Skip to main content
Log in

Global Existence and Decay Estimates of Solutions of a Parabolic–Elliptic–Parabolic System for Ion Transport Networks

  • Published:
Results in Mathematics Aims and scope Submit manuscript

Abstract

This paper deals with a parabolic-elliptic-parabolic system arising from ion transport networks originally proposed by Albi et al. (Anal Appl 14(1):185–206, 2016). Our investigation demonstrates that the local existence of the weak solution can always be achieved. Moreover, by establishing the uniform a-priori decay estimates of solution, the local solution can be extended to a global one provided that the initial data is suitably small. In addition, the time-decay rate of the weak solution is obtained. To our best knowledge, this seems to be the first analytical work on this system indeed.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Similar content being viewed by others

References

  1. Albi, G., Artina, M., Fornasier, M., Markowich, P.: Biological transportation networks: modeling and simulation. Anal. Appl. 14(1), 185–206 (2016)

    Article  MathSciNet  Google Scholar 

  2. Amann, H.: Linear and Quasilinear Parabolic Problems, Volume I: Abstract Linear Theory. Birkhauser, Basel (1995)

    Book  Google Scholar 

  3. Brezis, H., Cazenave, T.: A nonlinear heat equation with singular initial data. J. d’Analyse Math. 68(1), 277–304 (1996)

    Article  MathSciNet  Google Scholar 

  4. Burger, M., Haskovec, J., Markowich, P., Ranetbauer, H.: A mesoscopic model of biological transportation networks. Commun. Math. Sci. 17(5), 1213–1234 (2019)

    Article  MathSciNet  Google Scholar 

  5. Burger, M., Schlake, B., Wolfram, M.: Nonlinear Poisson–Nernst–Planck equations for ion flux through confined geometries. Nonlinearity 25(4), 961–990 (2012)

    Article  MathSciNet  Google Scholar 

  6. Cao, X., Lankeit, J.: Global classical small-data solutions for a three-dimensional chemotaxis Navier–Stokes system involving matrix-valued sensitivities. Calculus Var. Partial Differ. Equ. 55(4), paper No. 107, 39 pp (2016)

  7. Chen, J., Li, Y., Wang, W.: Global classical solutions to the Cauchy problem of conservation laws with degenerate diffusion. J. Differ. Equ. 260(5), 4657–4682 (2016)

    Article  MathSciNet  Google Scholar 

  8. Davies, E.: Heat Kernels and Spectral Theory. Cambridge University Press, Cambridge (1989)

    Book  Google Scholar 

  9. Evans, L.: Partial Differential Equations, 2nd edn. Amer. Math. Soc., Providence (2010)

    MATH  Google Scholar 

  10. Gilbarg, D., Trudinger, N.: Elliptic Partial Differential Equations of Second Order. Springer, New York (1983)

    Book  Google Scholar 

  11. Gokhale, B.: Numerical solutions for a one-dimensional silicon n-p-n transistor. IEEE Trans. Electron Devices 17(8), 594–602 (1970)

    Article  Google Scholar 

  12. Haskovec, J., Jönsson, H., Kreusser, L., Markowich, P.: Auxin transport model for leaf venation. Proc. R. Soc. A 475(2231): paper No. 20190015 (2019)

  13. Haskovec, J., Kreusser, L., Markowich, P.: ODE and PDE based modeling of biological transportation networks. Commun. Math. Sci. 17(5), 1235–1256 (2019)

    Article  MathSciNet  Google Scholar 

  14. Haskovec, J., Kreusser, L., Markowich, P.: Rigorous continuum limit for the discrete network formation problem. Commun. Partial Differ. Equ. 44(11), 1159–1185 (2019)

    Article  MathSciNet  Google Scholar 

  15. Haskovec, J., Markowich, P., Perthame, B.: Mathematical analysis of a PDE system for biological network formation. Commun. Partial Differ. Equ. 40(5), 918–956 (2015)

    Article  MathSciNet  Google Scholar 

  16. Haskovec, J., Markowich, P., Perthame, B., Schlottbom, M.: Notes on a PDE system for biological network formation. Nonlinear Anal. Theory Methods Appl. 138(5), 127–155 (2016)

