
We formulate and investigate a nonlinear parabolic reaction–diffusion equation describing the oxygen concentration in encapsulated pancreatic cells with a general core-shell geometry. This geometry introduces a discontinuous diffusion coefficient as the material properties of the core and shell differ. We apply monotone operator theory to show the well-posedness of the problem in the strong form. Furthermore, the stationary solutions are unique and asymptotically stable. These results rely on the gradient structure of the underlying PDE. Our results provide necessary theoretical steps for validation of the model.
Citation: Thomas Geert de Jong, Georg Prokert, Alef Edou Sterk. Reaction–diffusion transport into core-shell geometry: Well-posedness and stability of stationary solutions[J]. Networks and Heterogeneous Media, 2025, 20(1): 1-14. doi: 10.3934/nhm.2025001
[1] | Avner Friedman . PDE problems arising in mathematical biology. Networks and Heterogeneous Media, 2012, 7(4): 691-703. doi: 10.3934/nhm.2012.7.691 |
[2] | 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 |
[3] | Tong Li . Qualitative analysis of some PDE models of traffic flow. Networks and Heterogeneous Media, 2013, 8(3): 773-781. doi: 10.3934/nhm.2013.8.773 |
[4] | Magali Tournus, Aurélie Edwards, Nicolas Seguin, Benoît Perthame . Analysis of a simplified model of the urine concentration mechanism. Networks and Heterogeneous Media, 2012, 7(4): 989-1018. doi: 10.3934/nhm.2012.7.989 |
[5] | 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 |
[6] | Steinar Evje, Aksel Hiorth, Merete V. Madland, Reidar I. Korsnes . A mathematical model relevant for weakening of chalk reservoirs due to chemical reactions. Networks and Heterogeneous Media, 2009, 4(4): 755-788. doi: 10.3934/nhm.2009.4.755 |
[7] | Ken-Ichi Nakamura, Toshiko Ogiwara . Periodically growing solutions in a class of strongly monotone semiflows. Networks and Heterogeneous Media, 2012, 7(4): 881-891. doi: 10.3934/nhm.2012.7.881 |
[8] | Steinar Evje, Kenneth H. Karlsen . Hyperbolic-elliptic models for well-reservoir flow. Networks and Heterogeneous Media, 2006, 1(4): 639-673. doi: 10.3934/nhm.2006.1.639 |
[9] | 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 |
[10] | 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 |
We formulate and investigate a nonlinear parabolic reaction–diffusion equation describing the oxygen concentration in encapsulated pancreatic cells with a general core-shell geometry. This geometry introduces a discontinuous diffusion coefficient as the material properties of the core and shell differ. We apply monotone operator theory to show the well-posedness of the problem in the strong form. Furthermore, the stationary solutions are unique and asymptotically stable. These results rely on the gradient structure of the underlying PDE. Our results provide necessary theoretical steps for validation of the model.
Type 1 diabetes is a chronic condition that is caused by the immune system mistakenly destroying insulin-producing pancreatic Langerhans islets. Treatment requires continuous monitoring and maintenance of insulin levels via external means such as injections or insulin pumps. Islet transplantation from deceased donors provides a new treatment to recover natural insulin production [1]. This is not a permanent solution, as a significant percentage of patients do not achieve insulin independence at 5 years [1]. Furthermore, such a treatment requires immunosuppressants. As there are only a few suitable donors, it is desirable to increase the durability of the transplanted cells. This could be achieved by protecting the transplanted cells with a physical barrier. Alginate is a promising candidate as it can selectively diffuse or block certain molecules [2]. In addition, alginates are relatively inert [3,4,5]. This novel treatment shows promising results in mammalian trials [6,7,8,9]. Human trials are still in the preliminary phase [10,11,12].
King et al. [13,14] proposed a model that describes the reaction–diffusion of oxygen through a protective shell encapsulating a core so that hypoxia can be avoided within the donor cells. The encapsulation and core of donor cells are approximately spherical [15]. Hence, in [13,14], a spherical core and shell with a common center are considered. In that paper the authors derived the governing ODE for stationary solutions and numerically computed these solutions with oxygen concentration above the hypoxia threshold. The existence of these stationary solutions is made rigorous in [16] using topological shooting [17,18,19]. The latter result considers parameters used in cell encapsulation experiments [15,20,21]. Stationary solutions of oxygen concentration in a core without a shell have been extensively studied [22,23,24,25].
