}{{\rm{d}}t}\int_\Omega \xi_c^2 {\rm{d}}\mathit{\boldsymbol{x}}+\beta\int_\Omega e^{(\beta-\lambda) t}k\, \xi_c {\rm{d}}\mathit{\boldsymbol{x}} +D^{\min}\int_\Omega \left| {\nabla \xi_c} \right|^2 {\rm{d}}\mathit{\boldsymbol{x}} \nonumber\\&\;\;\;\;+ \frac{1}{2}\int_\Omega \mathit{\boldsymbol{u}}\cdot (\nabla \xi_c^2){\rm{d}}\mathit{\boldsymbol{x}} +\lambda \int_\Omega \xi_c^2{\rm{d}}\mathit{\boldsymbol{x}} \nonumber\\&\;\;\;\; \leq \int_\Omega e^{-\lambda t}G_{c}(e^{\lambda t}c_n, e^{\lambda t}s_n)\xi_c {\rm{d}}\mathit{\boldsymbol{x}} \nonumber\\&\;\;\;\;\leq C_6 \Biggl(\int_\Omega e^{(\beta-\lambda) t}k \xi_c {\rm{d}}\mathit{\boldsymbol{x}} +\int_\Omega \xi_c^2{\rm{d}}\mathit{\boldsymbol{x}}+\int_\Omega \xi_s^2{\rm{d}}\mathit{\boldsymbol{x}}\Biggl), \label{eq2:est-class} \end{align} $ | (3.9) |
where
$ \begin{align} &\frac{1}{2}\frac{d}{{\rm{d}}t}\int_\Omega \xi_s^2 {\rm{d}}\mathit{\boldsymbol{x}}+(\beta-\lambda)\int_\Omega e^{(\beta-\lambda) t}k\xi_s {\rm{d}}\mathit{\boldsymbol{x}} +\! D^{\min}\!\!\int_\Omega \left| {\nabla \xi_s} \right|^2 {\rm{d}}\mathit{\boldsymbol{x}}\nonumber\\&\;\;\;\;+\! \int_\Omega \xi_s \mathit{\boldsymbol{u}}\cdot \nabla s_n{\rm{d}}\mathit{\boldsymbol{x}} +\lambda \! \int_\Omega s_n \xi_s{\rm{d}}\mathit{\boldsymbol{x}} \nonumber \\ \label{eq2:est-class-eqsn}&\;\;\;\;\leq C_7 \Biggl(\int_\Omega e^{(\beta-\lambda) t}k \xi_s {\rm{d}}\mathit{\boldsymbol{x}} +\int_\Omega \xi_c^2{\rm{d}}\mathit{\boldsymbol{x}} +\int_\Omega \xi_s^2{\rm{d}}\mathit{\boldsymbol{x}}\Biggl), \end{align} $ | (3.10) |
for some constant
$ \int_\Omega \mathit{\boldsymbol{u}}\cdot \nabla (\xi_c)^2{\rm{d}}\mathit{\boldsymbol{x}} = 0, \;\;\;\;\int_\Omega \mathit{\boldsymbol{u}}\cdot \nabla (\xi_s)^2{\rm{d}}\mathit{\boldsymbol{x}} = 0, $ |
which inserted into (3.9) and (3.10), implies that
$ \begin{array}{lll} &\displaystyle\frac{1}{2}\frac{d}{{\rm{d}}t}\int_\Omega \xi_c^2 {\rm{d}}\mathit{\boldsymbol{x}} \\ &\;\;\;\;+\displaystyle\frac{1}{2}\frac{d}{{\rm{d}}t}\int_\Omega \xi_s^2 {\rm{d}}\mathit{\boldsymbol{x}} +(\beta-C_6-C_7)e^{(\beta-\lambda) t}k\Biggl(\int_\Omega \xi_c {\rm{d}}\mathit{\boldsymbol{x}}+\int_\Omega \xi_s {\rm{d}}\mathit{\boldsymbol{x}}\Biggl)\\ &\;\;\;\; \displaystyle +(\lambda-C_6-C_7) \int_\Omega \xi_c^2{\rm{d}}\mathit{\boldsymbol{x}} +(\lambda-C_6-C_7) \int_\Omega \xi_s^2{\rm{d}}\mathit{\boldsymbol{x}} \leq 0. \end{array}$ | (3.11) |
Finally for
$ \frac{d}{{\rm{d}}t}\int_\Omega \left| {\xi_c} \right|^2 {\rm{d}}\mathit{\boldsymbol{x}} +\frac{d}{{\rm{d}}t}\int_\Omega \left| {\xi_s} \right|^2 {\rm{d}}\mathit{\boldsymbol{x}} \le 0, $ |
and exploiting that
$ \begin{array}{ll}\displaystyle &\displaystyle\frac{1}{2}\frac{d}{{\rm{d}}t}\int_\Omega \left| {c_n} \right|^2{\rm{d}}\mathit{\boldsymbol{x}} +D^{\min}\int_\Omega \left| {\nabla c_n} \right|^2 {\rm{d}}\mathit{\boldsymbol{x}}+ \int_\Omega c_n \mathit{\boldsymbol{u}}\cdot \nabla c_n{\rm{d}}\mathit{\boldsymbol{x}} +\lambda \int_\Omega \left| {c_n} \right|^2{\rm{d}}\mathit{\boldsymbol{x}}\\ & \;\;\;\;\;\;\;\; \displaystyle\leq \int_\Omega e^{-\lambda t}G_{c}(e^{\lambda t}c_n, e^{\lambda t}s_n)c_n {\rm{d}}\mathit{\boldsymbol{x}} . \end{array}$ | (3.12) |
Invoking the boundedness of
$ \int_\Omega c_n \mathit{\boldsymbol{u}}\cdot \nabla c_n{\rm{d}}\mathit{\boldsymbol{x}} = \frac{1}{2}\int_\Omega \mathit{\boldsymbol{u}}\cdot \nabla (c_n)^2{\rm{d}}\mathit{\boldsymbol{x}} = 0, $ |
we deduce from (3.12) the following bound
$ \displaystyle\frac{1}{2}\frac{d}{{\rm{d}}t}\int_\Omega \left| {c_n} \right|^2 {\rm{d}}\mathit{\boldsymbol{x}}+ D^{\min} \int_\Omega \left| {\nabla c_n} \right|^2 {\rm{d}}\mathit{\boldsymbol{x}} \le C_8, $ |
valid for some constant