    Article  MathSciNet  Google Scholar 

  17. Horstmann, D.: From 1970 until now: the Keller-Segel model in chemotaxis and its consequences I. Jahresberichte der DMV 105, 103–165 (2003)

    MATH  Google Scholar 

  18. Horstmann, D.: From 1970 until now: the Keller-Segel model in chemotaxis and its consequences II. Jahresberichte der DMV 106, 51–69 (2004)

    MATH  Google Scholar 

  19. Hu, D.: Optimization, adaptation, and initialization of biological transport networks. Notes from Lecture (2013)

  20. Hu, D., Cai, D.: Adaptation and optimization of biological transport networks. Phys. Rev. Lett. 111(13): paper No. 138701 (2013)

  21. Hu, D., Cai, D.: An optimization principle for initiation and adaptation of biological transport networks. Commun. Math. Sci. 17(5), 1427–1436 (2019)

    Article  MathSciNet  Google Scholar 

  22. Ladyzhenskaya, O., Solonnikov, V., Ural’tseva, N.: Linear and Quasi-linear Equations of Parabolic Type, vol. 23. Amer. Math. Soc. Transl, Providence (1968)

    Book  Google Scholar 

  23. Ladyzhenskaya, O., Ural’tseva, N.: Linear and Quasi-linear Elliptic Equations. Academic Press, New York (1968)

    Book  Google Scholar 

  24. Li, B.: Long time behavior of the solution to a parabolic-elliptic system. Comput. Math. Appl. 78(10), 3345–3362 (2019)

    Article  MathSciNet  Google Scholar 

  25. Li, B.: On the blow-up criterion and global existence of a nonlinear PDE system in biological transport networks. Kinetic Relat. Models 12(5), 1131–1162 (2019)

    Article  MathSciNet  Google Scholar 

  26. Li, B., Shen, J.: Classical solution of a PDE system stemming from auxin transport model for leaf venation. Proc. Am. Math. Soc. (2020). https://doi.org/10.1090/proc/14951

    Article  MathSciNet  Google Scholar 

  27. Liu, J., Xu, X.: Partial regularity of weak solutions to a PDE system with cubic nonlinearity. J. Differ. Equ. 264(8), 5489–5526 (2018)

    Article  MathSciNet  Google Scholar 

  28. Markowich, P., Ringhofer, C., Schmeiser, C.: Semiconductor Equations. Springer, New York (1990)

    Book  Google Scholar 

  29. Mock, M.: An initial value problem from semiconductor device theory. SIAM J. Math. Anal. 5(4), 597–612 (1974)

    Article  MathSciNet  Google Scholar 

  30. Mock, M.: Asymptotic behavior of solutions of transport equations for semiconductor devices. J. Math. Anal. Appl. 49(1), 215–225 (1975)

    Article  MathSciNet  Google Scholar 

  31. Perthame, B.: Transport Equations in Biology. Birkhauser, Basel (2007)

    Book  Google Scholar 

  32. Quittner, P., Souplet, P.: Superlinear Parabolic Problems: Blow-up, Global Existence and Steady States. Springer, Berlin (2007)

    MATH  Google Scholar 

  33. Shen, J., Li, B.: A-priori estimates for a nonlinear system with some essential symmetrical structures. Symmetry 11(7): paper No. 852 (2019)

  34. Tao, Q., Yao, Z.: Global existence and large time behavior for a two-dimensional chemotaxis-shallow water system. J. Differ. Equ. 265(7), 3092–3129 (2018)

    Article  MathSciNet  Google Scholar 

  35. Weissler, F.: Semilinear evolution equations in banach spaces. J. Funct. Anal. 32(3), 277–296 (1979)

    Article  MathSciNet  Google Scholar 

  36. Winkler, M.: Aggregation vs. global diffusive behavior in the higher-dimensional Keller-Segel model. J. Differ. Equ. 248(12), 2889–2905 (2010)

    Article  MathSciNet  Google Scholar 

  37. Xu, X.: Regularity theorems for a biological network formulation model in two space dimensions. Kinetic Relat. Models 11(2), 397–408 (2018)