The oxygen concentration in the core and shell will initially be away from the stationary state. Furthermore, the geometry of the cell is only approximately spherical. Hence, it is important for the validation of the model to determine the stability of stationary solutions for general core-shell geometry, which is the goal of this work. This first requires formulating the corresponding PDE and showing its well-posedness, which was not considered in [13,14]. The resulting PDE is of parabolic type. On the boundary of the outside shell, we assume the oxygen concentration to be constant. As the shell and the core have different diffusion coefficients, there is a discontinuity of this coefficient at the interface separating them. This discontinuity makes the corresponding stationary problem a diffraction problem [26]. We make the natural assumption that the concentration and the flux of oxygen are continuous at the interface. Oxygen is consumed only by the donor cells in the core but not by the protective shell. This leads to a nonlinear reaction–diffusion PDE in the core, where the non-linearity corresponds to Michaelis–Menten consumption and consequently is bounded and monotone. In the shell, however, the problem reduces to linear diffusion.
Our main results for our newly formulated PDE model are
- Well-posedness of the strong form,
- Uniqueness and asymptotic stability of stationary solutions.
The results are necessary theoretical steps in the validation of the biological model. Furthermore, we will see that all the results hold for more general kinetics described by Hill's equation. We note that similar results have been obtained in the one-dimensional setting for the internal transition layer, which arises by scaling the discontinuous diffusion coefficient by a small parameter [27].
The sum of the reaction term and the spatial differential term is a nonlinear monotone operator. Hence, for the well-posedness result we can apply classical monotone operator theory. We apply the well-known theorems of Komura and Browder–Minty [28]. We emphasize that the monotonicity of the nonlinear operator defining our evolution problem results from the underlying structure of the problem as a gradient flow with respect to a convex functional. This structure also enables us to show the uniqueness and asymptotic stability of the stationary solution.
In Section 2, we present the governing equations. The well-posedness result is proved in Section 3. The stability results are presented in Section 4. In Section 5, we formulate the PDE donor cell model, transform it to the setting considered in Section 2, and apply the theorems from Sections 3 and 4. Finally, in Section 6, we conclude with remarks and an outlook on further questions.
We start with a description of the geometry of the core and its protective shell. The domain is denoted by Ω⊂RN with ¯Ω compact and N≥2. For the application N=3, but the results hold for N≥2. Let Γ be an (N−1)-dim surface that divides Ω into two open domains Ω1,Ω2 such that ∂Ω1=Γ is a closed connected hypersurface and Ω2 has 2 boundary components, S:=∂Ω and Γ, see Figure 1. We take S,Γ of class C2. We refer to Ω1 as the core, Ω2 as the shell, and Ω as the core-shell.
We consider the following PDE:
dudt−bΔu=f(u)in(Ω∖Γ)×(0,T), | (2.1) |
u|S=0, | (2.2) |
[u]Γ=0, | (2.3) |
[b∇u⋅ν]Γ=0, | (2.4) |
u(⋅,0)=u0inΩ. | (2.5) |
In Eq (2.1), b:Ω→R is given by
b(x)={b1ifx∈¯Ω1,b2ifx∈Ω2, |
with b1,b2>0, the map f:L2(Ω)→L2(Ω) given by
[f(u)](x)={φ1(u(x))ifx∈¯Ω1,φ2(u(x))ifx∈Ω2, | (2.6) |
with φi:R→R, i=1,2.
The following assumptions are the requirements for well-posedness of the strong form as well as uniqueness and asymptotic stability of stationary solutions:
Assumption 2.1. Let φi satisfy
- 0≤φi≤1;
- φi is decreasing;
- φi(z)z≤c for all z∈R with c>0;
- φi is Lipschitz with constant L>0.
In Eqs (2.3) and (2.4), the square brackets [⋅]Γ are denoting the jump of the quantity in the brackets across Γ, i.e., the trace from Ω2 minus the trace of Ω1. In Eq (2.4), ν is the normal directed towards Ω2.
Define V:=H10(Ω),H:=L2(Ω). The inner product on V is defined by (u,v)V=(u,v)H+(∇u,∇v)H. Denote by ⟨⋅,⋅⟩ the pairing between V∗ and V. So we have the evolution triplet
V⊂⊂H=H∗⊂V∗, |
where H∗ and V∗ denote the dual space of H and V, respectively. Moreover, we write V⊂⊂H to emphasize the compactness of the embedding of V in H.
Define the nonlinear operator A:D(A)→H given by
A(u)=−bΔu−f(u), | (3.1) |
with
D(A):={u∈V:u|Ωi∈H2(Ωi),usatisfies(2.4)}. |
We consider the equation
dudt+A(u)=0, | (3.2) |
as an equality in L2(0,T;H). We denote by Cw(0,T;H) the space of weakly continuous functions from [0,T] to H.