$ \begin{array}{ll}&\displaystyle \left| {\int_0^T\left \langle \partial_t c_n, \varphi\right \rangle {\rm{d}}t} \right|\\& = \displaystyle \Biggl | -\int_\Omega D_c(c_n)\nabla c_n \cdot \varphi{\rm{d}}\mathit{\boldsymbol{x}}+\int_\Omega c_n \mathit{\boldsymbol{u}}\cdot \nabla \varphi{\rm{d}}\mathit{\boldsymbol{x}} \\ & \displaystyle -\lambda \int_\Omega c_n \varphi{\rm{d}}\mathit{\boldsymbol{x}} + \int_\Omega e^{-\lambda t}G_{c}(e^{\lambda t}c_n, e^{\lambda t}s_n)\varphi {\rm{d}}\mathit{\boldsymbol{x}} \Biggl | \\ &\;\;\;\;\le D^{\max} \left\|{\nabla c_n}\right \|_{{\rm{L}}^2(\Omega_T)} \left\|{\varphi}\right \|_{{\rm{L}}^2(\Omega_T)}+ C_9 \left\|{ \mathit{\boldsymbol{u}}}\right \|_{0, \Omega} \left\|{\nabla \varphi}\right \|_{{\rm{L}}^2(\Omega_T)} \\ & \;\;\;\;+ C_{10}\Bigl(\left\|{c_n}\right \|_{{\rm{L}}^2(\Omega_T)} +\left\|{s_n}\right \|_{{\rm{L}}^2(\Omega_T)}\Bigl) \left\|{\varphi}\right \|_{{\rm{L}}^2(\Omega_T)}\\ &\le C_{11}\left\|{\varphi}\right \|_{{\rm{L}}^2(0, T;{\rm{H}}^1(\Omega))}, \end{array} $ |
for some constants
$ \left\|{ \partial_t c_n}\right \|_{{\rm{L}}^2(0, T;(H^1(\Omega))')} \le C_{11}. $ | (3.13) |
Working on the same lines as for
In summary, Lemmas 3.2, and 3.4 imply that there exists
$(c, s, \mathit{\boldsymbol{u}}, \mathit{\boldsymbol{\omega}}, p)\in{\rm{L}}^2(0, T;{\rm{H}}^1(\Omega))^2 \times{\boldsymbol{\rm{H}}}({\rm{div}};\Omega)\times{\boldsymbol{\rm{H}}}({\boldsymbol{\rm{curl}}};\Omega)\times {\rm{L}}_0^2(\Omega)$ |
such that, up to extracting subsequences if necessary,
$\begin{cases} &(c_n, s_n) \to (c, s)\;{\rm{in}}\;{\rm{L}}^2(\Omega_T, \mathbb{R}^2)\;{\rm{strongly\;and\;a.e.}},\;{\rm{and\;in}}\;{\rm{L}}^2(0, T;{\rm{H}}^1(\Omega))\;{\rm{weakly}}, \\ &(\mathit{\boldsymbol{u}}_n, \mathit{\boldsymbol{\omega}}_n, p_n) \to (\mathit{\boldsymbol{u}}, \mathit{\boldsymbol{\omega}}, p)\;{\rm{in}}\;{\boldsymbol{\rm{H}}}({\rm{div}};\Omega)\times{\boldsymbol{\rm{H}}}({\boldsymbol{\rm{curl}}};\Omega)\times {\rm{L}}_0^2(\Omega)\;{\rm{weakly}}, \\ \end{cases}$ |
and consequently, the continuity of
$ \mathcal{S} = \left\{ s\in{\rm{L}}^2(0, T;{\rm{H}}^1(\Omega)) \ : \ \partial_t s \in {\rm{L}}^2(0, T;({\rm{H}}^1(\Omega))') \right\}. $ | (3.14) |
Appealing to the theory of compact sets [33], the compactness of the map
$ \begin{align} &\int_0^T \langle {\partial _t} c, m^c\rangle{\rm{d}}t + \iint_{\Omega_T} (D_c(c)\nabla c-c\mathit{\boldsymbol{u}})\cdot \nabla m^c {\rm{d}}\mathit{\boldsymbol{x}}\;{\rm{d}}t = \iint_{\Omega_T} G_{c}(c, s) m^c {\rm{d}}\mathit{\boldsymbol{x}}\;{\rm{d}}t, \nonumber \\&\int_0^T \langle {\partial _t} s, m^s\rangle{\rm{d}}t + \iint_{\Omega_T} (D_s(s)\nabla s-s\mathit{\boldsymbol{u}})\cdot \nabla m^s {\rm{d}}\mathit{\boldsymbol{x}}\;{\rm{d}}t = \iint_{\Omega_T} G_{s}(c, s) m^s {\rm{d}}\mathit{\boldsymbol{x}}\;{\rm{d}}t, \nonumber \\ &\iint_{\Omega_T}\mathbb{K}^{-1}\mathit{\boldsymbol{u}}\cdot\mathit{\boldsymbol{v}}\;{\rm{d}}\mathit{\boldsymbol{x}}\;{\rm{d}}t + \sqrt{\mu}\iint_{\Omega_T}\mathit{\boldsymbol{v}}\cdot{\boldsymbol{\rm{curl}}}\;\mathit{\boldsymbol{\omega}}\;{\rm{d}}\mathit{\boldsymbol{x}}\;{\rm{d}}t -\iint_{\Omega_T}\;p\;{\rm{div}}\;\mathit{\boldsymbol{v}}\;{\rm{d}}\mathit{\boldsymbol{x}}\;{\rm{d}}t \label{wf-reg} \\& \;\;\;\;\;\;\;\;\iint_{\Omega_T} (s\mathit{\boldsymbol{g}}+\mathit{\boldsymbol{f}})\cdot \mathit{\boldsymbol{v}}\;{\rm{d}}\mathit{\boldsymbol{x}}\;{\rm{d}}t, \nonumber \\ &\sqrt{\mu}\iint_{\Omega_T}\mathit{\boldsymbol{u}}\cdot{\boldsymbol{\rm{curl}}}\mathit{\boldsymbol{z}} {\rm{d}}\mathit{\boldsymbol{x}}\;{\rm{d}}t - \iint_{\Omega_T}\mathit{\boldsymbol{\omega}}\cdot\mathit{\boldsymbol{z}}{\rm{d}}\mathit{\boldsymbol{x}}\;{\rm{d}}t = 0, \nonumber \\&-\iint_{\Omega_T}\;q\;{\rm{div}}\;\mathit{\boldsymbol{u}}\;{\rm{d}}\mathit{\boldsymbol{x}}\;{\rm{d}}t = 0, \nonumber \end{align} $ | (3.15) |
for all
The following result completes the analysis of unique solvability for the reaction-diffusion-Brinkman system.