    Article  MathSciNet  Google Scholar 

  38. Xu, X.: Partial regularity of weak solutions and life-span of smooth solutions to a PDE system with cubic nonlinearity (2018). arXiv:1706.06057v5

  39. Xu, X.: Global existence of strong solutions to a biological network formulation model in 2 + 1 dimensions (2019). arXiv:1911.01970v1

Download references

Acknowledgements

The author is deeply indebted to the anonymous referee for his/her valuable comments and helpful suggestions which help to significantly improve the manuscript, and to the editor for his/her kind help. The author also thanks Professor Zhaoyin Xiang for his suggesting of this problem and for his valuable discussions.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Bin Li.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendix

Appendix

In this concluding section, we provide the rigorous proofs of Lemmas 2.4 and 3.1.

Proof of Lemma 2.4

When \(\gamma =1\), by applying the theory of the linear parabolic equation, the solution \(\mathbf{m}\in C\big ([0,T],L^p(\Omega )\big )\) follows directly. Moreover, by applying [22, Theorem 3.7.1] to the Dirichlet problem associated with the linear heat equation with a potential

$$\begin{aligned} \mathbf{m}_t-\Delta \mathbf{m}+\mathbf{A}(t,x)\mathbf{m}=0\quad \mathrm {with}\quad \mathbf{A}:=I-\nabla \psi \otimes \nabla \psi , \end{aligned}$$

we conclude from \(\nabla \psi \in L^\infty \big ([0,T], H^2(\Omega )\big )\) and Sobolev’s embedding that \(\mathbf{m}\in C\big ((0,T],L^\infty (\Omega )\big )\). From this, \(\mathbf{m}\in C^{\alpha ,\frac{\alpha }{2}}(\Omega _T)\) by employing [22, Theorem 3.10.1].

Next, we consider the case of \(\gamma >1\). We begin with \(p=\infty \). A well-known construction, on the basis of the contraction mapping principle and Lemma 2.1, asserts the existence of solution of (2.8). This solution satisfies

$$\begin{aligned} \Vert \mathbf{m}(t)\Vert _{L^r(\Omega )}\le Ce^{kt}\Vert \mathbf{m}_0\Vert _{L^r(\Omega )},\ \ \Vert \mathbf{m}(t)\Vert _{L^\infty (\Omega )}\le Ce^{kt}t^{-\frac{n}{2r}}\Vert \mathbf{m}_0\Vert _{L^r(\Omega )} \end{aligned}$$
(5.1)

for any \(t\in (0,T]\) and \(r\in [1, \infty ]\), where k depends on \(\Vert \mathbf{m}\Vert _{L^\infty ((0,T),L^{\frac{p}{2(\gamma -1)}}(\Omega ))}\) and \(\Vert \nabla \psi \Vert _{L^\infty ((0,T),L^\sigma (\Omega ))}\) with some \(\sigma >n\), see Theorem A1 in [3]. Moreover, we see from [22, Theorem 3.10.1] that \(\mathbf{m}\in C^{\theta ,\frac{\theta }{2}}(\Omega _T)\) for some \(\theta \in (0,1)\), due to \(\psi \in L^\infty \big ([0,T];H^1_0(\Omega )\cap H^2(\Omega )\big )\).

It remains to show that the same conclusion holds for \(p\in [2,\infty )\) and \(p>n(\gamma -1)\) when \(\gamma >1\). In this case, we can establish the existence of the solution of (2.8) by employing the contraction mapping principle in a somewhat unusual space (this idea is due to [3, 35]). Let \(T\in (0,1)\) to be specified below. For simplicity, we set

$$\begin{aligned} X:=L^\infty \big ((0,T),L^p(\Omega )\big )\cap L^\infty _{\text {loc}}\big ((0,T),L^{p(2\gamma -1)}(\Omega )\big ), \end{aligned}$$

and denote the closed bounded convex subset \(X_R\) of X by

$$\begin{aligned} X_R:=\big \{f\in X: \Vert f\Vert _{L^p(\Omega )}\le R+1 \quad \text {and}\quad t^\alpha \Vert f\Vert _{L^\beta (\Omega )}\le R+1\quad \text {for}\ t\in (0,T)\big \}, \end{aligned}$$

where \(\beta :=p(2\gamma -1)\) and \(\alpha :=\frac{n(\gamma -1)}{p(2\gamma -1)}<\frac{1}{2\gamma -1}<1\), due to \(\gamma >1\). We equip \(X_R\) with the distance

$$\begin{aligned} d(f,g)=\sup _{t\in (0,T)}t^\alpha \Vert f(t)-g(t)\Vert _{L^{\beta }(\Omega )}, \end{aligned}$$

so that \((X_R,d)\) is a nonempty complete metric space.