Theorem 3.1. For u(0)=u0∈D(A), Eq (3.2) has a unique solution u∈C0(0,T;H) for any T>0 with
u∈Lip(0,T;D(A)),dudt∈Cw(0,T;H). |
Furthermore, u0↦u(t)∈C0(D(A),D(A)).
Proof. We will apply Theorem 31.A from [28]. Recall that A is called monotone if (A(u)−A(v),u−v)H≥0 for all u,v∈D(A). Let R denote the range of an operator and let I denote the identity operator. The assumptions to check are:
(H1): A is monotone,
(H2): R(I+A)=H.
Proof (H1): Since φ is decreasing - we obtain
(A(u)−A(v),u−v)H=∫Ωb∇(u−v)⋅∇(u−v)dx−∑i∫Ωi(φi(u)−φi(v))(u−v)⏟≤0dx≥0, | (3.3) |
for u,v∈D(A).
Proof (H2). We consider ˜A:V→V∗ given by
⟨˜A(u),w⟩:=∫Ω(uw+b∇u∇w)dx−∑i∫Ωiφi(u)wdxu,w∈V. |
Recall that ˜A is called hemicontinuous if the real function t↦⟨˜A(u+tv),w⟩ is continuous on [0,1] for all u,v,w∈V and that ˜A is called coercive if
⟨˜A(u),u⟩‖u‖V→+∞as‖u‖V→∞. | (3.4) |
We apply Theorem 26.A [28] to ˜A, which gives that for fixed g∈H, there exists a u∈V such that ˜Au=g. The assumptions to check are:
- ˜A monotone: Take u,v∈V, then similarly to Eq (3.3), we have that
⟨˜A(u)−˜A(v),u−v⟩≥∫Ωb∇(u−v)⋅∇(u−v)dx−∑i∫Ωi(φi(u)−φi(v))(u−v)dx≥0. |
- ˜A coercive: Recall from Assumptions 2.1 that φ(z)z≤c0 for all z∈R. Hence, we have that
⟨˜A(u),u⟩=∫Ω(u2+b|∇u|2)dx−∑i∫Ωiφi(u)udx≥c‖u‖2V−c0|Ω| |
and therefore, we have Eq (3.4).
- ˜A hemicontinuous: for t,s∈[0,1] u,v,w∈V, the Lipschitz continuity of φi by Assumption 2.1, we have
|⟨˜A(u+tv),w⟩−⟨˜A(u+sv),w⟩|≤C‖v‖V‖w‖V|t−s|+∑i|∫Ωi(φi(u+tv)−φi(u+sv))wdx|≤C‖v‖V‖w‖V|t−s|+L‖v‖H‖w‖H|t−s|. |
This implies the continuity of t↦⟨˜A(u+tv),w⟩.
Hence, for fixed g∈H, there exists a u∈V such that ˜Au=g, i.e.,
∫Ωb∇u∇wdx=∫Ω(−u+f(u)+g)wdx∀w∈V. | (3.5) |
It remains to show that u∈D(A), which can be done by adapting standard arguments on the regularity of weak solutions to elliptic boundary value problems to our two-phase setting [29]. More precisely, since −u+f(u)+a∈H and b is constant in Ωi, we can show, as in [30] Theorem 6.3.1, that Eq (3.5) implies u|Ωi∈H2loc(Ωi) and
−bΔu=−u+f(u)+ga.e. inΩi,i=1,2. |
To obtain H2-regularity up to the boundary, we modify the proof of [30] Theorem 6.3.4. (with partly changed notation) in the following way: We fix x0∈Γ and "flatten" Γ locally in a neighborhood Ux0 of x0 by an appropriate C2-diffeomorphism Φ:Ux0⟶B(0,1) so that we have
Φ(Ux0∩Ω1)=B(0,1)∩{y∈RN|yN>0},Φ(Ux0∩Ω2)=B(0,1)∩{y∈RN|yN<0}. |
For the transformed problem, the jump of b occurs along the plane yN=0. Therefore, the estimates for the difference quotients of the transformed solution in the first N−1 coordinate directions can be carried out in both half balls as in the proof of the cited theorem. Thus, we analogousy obtain u|Ωi∈H2(Ωi). Hence, applying integration by parts in the subdomains,
∫Ωb∇u∇wdx=∫Ω1b∇u∇wdx+∫Ω2b∇u∇wdx=−∫Ω1bΔuwdx−∫Ω2bΔuwdx−∫Γ[b∇u⋅ν]Γwds=∫Ω(−u+f(u)+g)wdx−∫Γ[b∇u⋅ν]Γwds∀w∈V, |
so by Eq (3.5), the boundary integral vanishes for all w∈V, which implies Eq (2.4), and u∈D(A) is proved.