Theorem 4.1. Assume (2.3)-(2.5) hold, and let
$ \label{est-thei-uniq} \begin{split} &\iint_{\Omega_t} \Bigl(\left| {\mathcal{C}} \right|^2 +\left| {\mathcal{S}} \right|^2 +\left| {\mathcal{U}} \right|^2+\left| {\mathcal{W}} \right|^2+\left| {\mathcal{P}} \right|^2\Bigl){\rm{d}}\mathit{\boldsymbol{x}}\;{\rm{d}}\sigma \\ & \le C \Biggl(\int_\Omega \Bigl(\left| {c_{1, 0}(\mathit{\boldsymbol{x}})-c_{2, 0}(\mathit{\boldsymbol{x}})} \right|^2+\left| {s_{1, 0}(\mathit{\boldsymbol{x}})-s_{2, 0}(\mathit{\boldsymbol{x}})} \right|^2 \Bigl){\rm{d}}\mathit{\boldsymbol{x}} \\ & \;\;\;\;\;\;\;\;+\iint_{\Omega_t}\Bigl(\left| {\mathit{\boldsymbol{g}}_1-\mathit{\boldsymbol{g}}_2} \right|^2+\left| {\mathit{\boldsymbol{f}}_1-\mathit{\boldsymbol{f}}_2} \right|^2\Bigl){\rm{d}}\mathit{\boldsymbol{x}}\;{\rm{d}}\sigma \!\!\Biggl), \end{split} $ | (4.1) |
where
Proof. First, we observe that the
$\begin{split} &\iint_{\Omega_t}\mathbb{K}^{-1}{\mathcal{U}}\cdot\mathit{\boldsymbol{v}}\;{\rm{d}}\mathit{\boldsymbol{x}}\;{\rm{d}}\sigma + \sqrt{\mu}\iint_{\Omega_t}\mathit{\boldsymbol{v}}\cdot{\boldsymbol{\rm{curl}}}\;{\mathcal{W}}\;{\rm{d}}\mathit{\boldsymbol{x}}\;{\rm{d}}\sigma - \iint_{\Omega_t}{\mathcal{P}}\;{\rm{div}}\;\mathit{\boldsymbol{v}}\;{\rm{d}}\mathit{\boldsymbol{x}}\;{\rm{d}}\sigma \\ & \;\;\;\;\;\;\;\;= \iint_{\Omega_t} (s_1\mathit{\boldsymbol{g}}_1-s_2\mathit{\boldsymbol{g}}_2) \cdot \mathit{\boldsymbol{v}}\; {\rm{d}}\mathit{\boldsymbol{x}}\;{\rm{d}}\sigma +\iint_{\Omega_t}(\mathit{\boldsymbol{f}}_1-\mathit{\boldsymbol{f}}_2)\cdot \mathit{\boldsymbol{v}} \;{\rm{d}}\mathit{\boldsymbol{x}}\;{\rm{d}}\sigma , \\ &\sqrt{\mu}\iint_{\Omega_t}{\mathcal{U}}\cdot{\boldsymbol{\rm{curl}}}\;\mathit{\boldsymbol{z}}\;{\rm{d}}\mathit{\boldsymbol{x}}\;{\rm{d}}\sigma - \iint_{\Omega_t}{\mathcal{W}}\cdot\mathit{\boldsymbol{z}}\;{\rm{d}}\mathit{\boldsymbol{x}}\;{\rm{d}}\sigma = 0, \\ &-\iint_{\Omega_t}\;q\;{\rm{div}}\;{\mathcal{U}}\;{\rm{d}}\mathit{\boldsymbol{x}}\;{\rm{d}}\sigma = 0, \end{split}$ | (4.2) |
for
After substituting
$ \begin{align}\nonumber&\iint_{\Omega_t}\left| {\mathcal{U}} \right|^2{\rm{d}}\mathit{\boldsymbol{x}}\;{\rm{d}}\sigma + \iint_{\Omega_t}\left| {\mathcal{W}} \right|^2 {\rm{d}}\mathit{\boldsymbol{x}}\;{\rm{d}}\sigma + \iint_{\Omega_t}\left| {\mathcal{P}} \right|^2{\rm{d}}\mathit{\boldsymbol{x}}\;{\rm{d}}\sigma \\& \leq C_{14}\iint_{\Omega_t}|\mathcal{S}|^2{\rm{d}}\mathit{\boldsymbol{x}}\;{\rm{d}}\sigma +C_{15}\Biggl(\iint_{\Omega_t}|\mathit{\boldsymbol{g}}_1-\mathit{\boldsymbol{g}}_2|^2{\rm{d}}\mathit{\boldsymbol{x}}\;{\rm{d}}\sigma +\iint_{\Omega_t}|\mathit{\boldsymbol{f}}_1-\mathit{\boldsymbol{f}}_2|^2{\rm{d}}\mathit{\boldsymbol{x}}\;{\rm{d}}\sigma \Biggl), \label{wf-global-uniq2} \end{align} $ | (4.3) |
for some constant
$\begin{split} &-\int_0^t \langle \mathcal{C}, {\partial _t} m^c\rangle{\rm{d}}\sigma + \iint_{\Omega_t} (D_c(c_1)\nabla c_1- D_c(c_2)\nabla c_2)\cdot \nabla m^c {\rm{d}}\mathit{\boldsymbol{x}}\;{\rm{d}}\sigma \\ & - \iint_{\Omega_t} c_1{\mathcal{U}}\cdot \nabla m^c {\rm{d}}\mathit{\boldsymbol{x}}\;{\rm{d}}\sigma -\iint_{\Omega_t} \mathcal{C}\mathit{\boldsymbol{u}}_2\cdot \nabla m^c {\rm{d}}\mathit{\boldsymbol{x}}\;{\rm{d}}\sigma \\ &\;\;\;\; = \int_{\Omega}\mathcal{C}_0(\mathit{\boldsymbol{x}})m^c(\mathit{\boldsymbol{x}}, 0){\rm{d}}\mathit{\boldsymbol{x}} +\iint_{\Omega_t} (G_{c}(c_1, s_1)-G_{c}(c_2, s_2)) m^c {\rm{d}}\mathit{\boldsymbol{x}}\;{\rm{d}}\sigma , \\ &-\int_0^t \langle \mathcal{S}, {\partial _t} m^s\rangle{\rm{d}}\sigma + \iint_{\Omega_t} (D_s(s_1)\nabla s_1- D_s(s_2)\nabla s_2)\cdot \nabla m^s {\rm{d}}\mathit{\boldsymbol{x}}\;{\rm{d}}\sigma \\ &- \iint_{\Omega_t} s_1{\mathcal{U}}\cdot \nabla m^s {\rm{d}}\mathit{\boldsymbol{x}}\;{\rm{d}}\sigma -\iint_{\Omega_t} \mathcal{S}\mathit{\boldsymbol{u}}_2\cdot \nabla m^s {\rm{d}}\mathit{\boldsymbol{x}}\;{\rm{d}}\sigma \\ &\;\;\;\; = \int_{\Omega}\mathcal{S}_0(\mathit{\boldsymbol{x}})m^c(\mathit{\boldsymbol{x}}, 0){\rm{d}}\mathit{\boldsymbol{x}} +\iint_{\Omega_t} (G_{s}(c_1, s_1)-G_{s}(c_2, s_2)) m^s {\rm{d}}\mathit{\boldsymbol{x}}\;{\rm{d}}\sigma , \nonumber \end{split}$ | (4.4) |
for
$ m^i(\mathit{\boldsymbol{x}}, t) = \begin{cases} &\displaystyle \int_{t}^{t_0}(\mathcal{D}_i(i_1)-\mathcal{D}_i(i_2)){\rm{d}}\sigma \text{ for }0\leq t < t_0, \\ &0\text{ for }t_0\leq t \leq T, \end{cases} $ |
where the function