We first fix an \(R\ge \max \{C_1,C_4\}\Vert \mathbf{m}_0\Vert _{L^p(\Omega )}\), where \(C_1\) and \(C_4\) are determined by (5.3) and (5.4) below. Then for any given \(\mathbf{m}\in X_R\), we set

$$\begin{aligned} \Psi (\mathbf{m})(t):=e^{\Delta t}\mathbf{m}_0+\int _0^te^{\Delta (t-s)}\big ((\mathbf{m}\cdot \nabla \psi )\nabla \psi -|\mathbf{m}|^{2(\gamma -1)}\mathbf{m}\big )(s)ds, \end{aligned}$$

where \(e^{\Delta t}\) satisfies the standard smoothing effect: let \(1\le a\le b\le \infty \), one has

$$\begin{aligned} \Vert e^{\Delta t}f\Vert _{L^b(\Omega )}\le C t^{-\frac{n}{2}\big (\frac{1}{a}-\frac{1}{b}\big )}\Vert f\Vert _{L^a(\Omega )}, \end{aligned}$$
(5.2)

where we only use the special case of (2.1), due to \(t\le T<1\). Since \(\psi \in L^\infty \big ([0,T];H^1_0(\Omega )\cap H^2(\Omega )\big )\), we have \(\nabla \psi \in L^\infty \big ([0,T], L^r(\Omega )\big )\) by Sobolev’s embedding, where \(r\in [1,\infty ]\) if \(n=1\), \(r\in [1,\infty )\) if \(n=2\) and \(r\in [1,6]\) if \(n=3\). Utilizing (5.2) we can obtain that, if \(p\ge 2\),

$$\begin{aligned} \Vert \Psi (\mathbf{m})(t)\Vert _{L^p(\Omega )}&\le C\Vert \mathbf{m}_0\Vert _{L^p(\Omega )} +C\int _0^t(t-s)^{-\frac{n}{6}}\Vert (\mathbf{m}\cdot \nabla \psi ) \nabla \psi \Vert _{L^\frac{3p}{3+p}(\Omega )}\\&\quad +\big \Vert |\mathbf{m}|^{2(\gamma -1)} \mathbf{m}(s)\big \Vert _{L^p(\Omega )}ds\\ \le&C\Vert \mathbf{m}_0\Vert _{L^p(\Omega )}+C\Vert \nabla \psi \Vert _{L^\infty \left( [0,T], L^6(\Omega )\right) }^2\int _0^t(t-s)^{-\frac{n}{6}} \Vert \mathbf{m}(s)\Vert _{L^p(\Omega )}ds\\&\quad +C\int _0^t\big \Vert \mathbf{m}(s)\big \Vert _{L^\beta (\Omega )}^{2\gamma -1}ds\\&\le C\Vert \mathbf{m}_0\Vert _{L^p(\Omega )}+C\Vert \nabla \psi \Vert _{L^\infty \left( [0,T], L^6(\Omega )\right) }^2(R+1)T^{1-\frac{n}{6}}\\&\quad +C\big (t^\alpha \Vert \mathbf{m}\Vert _{L^\beta (\Omega )}\big )^{2\gamma -1} \int _0^ts^{-\alpha (2\gamma -1)}ds, \end{aligned}$$

which implies that there exist three positive constants \(C_1, C_2\) and \(C_3\) such that

$$\begin{aligned} \Vert \Psi (\mathbf{m})(t)\Vert _{L^p(\Omega )}&\le C_1\Vert \mathbf{m}_0\Vert _{L^p(\Omega )} +C_2\Vert \nabla \psi \Vert _{L^\infty \left( [0,T], L^6(\Omega )\right) }^2(R+1)T^{1-\frac{n}{6}}\nonumber \\&\quad +C_3\big (R+1\big )^{2\gamma -1} T^{1-\alpha (2\gamma -1)}. \end{aligned}$$
(5.3)