Finally, applying Corollary 31.1 from [28] gives that u0↦u(t)∈C0(D(A),D(A)).
We define the functional E:V→R given by
E(u)=∫Ω12b|∇u|2dx−∑i∫ΩiFi(u)dx, | (4.1) |
where
Fi(s)=∫s0φi(σ)dσ. | (4.2) |
Lemma 4.1. E is Fréchet differentiable with derivative at u given by
E′(u)[h]=∫Ωb∇u∇hdx−∫Ωf(u)hdx. |
Proof. The first term in Eq (4.1) is a quadratic term in V. Fi has a bounded, integrable weak second derivative. Hence, for z,ζ∈R we have
|Fi(z+ζ)−Fi(z)−F′i(z)ζ|=|ζ2∫10(1−s)F″i(z+sζ)ds|≤12‖F″i‖∞ζ2. |
Thus, for u,h∈V
|∫ΩiFi(u+h)−Fi(u)−F′i(u)h|≤12‖F″i‖∞‖h‖2V, |
and ui↦∫ΩiFi(u)dx is Fréchet differentiable with derivative given by
h↦∫ΩiF′i(u)hdx=∫Ωiφi(u)hdx. |
Lemma 4.2. E′ is strongly monotone, i.e., for all u,v∈V, there exists γ>0 such that
(E′(u)−E′(v))[u−v]≥γ‖u−v‖2V. |
Proof. Take u,v∈V, then using φi decreasing and Poincaré's inequality we obtain
(E′(u)−E′(v))[u−v]=∫Ωb(∇u−∇v)⋅(∇u−∇v)dx−∑i∫Ωi(φi(u)−φi(v))(u−v)⏟≤0dx≥γ‖u−v‖2V. |
From Lemma 4.1, it follows that we can write Eq (3.2) as a gradient flow:
ut=−∇E(u), |
in the sense that
(ut,w)H=−E′(u)[w]. |
Recall that u∗∈V is called a critical point if E′(u∗)v=0. Hence, stationary solutions are critical points of E.
Theorem 4.3. There is precisely one stationary solution to Eq (3.2).
Proof. If E is a continuous, strictly convex, coercive functional then E has a unique critical point which is also a global minimum by Theorems 1.5.6 and 1.5.7 in [31]. E is continuous. Strict convexity of E follows from Lemma 4.2. So we only need to show coercivity. Observe that if s>0, then Fi(s)≤s, and if s≤0, then Fi(s)≤0. Hence, Fi(s)≤|s| for all s∈R. Therefore, using Poincaré's inequality, we have that
E(u)≥C‖u‖2V−C‖u‖V. |
Hence, E is coercive. The result follows from Lemma 4.1.
Denote the unique stationary solution by u∗∈V.
Theorem 4.4 (Global asymptotic stability in H). Let u be a solution of Eq (3.2), then t↦eγt‖u(t)−u∗‖H is bounded for t≥0.
Proof. Observe that E′(u∗)=0. From u∈L2(0,T;V), dudt∈L2(0,T;V∗) we have that
d‖u‖2Hdt=2(dudt,u)H, | (4.3) |
by Theorem 7.2 in [32]. Now let u be the solution of Eq (3.2). Then by Eq (4.3) and Lemma 4.2 we have that
12ddt(‖u−u∗‖2H)=(dudt,u−u∗)H=−E′(u)[u−u∗],≤(E′(u∗)−E′(u))[u−u∗]≤−γ‖u−u∗‖2V≤−γ‖u−u∗‖2H. |
Hence, we obtain that ‖u(t)−u∗‖2H≤e2γt‖u(0)−u∗‖2H.
Asymptotic stability in the V-norm does not follow from Theorem 4.4, but we can obtain a weak V-stability result:
Corollary 4.5. u(t)⇀u∗ in V as t→∞.
Proof. We first show that
t↦u(t)isboundedinV. | (4.4) |
We can bound ‖u‖V in terms of E(u) and ‖u‖H:
‖u‖2V≤C∫Ωb|∇u|2dx=2C(E(u)+∑i∫ΩiFi(u)dx)≤C′(E(u)+‖u‖H). |
Now Eq (4.4) follows because E(u) is decreasing and ‖u‖H is bounded by Theorem 4.4.
Now suppose the opposite of Corollary 4.5. Then, there exists ε>0 and ϕ∈V∗ and a sequence (tn),tn→∞ such that
|⟨ϕ,u(tn)−u∗⟩|≥ε, | (4.5) |
for all n. Write un:=u(tn). Since V is reflexive and (un) is bounded according to Eq (4.4), we obtain from Alaoglu's compactness theorem that there is a subsequence again denoted by (un) that converges weakly in V. As V⊂⊂H, the subsequence (un) converges strongly in H; therefore, the limit is u∗ which is in contradiction with Eq (4.5).