$ \begin{align}\label{equa-duality1}&-\int_0^T \langle \mathcal{C}, {\partial _t} m^c\rangle{\rm{d}}\sigma = \int_0^{t_0} \langle \mathcal{C}, (\mathcal{D}_c(c_1)-\mathcal{D}_c(c_2))\rangle{\rm{d}}\sigma \geq C_\mathcal{D} \iint_{\Omega_{t_0}} |\mathcal{C}|^2{\rm{d}}\mathit{\boldsymbol{x}}\;{\rm{d}}\sigma , \\ \nonumber&\iint_{\Omega_t} (D_c(c_1)\nabla c_1- D_c(c_2)\nabla c_2)\cdot \nabla m^c {\rm{d}}\mathit{\boldsymbol{x}}\;{\rm{d}}\sigma \\ \nonumber & = \frac{1}{2} \int_{\Omega} | \nabla \int_0^{t_0}(\mathcal{D}_c(c_1)-\mathcal{D}_c(c_2)) {\rm{d}}\sigma |^2 {\rm{d}}\mathit{\boldsymbol{x}}, \\ \nonumber&\int_{\Omega}\mathcal{C}_0(\mathit{\boldsymbol{x}})m^c(\mathit{\boldsymbol{x}}, 0){\rm{d}}\mathit{\boldsymbol{x}} = \int_{t}^{t_0}\!\!\int_{\Omega}\mathcal{C}_0(\mathit{\boldsymbol{x}})(\mathcal{D}_c(c_1)-\mathcal{D}_c(c_2)){\rm{d}}\mathit{\boldsymbol{x}}\;{\rm{d}}\sigma \\ \nonumber&\;\;\;\;\;\;\;\;{ \leq C_{16} \int_{\Omega}|\mathcal{C}_0(\mathit{\boldsymbol{x}})|^2{\rm{d}}\mathit{\boldsymbol{x}}+\frac{C_\mathcal{D}}{4} \iint_{\Omega_{t_0}} |\mathcal{C}|^2{\rm{d}}\mathit{\boldsymbol{x}}\;{\rm{d}}\sigma , }\end{align} $ | (4.5) |
for some constants
$\begin{split} \iint_{\Omega_t} c_1{\mathcal{U}}\cdot \nabla m^c {\rm{d}}\mathit{\boldsymbol{x}}\;{\rm{d}}\sigma & \leq C_{17} \int_0^{t_0}\iint_{\Omega_t}\left| {\mathcal{U}} \right|^2{\rm{d}}\mathit{\boldsymbol{x}}\;{\rm{d}}\sigma\; {\rm{d}}t \\ & +\frac{1}{8} \int_{\Omega} |\nabla \int_0^{t_0}(\mathcal{D}_c(c_1)-\mathcal{D}_c(c_2)) {\rm{d}}\sigma |^2 {\rm{d}}\mathit{\boldsymbol{x}}, \\ \iint_{\Omega_t} \mathcal{C}\mathit{\boldsymbol{u}}_2\cdot \nabla m^c {\rm{d}}\mathit{\boldsymbol{x}}\;{\rm{d}}\sigma &\leq C_{18} \int_0^{t_0} \iint_{\Omega_t}\left| {\mathcal{C}} \right|^2{\rm{d}}\mathit{\boldsymbol{x}}\;{\rm{d}}\sigma \; {\rm{d}}t \\ & +\frac{1}{8} \int_{\Omega} |\nabla \int_0^{t_0}(\mathcal{D}_c(c_1)-\mathcal{D}_c(c_2)) {\rm{d}}\sigma |^2 {\rm{d}}\mathit{\boldsymbol{x}}, \end{split}$ | (4.6) |
for some constants
$\begin{split} \iint_{\Omega_t} (G_{s}(c_1, s_1)-G_{s}(c_2, s_2)) m^s {\rm{d}}\mathit{\boldsymbol{x}}\;{\rm{d}}\sigma &\leq C_{19}\int_{0}^{t_0} \Bigl(\iint_{\Omega_t}(\left| {\mathcal{C}} \right|^2+|\mathcal{S}|^2){\rm{d}}\mathit{\boldsymbol{x}}\;{\rm{d}}\sigma \Bigl){\rm{d}}t\\ & +\frac{C_\mathcal{D}}{4} \iint_{\Omega_{t_0}} |\mathcal{C}|^2{\rm{d}}\mathit{\boldsymbol{x}}\;{\rm{d}}\sigma , \end{split}$ | (4.7) |
for some constant
$\begin{split} \frac{C_\mathcal{D}}{2} \iint_{\Omega_{t_0}} |\mathcal{C}|^2{\rm{d}}\mathit{\boldsymbol{x}}\;{\rm{d}}\sigma &\leq C_{16} \int_{\Omega}|\mathcal{C}_0(\mathit{\boldsymbol{x}})|^2{\rm{d}}\mathit{\boldsymbol{x}} \\ & +C_{20}\Biggl(\iint_{\Omega_t}|\mathit{\boldsymbol{g}}_1-\mathit{\boldsymbol{g}}_2|^2{\rm{d}}\mathit{\boldsymbol{x}}\;{\rm{d}}\sigma +\iint_{\Omega_t}|\mathit{\boldsymbol{f}}_1-\mathit{\boldsymbol{f}}_2|^2{\rm{d}}\mathit{\boldsymbol{x}}\;{\rm{d}}\sigma \Biggl) \\ &+C_{21}\int_{0}^{t_0} \Bigl(\iint_{\Omega_t}(\left| {\mathcal{C}} \right|^2+|\mathcal{S}|^2){\rm{d}}\mathit{\boldsymbol{x}}{\rm{d}}\sigma\Bigl){\rm{d}}t, \end{split}$ | (4.8) |
for some constants
$\begin{split} \frac{C_\mathcal{D}}{2} \iint_{\Omega_{t_0}} |\mathcal{S}|^2{\rm{d}}\mathit{\boldsymbol{x}}\;{\rm{d}}\sigma &\leq C_{22} \int_{\Omega}|\mathcal{S}_0(\mathit{\boldsymbol{x}})|^2{\rm{d}}\mathit{\boldsymbol{x}}\\ & +C_{23}\Biggl(\iint_{\Omega_t} |\mathit{\boldsymbol{g}}_1-\mathit{\boldsymbol{g}}_2|^2{\rm{d}}\mathit{\boldsymbol{x}}{\rm{d}}\sigma+\iint_{\Omega_t}|\mathit{\boldsymbol{f}}_1-\mathit{\boldsymbol{f}}_2|^2{\rm{d}}\mathit{\boldsymbol{x}}{\rm{d}}\sigma\Biggl) \\ &+C_{24}\int_{0}^{t_0} \Bigl(\iint_{\Omega_t}(\left| {\mathcal{C}} \right|^2+|\mathcal{S}|^2){\rm{d}}\mathit{\boldsymbol{x}}{\rm{d}}\sigma\Bigl){\rm{d}}t, \end{split}$ | (4.9) |
for some constants
$\begin{split} \iint_{\Omega_{t_0}} (|\mathcal{C}|^2+|\mathcal{S}|^2){\rm{d}}\mathit{\boldsymbol{x}}\;{\rm{d}}\sigma &\leq C_{25} \int_{\Omega}\Bigl(|\mathcal{C}_0(\mathit{\boldsymbol{x}})|^2+|\mathcal{S}_0(\mathit{\boldsymbol{x}})|^2\Bigl){\rm{d}}\mathit{\boldsymbol{x}} \\ & +C_{26}\int_{0}^{t_0} \Bigl(\iint_{\Omega_t}(\left| {\mathcal{C}} \right|^2+|\mathcal{S}|^2){\rm{d}}\mathit{\boldsymbol{x}}{\rm{d}}\sigma\Bigl){\rm{d}}t\\ & +C_{27}\Biggl(\iint_{\Omega_t}|\mathit{\boldsymbol{g}}_1-\mathit{\boldsymbol{g}}_2|^2{\rm{d}}\mathit{\boldsymbol{x}}{\rm{d}}\sigma+\!\!\iint_{\Omega_t}|\mathit{\boldsymbol{f}}_1-\mathit{\boldsymbol{f}}_2|^2{\rm{d}}\mathit{\boldsymbol{x}}{\rm{d}}\sigma\Biggl), \end{split}$ | (4.10) |
for some constants
In this section, we construct a primal-mixed fully-discrete scheme for the approximation of the coupled system (2.1). The finite element spaces characterising the semi-discrete problem are then specified, and a proof of convergence to the unique weak solution (presented in Definition 2.1) is outlined.