Similarly, we have

$$\begin{aligned} t^\alpha \Vert \Psi (\mathbf{m})(t)\Vert _{L^\beta (\Omega )}&\le Ct^\alpha t^{-\frac{n}{2}(\frac{1}{p}-\frac{1}{\beta })} \Vert \mathbf{m}_0\Vert _{L^p(\Omega )}\\&\quad +Ct^\alpha \int _0^t(t-s)^{-\frac{n}{6}} \Vert (\mathbf{m}\cdot \nabla \psi )\nabla \psi \Vert _{L^{\frac{3\beta }{3+\beta }}(\Omega )}ds\\&\quad +Ct^\alpha \int _0^t(t-s)^{-\alpha }\big \Vert |\mathbf{m}|^{2(\gamma -1)} \mathbf{m}\big \Vert _{L^p(\Omega )}ds\\&\le C\Vert \mathbf{m}_0\Vert _{L^p(\Omega )}+Ct^\alpha \int _0^t\big ((t-s)^{-\frac{n}{6}} \Vert \nabla \psi \Vert _{L^6(\Omega )}^2\Vert \mathbf{m}\Vert _{L^\beta (\Omega )}\\&\quad +(t-s)^{-\alpha }\Vert \mathbf{m}\Vert _{L^\beta (\Omega )}^{2\gamma -1}\big )ds\\&\le C\Vert \mathbf{m}_0\Vert _{L^p(\Omega )}+C\Vert \nabla \psi \Vert _{L^\infty \left( [0,T], L^6(\Omega )\right) }^2(R+1)t^\alpha \int _0^t(t-s)^{-\frac{n}{6}}s^{-\alpha }ds\\&\quad +C\big (R+1\big )^{2\gamma -1}t^\alpha \int _0^t(t-s)^{-\alpha }s^{-\alpha (2\gamma -1)}ds, \end{aligned}$$

where we use the fact \(\beta \ge 2\) due to \(p\ge 2\) and \(\gamma >1\). Hence, there exist three positive constants \(C_4, C_5\) and \(C_6\) such that

$$\begin{aligned} t^\alpha \Vert \Psi (\mathbf{m})(t)\Vert _{L^\beta (\Omega )}&\le C_4\Vert \mathbf{m}_0\Vert _{L^p(\Omega )}+C_5\Vert \nabla \psi \Vert _{L^\infty \left( [0,T], L^6(\Omega )\right) }^2(R+1)T^{1-\frac{n}{6}}\nonumber \\&\quad +C_6\big (R+1\big )^{2\gamma -1}T^{1-\alpha (2\gamma -1)}. \end{aligned}$$
(5.4)

By taking T suitable small, we get that

$$\begin{aligned} C_2\Vert \nabla \psi \Vert _{L^\infty \left( [0,T], L^6(\Omega )\right) }^2(R+1)T^{1-\frac{n}{6}} +C_3\big (R+1\big )^{2\gamma -1}T^{1-\alpha (2\gamma -1)}<1,\\ C_5\Vert \nabla \psi \Vert _{L^\infty \left( [0,T], L^6(\Omega )\right) }^2(R+1)T^{1-\frac{n}{6}} +C_6\big (R+1\big )^{2\gamma -1}T^{1-\alpha (2\gamma -1)}<1. \end{aligned}$$

Hence, we infer from (5.3) and (5.4) that \(\Psi \) maps \(X_R\) into itself when T is suitable small.