We will formulate a PDE for the model considered in [13]. We assume that the transplanted cells are injected in an oxygen-stable environment [1]. Hence, we consider on S a Dirichlet boundary condition. At Γ we require the concentration and flux to be continuous. Finally, on Ω1 there is a non-linear term corresponding to Michaelis–Menten consumption [13]. It is assumed that on Ω2 there is no oxygen consumption. We also assume that Michaelis–Menten consumption is zero when the oxygen concentration is zero. This condition ensures that the non-linear term is bounded and monotone. The equations for the non-dimensional oxygen concentration, v=v(x,t), are then given by
dvdt−bΔv=−g(v)in(Ω∖Γ)×(0,T),v|S=c0,[v]|Γ=0,[b∂v∂n]|Γ=0,v(x,0)=v0(x)inΩ, |
where g:L2(Ω)→L2(Ω) is defined by
[g(v)](x)={ϕ(v(x))ifx∈¯Ω1,0ifx∈Ω2, |
with ϕ:R→R given by
ϕ(z)={zz+ˆcifz≥0,0else, |
with ˆc>0. Let u=−v+c0 and we obtain Eqs (2.1)–(2.5). Setting c1:=c0+ˆc we have that
φ1(z)={c0−zc1−z ifz≤c0,0 ifz>c0, | (5.1) |
where 0<c0<c1, see Figure 2.
It is straightforward to see that φ1 satisfies the properties in Assumptions 2.1. More specifically, the Lipschitz constant is 1/(c1−c0), and zφ1(z)≤c0 for all z∈R. Consequently, we can apply Theorems 3.1, 4.3, and 4.4 to obtain well-posedness and global asymptotic stability of the unique stationary solution.
For the most general form of kinetics, Hill's equation, the theorems can also be applied. In the transformed variables we then obtain that
φi(z)={(c0−z)n(c0−z)n+ˆcifz≤c0,0ifz>c0, | (5.2) |
with n,ˆc>0 which satisfies Assumption 2.1. Note that for n=1, Eq (5.2) becomes Michaelis–Menten Eq (5.1).
In this work, we have shown the well-posedness of a nonlinear reaction–diffusion equation for general core-shell geometry. Furthermore, the corresponding stationary solutions are unique and asymptotically stable in a suitable topology. These results extend the model by [13], which only considers stationary solutions for spherical core-shell geometry.
The well-posedness theorem, Theorem 3.1, allows us to define a semi-dynamical system: (D(A),{S(t)}t≥0). This can be used to prove the existence of a global attractor following techniques in [32]. We expect that Theorem 10.13 from [32] can be applied and consequently that the global attractor is equal to the unique stationary solution. We note that the techniques from [32,33] can be used to prove well-posedness by only relying on an L2 bound on f and not on the monotonicity of f [34].
In view of general results on parabolic PDE systems, we expect our evolution problem to be well-posed in Hölder spaces as well, if Γ is smooth enough. However, the proof would be rather technical, and the improvement might not be essential from the point of view of the application. Consequently, it was not considered in this work.
In an experimental set-up, f might be unknown. Suppose that the following are known: b, Ω1,Ω2, and spatially dependent data for u when u has become stationary. The nonlinear f can be approximated using the Sparse Identification of Nonlinear Dynamics (SINDy) framework, which in short, is a regression-type fitting process, typically LASSO, over a set of library functions [35]. There are specialized SINDy frameworks for PDEs [36]. Observe that the data concerns stationary u in time so we can ignore the time derivative, which reduces the complexity of the fitting process. We note that the problem can be further reduced if we can control Ω1 and Ω2 in the experimental set-up. We would then select a spherical geometry such that Ω1 and Ω2 have a common center. Assuming that the obtained data for u also has spherical geometry, the problem reduces to applying the SINDy-framework to a first-order ODE, which is a vanilla-type SINDy.
It is tempting to make the connection between this work and bulk-surface PDEs [37,38,39,40]. In our setting bulk-surface geometry means that the outer shell is a surface. If we would consider our governing equations and then take the limit to zero for the thickness of the shell it appears to be necessary to assume that Neumann boundary conditions are imposed on the shell, to arrive at the Bulk-Surface PDE [41].
Besides oxygen transport, glucose transport is also needed to sustain encapsulated donor cells. In [42], a coupled glucose–oxygen transport model is proposed based on biological assumptions from [21,43]. As in [13], stationary solutions are considered, and a numerical study is performed to find solutions that have concentrations which are above the donor cell survival threshold. The well-posedness and stability of the corresponding PDE is a topic we would like to explore in future work.