Let
$ \begin{align}\nonumber {\rm{V}}_h&\! = \! \bigl\{m_{h} \in {\rm{H}}^1(\Omega): m_{h}|_{K} \in \mathcal{P}_{1}(K)\, \forall K \in \mathcal{T}_h\bigr\}, \\ \nonumber {\boldsymbol{\rm{H}}}_h &\! = \! \bigl\{\mathit{\boldsymbol{v}}_h \in {{\boldsymbol{\rm{H}}}_0({\rm{div}};\Omega)}: \mathit{\boldsymbol{v}}_h|_{K} \in {\boldsymbol{\rm{RT}}}_{0}(K)\, \forall K\in \mathcal{T}_h \bigr\}, \\ \nonumber {\boldsymbol{\rm{Z}}}_h&\! = \! \bigl\{\mathit{\boldsymbol{z}}_h \in {\boldsymbol{\rm{H}}}({\boldsymbol{\rm{curl}}}, \Omega): \mathit{\boldsymbol{z}}_h|_{K} \in {\boldsymbol{\rm{ND}}}_1(K)\, \forall K \in \mathcal{T}_h \bigr\}, \\ {\rm{Q}}_{h} \!& = \! \bigl\{q_{h} \in {\rm{L}}_0^{2}(\Omega): q_{h}|_{K} \in \mathcal{P}_{0}(K)\, \forall K \in \mathcal{T}_h\bigr\}, \label{eq:spaces-h} \end{align} $ | (5.1) |
where for any
Then, a Galerkin semi-discretisation associated to the formulation introduced in Definition 2.1 reads: For
$\begin{split} &\int_\Omega {\partial _t} c_h(t) m_h^c{\rm{d}}\mathit{\boldsymbol{x}} + \int_\Omega (D_c(c_h(t))\nabla c_h(t) -c_h(t)\mathit{\boldsymbol{u}}_h)\cdot \nabla m_h^c{\rm{d}}\mathit{\boldsymbol{x}} \\ & \;\;\;\;= \int_\Omega G_c(c_h(t), s_h(t)) m_h^c{\rm{d}}\mathit{\boldsymbol{x}}, \\ &\int_\Omega {\partial _t} s_h(t) m_h^s{\rm{d}}\mathit{\boldsymbol{x}} + \int_\Omega(D_s(s_h(t))\nabla s_h(t) -s_h(t)\mathit{\boldsymbol{u}}_h)\cdot \nabla m_h^s{\rm{d}}\mathit{\boldsymbol{x}} \\ &\;\;\;\; = \int_\Omega G_s(c_h(t), s_h(t)) m_h^s{\rm{d}}\mathit{\boldsymbol{x}}, \\ &\int_\Omega \mathbb{K}^{-1}\mathit{\boldsymbol{u}}_h(t)\cdot\mathit{\boldsymbol{v}}_h\;{\rm{d}}\mathit{\boldsymbol{x}}+ \sqrt{\mu}\int_\Omega {\boldsymbol{\rm{curl}}}\;\mathit{\boldsymbol{v}}_h\cdot \mathit{\boldsymbol{\omega}}_h(t){\rm{d}}\mathit{\boldsymbol{x}} -\int_\Omega p_h(t)\, \;{\rm{div}}\;\mathit{\boldsymbol{v}}_h\;{\rm{d}}\mathit{\boldsymbol{x}} \\ & \;\;\;\;= \int_\Omega (s_h(t)\mathit{\boldsymbol{g}}+\mathit{\boldsymbol{f}})\cdot \mathit{\boldsymbol{v}}_h\;{\rm{d}}\mathit{\boldsymbol{x}}, \\ &\sqrt{\mu}\int_\Omega {\boldsymbol{\rm{curl}}}\;\mathit{\boldsymbol{u}}_h(t)\cdot \mathit{\boldsymbol{z}}_h\;{\rm{d}}\mathit{\boldsymbol{x}}- \int_\Omega\mathit{\boldsymbol{\omega}}_h(t)\cdot\mathit{\boldsymbol{z}}_h\;{\rm{d}}\mathit{\boldsymbol{x}} = 0, \\ & -\int_\Omega q_h\, {\rm{div}}\mathit{\boldsymbol{u}}_h(t) \;{\rm{d}}\mathit{\boldsymbol{x}} = 0, \end{split}$ | (5.2) |
for all
Let
$ \label{finite-element-scheme} \begin{split} \int_\Omega \frac{c_{h}^n\!-\!c_{h}^{n-1}}{\Delta t}\, m_h^c{\rm{d}}\mathit{\boldsymbol{x}} + \int_{\Omega} (D_c(c_h^n)\nabla c_h^n -c_h^n\mathit{\boldsymbol{u}}^n_h)\cdot \nabla m_h^c {\rm{d}}\mathit{\boldsymbol{x}} \\ = \int_{\Omega} G_c(c_h^{n-1}, s_h^{n-1}) m_h^c {\rm{d}}\mathit{\boldsymbol{x}}, \\ \int_\Omega \frac{s_{h}^n\!-\!s_{h}^{n-1}}{\Delta t}\, m_h^c{\rm{d}}\mathit{\boldsymbol{x}} + \int_{\Omega} (D_s(s_h^n)\nabla s_h^n -s_h^n\mathit{\boldsymbol{u}}^n_h)\cdot \nabla m_h^s {\rm{d}}\mathit{\boldsymbol{x}} \\ = \int_{\Omega} G_s(c_h^{n-1}, s_h^{n-1}) m_h^s {\rm{d}}\mathit{\boldsymbol{x}}, \\ \int_{\Omega}\mathbb{K}^{-1}\mathit{\boldsymbol{u}}^n_h\cdot\mathit{\boldsymbol{v}}_h {\rm{d}}\mathit{\boldsymbol{x}} +\sqrt{\mu}\int_{\Omega}\mathit{\boldsymbol{v}}_h\cdot{\boldsymbol{\rm{curl}}}\mathit{\boldsymbol{\omega}}^n_h {\rm{d}}\mathit{\boldsymbol{x}} -\int_{\Omega}p^n_h\;{\rm{div}}\;\mathit{\boldsymbol{v}}_h\; {\rm{d}}\mathit{\boldsymbol{x}} \\ = \int_{\Omega} (s_h^{n}\mathit{\boldsymbol{g}}+\mathit{\boldsymbol{f}})\cdot \mathit{\boldsymbol{v}}_h {\rm{d}}\mathit{\boldsymbol{x}}, \\ \sqrt{\mu}\int_{\Omega}\mathit{\boldsymbol{u}}^n_h\cdot{\boldsymbol{\rm{curl}}}\mathit{\boldsymbol{z}}_h {\rm{d}}\mathit{\boldsymbol{x}}\;{\rm{d}}t - \int_{\Omega}\mathit{\boldsymbol{\omega}}^n_h\cdot\mathit{\boldsymbol{z}}_h{\rm{d}}\mathit{\boldsymbol{x}} = 0, \\ -\int_{\Omega} q_h\;{\rm{div}}\;\mathit{\boldsymbol{u}}^n_h\; {\rm{d}}\mathit{\boldsymbol{x}} = 0, \end{split} $ | (5.3) |
for all
Note that the positive and negative cuts test functions may not be admissible in finite element context. Therefore, we cannot obtain the maximum principle as in the continuous case. Comparing to the continuous case, here we assume that the initial condition
In order to derive stability estimates, we employ the same kind of test functions in the same combination of equations used in Section 3; whereas the chain rule for time derivatives is now replaced by the convexity inequality
$ a^{n}(a^{n}-a^{n-1})\geq \frac{|a^n|^2}{2}-\frac{|a^{n-1}|^2}{2}, $ |
to treat the finite difference discretisation of the time derivatives. Proceeding in this way, we can obtain estimates (uniform in both
$ \Vert\mathit{\boldsymbol{u}}_h\Vert_{{\boldsymbol{\rm{L}}}^2(0, T;{\boldsymbol{\rm{H}}}({\rm{div}};\Omega))}+\|\mathit{\boldsymbol{\omega}}_h\|_{{\boldsymbol{\rm{L}}}^2(0, T;{\boldsymbol{\rm{H}}}({\boldsymbol{\rm{curl}}};\Omega))} +\|p_h\|_{{\rm{L}}^2(0, T;{\rm{L}}_0^2(\Omega))} \leq C, $ | (5.4) |
and for the advection-reaction-diffusion system
$ \label{est1-u-weak-discrete} \begin{array}{l}\displaystyle ||c_h||_{{\rm{L}}^\infty(0, T;{\rm{L}}^2(\Omega))} +||s_h||_{{\rm{L}}^\infty(0, T;{\rm{L}}^2(\Omega))} +||\nabla c_h||_{{\rm{L}}^2(\Omega_T)} +||\nabla s_h||_{{\rm{L}}^2(\Omega_T)}\le C, \end{array} $ | (5.5) |
for some constant