We now turn to the proof of the contractivity of the mapping \(\Psi \). For any given \(\mathbf{m}^1, \mathbf{m}^2\in X_R\), we let \(\delta \mathbf{m}:=\mathbf{m}^1-\mathbf{m}^2\) and \(\delta \Psi (\mathbf{m}):=\Psi (\mathbf{m}^1)-\Psi (\mathbf{m}^2)\). The key is to estimate \(t^\alpha \Vert \delta \Psi (\mathbf{m})(t)\Vert _{L^\beta (\Omega )}\). In fact, we have

$$\begin{aligned}&t^\alpha \Vert \delta \Psi (\mathbf{m})(t)\Vert _{L^\beta (\Omega )}\\&\quad \le Ct^\alpha \int _0^t(t-s)^{-\frac{n}{6}}\Vert (\delta \mathbf{m}\cdot \nabla \psi ) \nabla \psi \Vert _{L^{\frac{3\beta }{3+\beta }}(\Omega )}ds\\&\qquad +\,Ct^\alpha \int _0^t(t-s)^{-\alpha }\big \Vert \big (|\mathbf{m}^1|^{2(\gamma -1)} \mathbf{m}^1-|\mathbf{m}^2|^{2(\gamma -1)}\mathbf{m}^2\big )(s)\big \Vert _{L^p(\Omega )}ds, \end{aligned}$$

where we have used the fact \(\Vert \delta \mathbf{m}_0\Vert _{L^p(\Omega )}=0\). Meanwhile, noticing the elementary inequality

$$\begin{aligned} \big ||x|^\kappa x-|y|^\kappa y\big |\le C\big (|x|+|y|\big )^\kappa |x-y| \le C(|x|^\kappa +|y|^\kappa )|x-y| \end{aligned}$$
(5.5)

for all \(\kappa \ge 0\) and \(x, y\in {\mathbb {R}}^n\), we have

$$\begin{aligned} t^\alpha \Vert \delta \Psi (\mathbf{m})(t)\Vert _{L^\beta (\Omega )}\le & {} Ct^\alpha \int _0^t(t-s)^{-\frac{n}{6}}\Vert \delta \mathbf{m}\Vert _{L^\beta (\Omega )} \Vert \nabla \psi \Vert _{L^6(\Omega )}^2ds\\&+\,Ct^\alpha \int _0^t(t-s)^{-\alpha }\big \Vert \big (|\mathbf{m}^1|^{2(\gamma -1)} +|\mathbf{m}^2|^{2(\gamma -1)}\big )\delta \mathbf{m}\big \Vert _{L^p(\Omega )}ds\\\le & {} C\Vert \nabla \psi \Vert _{L^\infty \left( [0,T], L^6(\Omega )\right) }^2t^\alpha d(\mathbf{m}^1,\mathbf{m}^2)\int _0^t(t-s)^{-\frac{n}{6}}s^{-\alpha }ds\\&+\,Ct^\alpha \int _0^t(t-s)^{-\alpha } \big (\Vert \mathbf{m}^1\Vert _{L^\beta (\Omega )}^{2(\gamma -1)} +\Vert \mathbf{m}^2\Vert _{L^\beta (\Omega )}^{2(\gamma -1)}\big )\Vert \delta \mathbf{m}\Vert _{L^\beta (\Omega )}ds\\\le & {} C\Vert \nabla \psi \Vert _{L^\infty \left( [0,T], L^6(\Omega )\right) }^2T^{1-\frac{n}{6}}d(\mathbf{m}^1,\mathbf{m}^2)\\&+\,Ct^\alpha \big (R+1\big )^{2(\gamma -1)}\int _0^t(t-s)^{-\alpha } s^{-2(\gamma -1)\alpha }\Vert \delta \mathbf{m}\Vert _{L^\beta (\Omega )}ds, \end{aligned}$$

which implies that there exists a positive constant \(C_7\) such that

$$\begin{aligned} t^\alpha \Vert \delta \Psi (\mathbf{m})(t)\Vert _{L^\beta (\Omega )}\le & {} C_7\Big (\Vert \nabla \psi \Vert _{L^\infty \left( [0,T], L^6(\Omega )\right) }^2T^{1-\frac{n}{6}}\\&+\big (R+1\big )^{2(\gamma -1)} T^{1-\alpha (2\gamma -1)}\Big )d(\mathbf{m}^1,\mathbf{m}^2). \end{aligned}$$

From this, we can assert that, if T (depending on R) is small enough such that

$$\begin{aligned} C_7\Big (\Vert \nabla \psi \Vert _{L^\infty \left( [0,T], L^6(\Omega )\right) }^2T^{1-\frac{n}{6}}+\big (R+1\big )^{2(\gamma -1)} T^{1-\alpha (2\gamma -1)}\Big )<1, \end{aligned}$$

then \(\Psi \) is a strict contraction mapping from \(X_R\) to \(X_R\). Thus \(\Psi \) has a unique fixed point \(\mathbf{m}\) in \(X_R\). Moreover, \(\mathbf{m}\) is a solution of (2.8).