Thomas de Jong: Conceptualization, Writing original draft. Georg Prokert: Review, Methodology. Alef Sterk: Review, Methodology.
The authors declare they have not used artificial intelligence (AI) tools in the creation of this article.
During this research Thomas de Jong was also affiliated with Xiamen University and University of Groningen. This research was supported by JST CREST grant number JPMJCR2014.
The authors declare that there are no conflicts of interest.
[1] | A. M. J. Shapiro, M. Pokrywczynska, C. Ricordi, Clinical pancreatic islet transplantation, Nat. Rev. Endocrinol., 13 (2017), 268–277. https://doi.org/nrendo.2016.178 |
[2] |
M. F. Goosen, G. M. O'Shea, H. M. Gharapetian, S. Chou, A. M. Sun, Optimization of microencapsulation parameters: Semipermeable microcapsules as a bioartificial pancreas, Biotechnol. Bioeng., 2 (1985), 146–150. https://doi.org/10.1002/bit.260270207 doi: 10.1002/bit.260270207
![]() |
[3] |
A. D. Augst, H. J. Kong, D. J. Mooney, Alginate hydrogels as biomaterials, Macromol. Biosci., 8 (2006), 623–633. https://doi.org/10.1002/mabi.200600069 doi: 10.1002/mabi.200600069
![]() |
[4] |
A. Espona-Noguera, J. Ciriza, A. Cañibano-Hernández, L. Fernandez, I. Ochoa, L. S. Del Burgo, et al., Tunable injectable alginate-based hydrogel for cell therapy in type 1 Diabetes Mellitus, Int. J. Biol. Macromol., 107 (2018), 1261–1269. https://doi.org/10.1016/j.ijbiomac.2017.09.103 doi: 10.1016/j.ijbiomac.2017.09.103
![]() |
[5] |
A. Espona-Noguera, J. Ciriza, A. Canibano-Hernandez, R. Villa, L. S. Del Burgo, M. Alvarez, et al., 3D printed polyamide macroencapsulation devices combined with alginate hydrogels for insulin-producing cell-based therapies, Int. J. Pharm., 566 (2019), 604–614. https://doi.org/10.1016/j.ijpharm.2019.06.009 doi: 10.1016/j.ijpharm.2019.06.009
![]() |
[6] |
R. B. Elliott, L. Escobar, P. L. J. Tan, O. Garkavenko, R. Calafiore, P. Basta, et al., Intraperitoneal alginate-encapsulated neonatal porcine islets in a placebo-controlled study with 16 diabetic cynomolgus primates, Transplantat. Proc., 37 (2005), 3505–3508. https://doi.org/10.1016/j.transproceed.2005.09.038 doi: 10.1016/j.transproceed.2005.09.038
![]() |
[7] |
D. Dufrane, R. M. Goebbels, P. Gianello, Alginate macroencapsulation of pig islets allows correction of streptozotocin-induced diabetes in primates up to 6 months without immunosuppression, Transplantation, 90 (2010), 1054–1062. https://doi.org/10.1097/TP.0b013e3181f6e267 doi: 10.1097/TP.0b013e3181f6e267
![]() |
[8] | M. A. Bochenek, O. Veiseh, A. J. Vegas, J. J. McGarrigle, M. Qi, E. Marchese, M. Omami, et al., Alginate encapsulation as long-term immune protection of allogeneic pancreatic islet cells transplanted into the omental bursa of macaques, Nat. Biomed. Eng., 2 (2018), 810–821. https://doi.org/s41551-018-0275-1 |
[9] |
D. A. Alagpulinsa, J. J. L. Cao, R. K. Driscoll, R. F. Sîrbulescu, M. F. E. Penson, M. Sremac, et al., Alginate-microencapsulation of human stem cell–derived β cells with CXCL12 prolongs their survival and function in immunocompetent mice without systemic immunosuppression, Am. J. Transplant., 19 (2019), 1930–1940. https://doi.org/10.1111/ajt.15308 doi: 10.1111/ajt.15308
![]() |
[10] | R. Calafiore, G. Basta, G. Luca, A. Lemmi, M. P. Montanucci, G. Calabrese, et al., Microencapsulated pancreatic islet allografts into nonimmunosuppressed patients with type 1 diabetes: First two cases, Diabetes Care, 29 (2006), 137–138. |
[11] | B. E. Tuch, G. W. Keogh, L. J. Williams, W. Wu, J. L. Foster, V. Vaithilingam, et al., Safety and viability of microencapsulated human islets transplanted into diabetic humans, Diabetes Care, 32 (1999), 1887–1889. |