The next goal is to establish the relative compactness in
Lemma 5.1. There exists a constant
$ \iint_{\Omega_{\mathit{\boldsymbol{r}}} \times (0, T)}\bigl[|c_{h}(\mathit{\boldsymbol{x}}+\mathit{\boldsymbol{r}}, t)-c_{h}(\mathit{\boldsymbol{x}}, t)|^2+|s_{h}(\mathit{\boldsymbol{x}}+\mathit{\boldsymbol{r}}, t)-s_{h}(\mathit{\boldsymbol{x}}, t)|^2\bigr]{\rm{d}}\mathit{\boldsymbol{x}}\;{\rm{d}}t \nonumber \\\;\;\;\;\le \mathcal{C}\, |\mathit{\boldsymbol{r}}|^2, $ | (5.6) |
$ \iint_{\Omega \times (0, T-\tau)} \bigl[|c_{h}(\mathit{\boldsymbol{x}}, t+\tau)-c_{h}(\mathit{\boldsymbol{x}}, t)|^2+|s_{h}(\mathit{\boldsymbol{x}}, t+\tau)-s_{h}(\mathit{\boldsymbol{x}}, t)|^2\bigr]{\rm{d}}\mathit{\boldsymbol{x}}\;{\rm{d}}t \nonumber \\\;\;\;\;\le C(\tau+\Delta t). $ | (5.7) |
for all
Proof. Let us introduce the space translates
$(J^{\mathit{\boldsymbol{r}}} c_h)(\mathit{\boldsymbol{x}}, \cdot) = c_h(\mathit{\boldsymbol{x}}+\mathit{\boldsymbol{r}}, \cdot)-c_h(\mathit{\boldsymbol{x}}, \cdot), \;\;\;\;\text{and} \;\;\;\; (J^{\mathit{\boldsymbol{r}}} s_h)(\mathit{\boldsymbol{x}}, \cdot) = s_h(\mathit{\boldsymbol{x}}+\mathit{\boldsymbol{r}}, \cdot)-s_h(\mathit{\boldsymbol{x}}, \cdot).$ |
From the
$ \int_0^T\!\!\int_{\Omega_{\mathit{\boldsymbol{r}}}} |J^{\mathit{\boldsymbol{r}}} c_{h}|^2{\rm{d}}\mathit{\boldsymbol{x}}\;{\rm{d}}t +\int_0^T\!\!\int_{\Omega_{\mathit{\boldsymbol{r}}}} |J^{\mathit{\boldsymbol{r}}} s_{h}|^2{\rm{d}}\mathit{\boldsymbol{x}}\;{\rm{d}}t\leq \mathcal{C} |\mathit{\boldsymbol{r}}|^2, $ | (5.8) |
easily follows. It is then clear that the right-hand side in (5.8) vanishes as
Next we introduce the time translates
$ (T^h c_{h})(\cdot, t): = c_{h}(\cdot, t+\tau)-c_{h}(\cdot, t) \;{\rm{and}}\;(T^h s_{h})(\cdot, t): = s_{h}(\cdot, t+\tau)-s_{h}(\cdot, t), $ |
and notice that for all
Let us now define
$\begin{split} \int_{\Omega} {\partial _t} \overline{c}_h \, m_h^c{\rm{d}}\mathit{\boldsymbol{x}} + \int_{\Omega} (D_c(c_h)\nabla c_h- c_h\mathit{\boldsymbol{u}}_h) \cdot\nabla m_h^c {\rm{d}}\mathit{\boldsymbol{x}}& = \int_{\Omega} G_c(c_h, s_h) m_h^c {\rm{d}}\mathit{\boldsymbol{x}}, \\ \int_{\Omega} {\partial _t} \overline{s}_h \, m_h^s{\rm{d}}\mathit{\boldsymbol{x}} + \int_{\Omega} (D_s(s_h)\nabla s_h- s_h\mathit{\boldsymbol{u}}_h) \cdot \nabla m_h^s {\rm{d}}\mathit{\boldsymbol{x}}& = \int_{\Omega} G_s(c_h, s_h) m_h^s {\rm{d}}\mathit{\boldsymbol{x}}, \end{split}$ | (5.9) |
for all
$\begin{align*} &\int_0^{T-\tau} \int_{\Omega} |(T^h \bar c_{h})(\mathit{\boldsymbol{x}}, t)|^2 {\rm{d}}\mathit{\boldsymbol{x}} {\rm{d}}t +\int_0^{T-\tau} \int_{\Omega} |(T^h \bar s_{h})(\mathit{\boldsymbol{x}}, t)|^2 {\rm{d}}\mathit{\boldsymbol{x}}\;{\rm{d}}t\\ &\;\;\;\; = \int_0^{T-\tau} \!\!\!\int_{\Omega} \Bigl( \int_{t}^{t+ \tau} \!\!\! \partial_\sigma \bar c_{h} (\mathit{\boldsymbol{x}}, \sigma){\rm{d}}\sigma \! \Bigl)(T^h c_{h})(\mathit{\boldsymbol{x}}, t) {\rm{d}}\mathit{\boldsymbol{x}}\;{\rm{d}}t \\ & \;\;\;\;+ \! \int_0^{T-\tau}\!\!\! \int_{\Omega} \Bigl( \int_{t}^{t+ \tau}\!\!\! \partial_\sigma \bar s_{h} (\mathit{\boldsymbol{x}}, \sigma){\rm{d}}\sigma \! \Bigl)(T^h s_{h})(\mathit{\boldsymbol{x}}, t) {\rm{d}}\mathit{\boldsymbol{x}}\;{\rm{d}}t\\ &\;\;\;\; = -\int_0^{T-\tau} \int_{\Omega} \int_{t}^{t+ \tau} (D_c(c_h)\nabla c_{h}(\mathit{\boldsymbol{x}}, \sigma)-c_h(\mathit{\boldsymbol{x}}, t))\mathit{\boldsymbol{u}}_h(\mathit{\boldsymbol{x}}))\cdot \nabla (T^h c_{h})(\mathit{\boldsymbol{x}}, t){\rm{d}}\mathit{\boldsymbol{x}}\;{\rm{d}}\sigma \;{\rm{d}}t\\ & \;\;\;\;-\int_0^{T-\tau} \int_{\Omega} \int_{t}^{t+ \tau} (D_s(s_h)\nabla s_{h}(\mathit{\boldsymbol{x}}, \sigma)-s_h(\mathit{\boldsymbol{x}}, t))\mathit{\boldsymbol{u}}_h(\mathit{\boldsymbol{x}}))\cdot \nabla (T^h s_{h})(\mathit{\boldsymbol{x}}, t){\rm{d}}\mathit{\boldsymbol{x}}\;{\rm{d}}\sigma\; {\rm{d}}t \\ & \;\;\;\; +\int_0^{T-\tau} \int_{\Omega} \int_{t}^{t+ \tau} G_c(c_h(\mathit{\boldsymbol{x}}, t), s_h(\mathit{\boldsymbol{x}}, t)) (T^h c_{h})(\mathit{\boldsymbol{x}}, t){\rm{d}}\mathit{\boldsymbol{x}}\;{\rm{d}}\sigma\;{\rm{d}}t\\ & \;\;\;\;+\int_0^{T-\tau} \int_{\Omega} \int_{t}^{t+ \tau} G_s(c_h(\mathit{\boldsymbol{x}}, t), s_h(\mathit{\boldsymbol{x}}, t)) (T^h s_{h})(\mathit{\boldsymbol{x}}, t){\rm{d}}\mathit{\boldsymbol{x}}\;{\rm{d}}\sigma\;{\rm{d}}t\\ & = :I_1+I_2+I_3+I_4. \end{align*}$ |
Now, we examine these integrals separately. For
$\begin{align*} & |I_1| \leq C \Biggl[\int_0^{T-\tau} \! \! \int_{\Omega} \Bigl(\int_{t}^{t+\tau} |\nabla c_{h}(\mathit{\boldsymbol{x}}, \sigma)|^2{\rm{d}}\sigma \Bigl)^{2} {\rm{d}}\mathit{\boldsymbol{x}}\;{\rm{d}}t\Biggl]^{\frac{1}{2}} \\ &\;\;\;\;\;\;\;\; \times \Biggl[\int_0^{T-\tau} \!\! \int_{\Omega} |\nabla (T^ h c_{h})(\mathit{\boldsymbol{x}}, t)|^2 {\rm{d}}\mathit{\boldsymbol{x}}\;{\rm{d}}t \Biggl]^{\frac{1}{2}} \leq C\, \tau, \end{align*}$ |
and similarly
$ \int_0^{T-\tau}\!\!\int_{\Omega} \Bigl( |\bar c_h(\cdot, t+\tau)-\bar c_h(\cdot, t)|^2 + |\bar s_h(\cdot, t+\tau)-\bar s_h(\cdot, t)|^2 \Bigr){\rm{d}}\mathit{\boldsymbol{x}}\;{\rm{d}}t \leq \mathcal C\, \tau. $ |
Furthermore, the definition of (
$\begin{align*} \|\bar c_h - c_h\|^2_{{\rm{L}}^2(\Omega_{T})}\leq \sum\limits_{n = 1}^N \Delta t \|c_h^{n}-c_h^{n-1}\|^2_{{\rm{L}}^2(\Omega)}\leq \, \mathcal{C}(\Delta t)\ &\to 0 \;\;\text{as}\;\;\Delta t \to 0, \\ \|\bar s_h -s_h\|^2_{{\rm{L}}^2(\Omega_{T})}\leq \sum_{n = 1}^N \Delta t \|s_h^{n}-s_h^{n-1}\|^2_{{\rm{L}}^2(\Omega)}\leq \, \mathcal{C}(\Delta t)\ & \to 0 \; \; \text{as}\;\;\Delta t \to 0, \end{align*}$ |