[12] |
G. Basta, P. Montanucci, G. Luca, C. Boselli, G. Noya, B. Barbaro, et al., Long-term metabolic and immunological follow-up of nonimmunosuppressed patients with type 1 diabetes treated with microencapsulated islet allografts: Four cases, Diabetes Care, 34 (2011), 2406–2409. https://doi.org/10.2337/dc11-0731 doi: 10.2337/dc11-0731
![]() |
[13] |
C. C. King, A. A. Brown, I. Sargin, K. M. Bratlie, S. P. Beckman, Modelling of reaction-diffusion transport into core-shell geometry, J. Theor. Biol., 260 (2019), 204–208. https://doi.org/10.1016/j.jtbi.2018.09.026 doi: 10.1016/j.jtbi.2018.09.026
![]() |
[14] |
C. C. King, A. A. Brown, I. Sargin, K. M. Bratlie, S.P. Beckman, Corrigendum to: Modeling of reaction-diffusion transport into a core-shell geometry, J. Theor. Biol., 507 (2020), 110439. https://doi.org/10.1016/j.jtbi.2020.110439 doi: 10.1016/j.jtbi.2020.110439
![]() |
[15] |
M. Ma, A. Chiu, G. Sahay, J. C. Doloff, N. Dholakia, R. Thakrar, et al., Core-shell hydrogel microcapsules for improved islets encapsulation, Adv. Healthcare Mater., 2 (2013), 1–12. https://doi.org/10.1002/adhm.201200341 doi: 10.1002/adhm.201200341
![]() |
[16] |
T. G. de Jong, A. E. Sterk, Topological shooting of solutions for fickian diffusion into core-shell geometry, Nonlinear Dyn. Discrete Contin. Syst., (2021), 103–116. https://doi.org/10.1007/978-3-030-53006-8_7 doi: 10.1007/978-3-030-53006-8_7
![]() |
[17] | W. C. Troy, L. A. Peletier, A topological shooting method and the existence of kinks of the extended fisher-kolmogorov equation, Topol. Methods Nonlinear Anal., 6 (1995), 331–355. |
[18] | J. B. McLeod, S. P. Hastings, Classical Methods in Ordinary Differential Equations, American Mathematical Society, Rhode Island, 2012. |
[19] |
T. G. de Jong, A. E. Sterk, H. W. Broer, Fungal tip growth arising through a codimension-1 bifurcation, Int. J. of Bifurcation Chaos, 30 (2020), 2050107. https://doi.org/10.1142/S0218127420501072 doi: 10.1142/S0218127420501072
![]() |
[20] |
E. S. Avgoustiniatos, K. E. Dionne, D. F Wilson, M. L. Yarmush, Measurements of the effective diffusion coefficient of oxygen in pancreatic islets, Ind. Eng. Chem. Res., 46 (2007), 6157–6163. https://doi.org/10.1021/ie070662y doi: 10.1021/ie070662y
![]() |
[21] | P. Buchwald, FEM-based oxygen consumption and cell viability models for avascular pancreatic islets, Theor. Biol. Med. Modell., 6 (2009). https://doi.org/10.1186/1742-4682-6-5 |
[22] |
S. H. Lin, Oxygen diffusion in a spherical cell with nonlinear oxygen uptake kinetics, J. Theor. Biol., 60 (1976), 449–457. https://doi.org/10.1016/0022-5193(76)90071-0 doi: 10.1016/0022-5193(76)90071-0
![]() |
[23] |
D. L. S. McElwain, A re-examination of oxygen diffusion in a spherical cell with michaelis-menten oxygen uptake kinetics, J. Theor. Biol., 71 (1978), 255–263. https://doi.org/10.1016/0022-5193(78)90270-9 doi: 10.1016/0022-5193(78)90270-9
![]() |
[24] |
P. Hiltmann, P. Lory, On oxygen diffusion in a spherical cell with michaelis-menten oxygen uptake kinetics, Bull. Math. Biol., 45 (1983), 661–664. https://doi.org/10.1007/BF02460043 doi: 10.1007/BF02460043
![]() |
[25] |
M. J. Simpson, A. J. Ellery, An analytical solution for diffusion and nonlinear uptake of oxygen in a spherical cell, Appl. Math. Modell., 36 (2012), 3329–3334. https://doi.org/10.1016/j.apm.2011.09.071 doi: 10.1016/j.apm.2011.09.071
![]() |
[26] | O. A. Ladyzhenskaya, The Boundary Value Problems of Mathematical Physics, Springer Science & Business Media, 2013. |
[27] |
N. T. Levashova, N. N. Nefedov, O. A. Nikolaeva, A. O. Orlov, A. A. Panin, The solution with internal transition layer of the reaction-diffusion equation in case of discontinuous reactive and diffusive terms, Math. Methods Appl. Sci., 41 (2018), 9203–9217. https://doi.org/10.1002/mma.5134 doi: 10.1002/mma.5134
![]() |
[28] | E. Zeidler, Nonlinear Functional Analysis and Its Applications: II/B: Nonlinear Monotone Operators, Springer Science & Business Media, 2013. |
[29] | V. A. Il'in, I. A. Šišmarev, The connection between generalized and classical solutions of the dirichlet problem (in Russian), Izv. Akad. Nauk. SSSR, 24 (1960), 521–530. |
[30] | L. C. Evans, Partial Differential Equations, American Mathematical Society, 1998. |
[31] | M. Badiale, E. Serra, Semilinear Elliptic Equations for Beginners: Existence Results via the Variational Approach, Springer Science & Business Media, 2010. |
[32] | J. C. Robinson, Infinite-Dimensional Dynamical Systems: An Introduction to Dissipative Parabolic PDEs and the Theory of Global Attractors, Cambridge University Press, 2001. |
[33] |
M. Marion, Attractors for reaction-diffusion equations: Existence and estimate of their dimension, Appl. Anal., 25 (1987), 101–147. https://doi.org/10.1080/00036818708839678 doi: 10.1080/00036818708839678
![]() |
[34] | Y. Nakamura, K. S. Putri, A. E. Sterk, T. G. de Jong, Well-posedness results for general reaction–diffusion transport of oxygen in encapsulated cells, preprint, arXiv: 2401.09863. |
[35] |
S. L. Brunton, J. L. Proctor, J. N. Kutz, Discovering governing equations from data by sparse identification of nonlinear dynamical systems, Proc. Natl. Acad. Sci., 113 (2016), 3932–3937. https://doi.org/10.1073/pnas.1517384113 doi: 10.1073/pnas.1517384113
![]() |
[36] |
D. A. Messenger, D. M. Bortz, Weak sindy for partial differential equations, J. Comput. Phys., 443 (2021), 110525. https://doi.org/10.1016/j.jcp.2021.110525 doi: 10.1016/j.jcp.2021.110525
![]() |
[37] |
A. Rätz, M. Röger, Symmetry breaking in a bulk–surface reaction–diffusion model for signalling networks, Nonlinearity, 27 (2014), 1805. https://doi.org/10.1088/0951-7715/27/8/1805 doi: 10.1088/0951-7715/27/8/1805
![]() |
[38] |
F. Paquin-Lefebvre, B. Xu, K. L. DiPietro, A. E. Lindsay, A. Jilkine, Pattern formation in a coupled membrane-bulk reaction-diffusion model for intracellular polarization and oscillations, J. Theor. Biol., 497 (2020), 110242. https://doi.org/10.1016/j.jtbi.2020.110242 doi: 10.1016/j.jtbi.2020.110242
![]() |
[39] |
A. Madzvamuse, A. Chung, C. Venkataraman, Stability analysis and simulations of coupled bulk-surface reaction–diffusion systems, Proc. R. Soc. A, 471 (2015), 20140546. https://doi.org/10.1098/rspa.2014.0546 doi: 10.1098/rspa.2014.0546
![]() |
[40] |
A. Diez, A. L. Krause, P. K. Maini, E. A. Gaffney, S. Seirin-Lee, Turing pattern formation in reaction-cross-diffusion systems with a bilayer geometry, Bull. Math. Biol., 86 (2023). https://doi.org/10.1007/s11538-023-01237-1 doi: 10.1007/s11538-023-01237-1
![]() |
[41] |
J. Li, L. Su, X. Wang Y. Wang, Bulk-surface coupling: Derivation of two models, J. Differ. Equations, 289 (2021), 1–34. https://doi.org/10.1016/j.jde.2021.04.011 doi: 10.1016/j.jde.2021.04.011
![]() |
[42] |
C. C. King, S. P. Beckman, Coupled reaction–diffusion transport into a core–shell geometry, J. Theor. Biol., 546 (2022), 111138. https://doi.org/10.1016/j.jtbi.2022.111138 doi: 10.1016/j.jtbi.2022.111138
![]() |
[43] |
P. Buchwald, A local glucose-and oxygen concentration-based insulin secretion model for pancreatic islets, Theor. Biol. Med. Modell., 8 (2011), 1–25. https://doi.org/10.1186/1742-4682-8-20 doi: 10.1186/1742-4682-8-20
![]() |