which establishes (5.7), therefore finishing the proof.
Note that as a consequence of (5.4)-(5.5), Lemma 5.1, and Kolmogorov's compactness criterion (cf. [10,Theorem Ⅳ.25]), we can assert that there exists a subsequence of
$\begin{align*} &(c_h, s_h) \to (c, s)\;{\rm{strongly\;in}}\;{\rm{L}}^2(\Omega_T, \mathbb{R}^2)\; {\rm{and\;in}}\;{\rm{L}}^2(0, T;{\rm{H}}^1(\Omega, \mathbb{R}^2))\;{\rm{weakly}}, \\ &(\mathit{\boldsymbol{u}}_h, \mathit{\boldsymbol{\omega}}_h, p_h) \to (\mathit{\boldsymbol{u}}, \mathit{\boldsymbol{\omega}}, p) \\ & \;\;\;\;\;\;{\rm{in}}\; {\boldsymbol{\rm{L}}}^2(0, T;{\boldsymbol{\rm{H}}}({\rm{div}};\Omega))\times{\boldsymbol{\rm{L}}}^2(0, T;{\boldsymbol{\rm{H}}}({\boldsymbol{\rm{curl}}};\Omega))\times {\rm{L}}^2(0, T;{\rm{L}}_0^2(\Omega))\;{\rm{weakly}}. \end{align*}$ |
These convergences allow us to identify the limit
Theorem 5.2. Assume that conditions (2.3), (2.4) and (2.5) hold. If
We observe that the error generated by the fully discrete scheme (5.3) has two components: one due to the spatial discretisation and depending on the meshsize
$ \label{convergence-rate} \Vert(c(\cdot, t_n), s(\cdot, t_n), \mathit{\boldsymbol{u}}(\cdot, t_n), \mathit{\boldsymbol{\omega}}(\cdot, t_n), p(\cdot, t_n)) -(c^n_h, s^n_h, \mathit{\boldsymbol{u}}^n_h, \mathit{\boldsymbol{\omega}}^n_h, p^n_h)\Vert\le C_{1}h+C_{2}\Delta t, $ | (5.10) |
with
We finally present a set of elementary examples to illustrate the properties of the model and of the proposed finite element method. The coupling between Brinkman and reaction-diffusion equations will be implemented either via: a) a fully monolithic solution, b) an iterative Picard method splitting linear Brinkman and fully explicit reaction-diffusion equations, and c) an iterative Picard method with an embedded Newton algorithm for the linearisation of the reaction-diffusion system. Rigorous convergence proofs for the Newton iteration and various linear iterative schemes applied to similar problems in non-viscous flow in porous media can be found in e.g. [27,34]. For our particular scenario, efficient coupling strategies and thorough comparisons between dedicated solvers have been reported in [24].
Example 1. Spatio-temporal accuracy. Let us consider the spatio-temporal domain
$\begin{split} c(\mathit{\boldsymbol{x}}, t)& = \exp(-4D_c\pi^2 t) [\cos(2\pi x)+ \cos (2\pi y)], \\ s(\mathit{\boldsymbol{x}}, t)& = \exp(-4D_s\pi^2 t) [\cos(2\pi x)+ \cos (2\pi y)], p(\mathit{\boldsymbol{x}}, t) = x^4-y^4, \\ \mathit{\boldsymbol{u}}(\mathit{\boldsymbol{x}}, t)& = 3\sin(t) x^2(x-1)^2y^2(y-1)^2 \bigl(x(x-1)(2y-1), -(2x-1)y(y-1)\bigr)^{T}, \\ \mathit{\boldsymbol{\omega}}(\mathit{\boldsymbol{x}}, t)& = -6\sin(t)\sqrt{\mu}[x(5x^3-10x^2+6x-1)y^3(y-1)^3\\ & +y(5y^3-10y^2+6y-1)x^3(x-1)^3]. \end{split}$ | (6.1) |
The model parameters are set as
$\begin{align*} G_j(c, s)& = 6\pi \sin(t) \exp(-4D_j\pi^2 t) x^2(x-1)^2y^2(y-1)^2\\ & \times [(2x-1)y(y-1)\sin(2\pi y)-x(x-1)(2y-1)\sin(2\pi x)],\;\;\;\; j = c, s, \end{align*}$ |
whereas
DoF | 163 | 579 | 2179 | 8451 | 33283 | 526339 |
avg(iter) | 4.2 | 4.0 | 4.0 | 3.8 | 4.2 | 4.3 |
Example 1. Average number of Newton iterations to reach the residual tolerance tol
Errors associated to splitting algorithms are avoided by invoking the exact full Jacobian and performing Newton iterations until convergence at each time step. The mesh convergence is investigated by fixing
Example 2. Bacterial bioconvection. Next we focus on the interaction between bacteria (with concentration
Example 2: snapshots at
Example 3. FitzHugh-Nagumo dynamics in a porous cavity. This test emphasises the effects of high contrast permeabilities on the propagation of travelling waves dictated by the well-known FitzHugh-Nagumo reactions
We also perform the 3D counterpart of this example, defined in the polygonal domain
Example 4. Intracellular calcium-induced calcium release. In closing this section we present a simulation of the interaction between two species
$\begin{align*} & D_c = (1 + \frac{1}{4}\eta(\mathit{\boldsymbol{x}}))\begin{pmatrix} 0.6&0 \\ 0&0.3\end{pmatrix}, \ {D_s = 0.01}, \\ G_c (c, s)& = \nu_1-\nu_2 \frac{c^2}{k_2+c^2} +\nu_3 \frac{c^4 s^2}{(k_3+s^2)(k_4+c^4)} -\nu_4 c, \\ & G_s (c, s) = \nu_2 \frac{c^2}{k_2+c^2} -\nu_3 \frac{c^4 s^2}{(k_3+s^2)(k_4+c^4)} -\nu_5 s, \end{align*}$ |
where
The interaction of viscous flow in porous media and reaction-diffusion phenomena occur in a variety of applicative scenarios. Here we have advocated the solvability analysis and the numerical approximation of a special class of problems related to Brinkman and nonlinear advection-diffusion-reaction equations. The analysis of the coupled set of equations is framed under the theory of fixed-point schemes combined with an abstract result for saddle-point problems. A mixed-primal finite element method has been proposed and we have discussed its stability and convergence properties. We have provided numerical validation of the spatio-temporal convergence of the method, and have also addressed the simulation of a number of biochemically-oriented examples.
Our theoretical framework currently does not account for higher order terms nor the chemotactical convection simulated in Example 2, and further investigation is needed in this direction. We also need to cover the dependence of the momentum force on the gradient of the chemical concentrations (studied numerically in Example 4). Regarding other model generalisations, interesting topics be explored include cross-diffusion effects, nonlinear convective fluxes, a non-laminar flow regime, viscosity depending on the chemical concentrations, or variable density flows. Finally, we mention that the study of mixed formulations involving embedded saddle-point problems and the convergence analysis of the associated fully-mixed finite element schemes can be the subject of suitable extensions to the mathematical aspects of this contribution.
We thank the inputs from two anonymous referees, which have resulted in a number of improvements with respect to the initial version of the manuscript. Also, special thanks are due to Davin Lunz, Tori Pereira, Lindon Roberts, Caoimhe Rooney, Ian Roper, Valentin Sulzer, and Jessica Williams (cohort two of the Oxford Centre for Doctoral Training on Industrially Focused Mathematical Modelling) for the design and preliminary implementation of Example 2 (bacterial bioconvection on a disk). This work was partially supported by CONICYT-Chile through FONDECYT projects 11160706 and 1140791; by DIUBB through project 165608--3/R; and by the EPSRC through the Research Grant EP/R00207X/1.
1. | David Mora, Alberth Silgado, 2022, Chapter 8, 978-3-030-95318-8, 321, 10.1007/978-3-030-95319-5_8 | |
2. | Nitesh Verma, Sarvesh Kumar, VIRTUAL ELEMENT APPROXIMATIONS FOR TWO SPECIES MODEL OF THE ADVECTION-DIFFUSION-REACTION IN POROELASTIC MEDIA, 2022, 27, 1392-6292, 668, 10.3846/mma.2022.15542 | |
3. | Mario Alvarez, Gabriel N. Gatica, Ricardo Ruiz-Baier, A posteriori error estimation for an augmented mixed-primal method applied to sedimentation–consolidation systems, 2018, 367, 00219991, 322, 10.1016/j.jcp.2018.04.040 | |
4. | Mostafa Bendahmane, Fahd Karami, Mohamed Zagour, Kinetic-fluid derivation and mathematical analysis of the cross-diffusion-Brinkman system, 2018, 41, 01704214, 6288, 10.1002/mma.5139 | |
5. | N. A. Barnafi, B. Gómez-Vargas, W. J. Lourenço, R. F. Reis, B. M. Rocha, M. Lobosco, R. Ruiz-Baier, R. Weber dos Santos, Finite Element Methods for Large-Strain Poroelasticity/Chemotaxis Models Simulating the Formation of Myocardial Oedema, 2022, 92, 0885-7474, 10.1007/s10915-022-01944-2 | |
6. | Mostafa Bendahmane, Elmahdi Erraji, Fahd Karami, A 3D reaction–diffusion system describing calcium dynamics in cardiac cell, 2019, 14, 0973-5348, 205, 10.1051/mmnp/2018064 | |
7. | Nitesh Verma, Bryan Gómez-Vargas, Luis Miguel De Oliveira Vilaca, Sarvesh Kumar, Ricardo Ruiz-Baier, Well-posedness and discrete analysis for advection-diffusion-reaction in poroelastic media, 2022, 101, 0003-6811, 4914, 10.1080/00036811.2021.1877677 | |
8. | V. Anaya, A. Bouharguane, D. Mora, C. Reales, R. Ruiz-Baier, N. Seloula, H. Torres, Analysis and Approximation of a Vorticity–Velocity–Pressure Formulation for the Oseen Equations, 2019, 80, 0885-7474, 1577, 10.1007/s10915-019-00990-7 | |
9. | Raimund Bürger, Paul E. Méndez, Ricardo Ruiz-Baier, On $H(div)$-conforming Methods for Double-diffusion Equations in Porous Media, 2019, 57, 0036-1429, 1318, 10.1137/18M1196108 | |
10. | Huadong Gao, Wen Xie, A new optimal error analysis of a mixed finite element method for advection–diffusion–reaction Brinkman flow, 2024, 40, 0749-159X, 10.1002/num.23097 | |
11. | Raimund Bürger, Arbaz Khan, Paul E Méndez, Ricardo Ruiz-Baier, Divergence-conforming methods for transient double-diffusive flows: a priori and a posteriori error analysis, 2023, 0272-4979, 10.1093/imanum/drad090 | |
12. | Ruben Caraballo, Chansophea Wathanak In, Alberto F. Martín, Ricardo Ruiz‐Baier, Robust finite element methods and solvers for the Biot–Brinkman equations in vorticity form, 2024, 40, 0749-159X, 10.1002/num.23083 | |
13. | Raimund Bürger, Arbaz Khan, Paul E. Méndez, Ricardo Ruiz‐Baier, A divergence‐conforming method for flow and double‐diffusive transport, 2024, 1617-7061, 10.1002/pamm.202400201 |
DoF | 163 | 579 | 2179 | 8451 | 33283 | 526339 |
avg(iter) | 4.2 | 4.0 | 4.0 | 3.8 | 4.2 | 4.3 |
Example 1. Convergence tests for the spatial (left) and temporal (right) discretisation via mixed
Example 2: snapshots at
Example 3A: snapshots of FitzHugh-Nagumo dynamics on a porous mixture at early (left) and advanced (right) times. Computed solutions from top to bottom: membrane voltage, vorticity, and velocity.
Example 3A: Number of inner Newton steps and outer Picard steps needed to reach residual convergence to a tolerance of 1e-6.
Approximate membrane voltage, velocity, and pressure for the FitzHugh-Nagumo dynamics on a porous mixture at early (top), moderate (middle row), and advanced (bottom panels) times.
Example 4: Computed solutions (cytosolic calcium, sarcoplasmic calcium, vorticity, velocity, and pressure) for the intracellular calcium dynamics at early (left) and advanced (right) times.