We claim that there exists \(\theta \in (0,1)\) such that \(\mathbf{m}\in C^{\theta ,\frac{\theta }{2}}(\Omega _T)\). To this end, it suffices to verify that \(\mathbf{m}\in C\big ([0,T], L^p(\Omega )\big )\cap L^\infty _{\text {loc}}\big ((0,T), L^\infty (\Omega )\big )\) by applying [22, Theorem 3.10.1] to the Dirichlet problem associated with the linear heat equation with a potential

$$\begin{aligned} \mathbf{m}_t-\Delta \mathbf{m}+\mathbf{A}_\gamma (t,x)\mathbf{m}=0\quad \mathrm {with}\quad \mathbf{A}_\gamma :=|\mathbf{m}|^{2(\gamma -1)}I-\nabla p\otimes \nabla p. \end{aligned}$$

In fact, due to \(\mathbf{m}\in L^\infty _{\text {loc}}((0,T),L^\beta (\Omega ))\), we can apply Theorem A1 in [3] with \(r=\beta \) and \(\sigma =\frac{\beta }{2(\gamma -1)}\) on the interval \((\epsilon ,T-\epsilon )\) for any \(\epsilon \in (0,\frac{T}{2})\), which also implies

$$\begin{aligned} \mathbf{m}\in C\big ([0,T], L^p(\Omega )\big )\cap L^\infty _{\text {loc}}\big ((0,T), L^\infty (\Omega )\big ). \end{aligned}$$

This completes the proof of Lemma 2.4. \(\square \)

Proof of Lemma 3.1

Here we will omit the superscript \(\varepsilon \) for notational convenience. Without loss of generality, we just consider the case \(\Vert S\Vert _{C^\theta (\Omega )}\ne 0\), since the case \(\Vert S\Vert _{C^\theta (\Omega )}=0\) can be similarly dealt with. Let \(T\in (0,1)\) to be specified below. For simplicity, we also set

$$\begin{aligned} E:=C\big ([0,T]; L^2(\Omega )\big ). \end{aligned}$$

For any \(R>0\), we will denote the closed bounded convex subset \(E_R\) of E by

$$\begin{aligned} E_R:=\big \{g\in E: \Vert g(\cdot ,t)\Vert _{L^2(\Omega )}\le R+1 \quad \text {for all}\ t\in [0,T]\big \}. \end{aligned}$$

We first fix an \(R\ge C_1\Vert \varrho _0\Vert _{L^2(\Omega )}\), where \(C_1\) is given in (5.10). Then for any given \({\bar{\varrho }}\in E_R\), we see from the standard elliptic theory [10] that the boundary value problem

$$\begin{aligned} \left\{ \begin{aligned}&-\Delta \psi ={\bar{\varrho }}, \quad x\in \Omega ,\\&\psi (x,t)=0, \quad x\in \partial \Omega \end{aligned} \right. \end{aligned}$$
(5.6)

possesses a unique solution \({\bar{\psi }}=\psi [{\bar{\varrho }}](\cdot , t)\in H^1_0(\Omega )\cap H^2(\Omega )\) satisfying

$$\begin{aligned} \Vert {\bar{\psi }}\Vert _{L^\infty ([0,T]; H^2(\Omega ))}\le C\Vert {\bar{\varrho }}\Vert _{L^\infty ([0,T];L^2(\Omega ))}. \end{aligned}$$
(5.7)

Invoking this, we can get from Lemma 2.4 that, if \(\mathbf{m}_0\in L^\infty (\Omega )\), then the initial-boundary value problem

$$\begin{aligned} \left\{ \begin{aligned}&\partial _t\mathbf{m}-\Delta \mathbf{m}-\big (\mathbf{m}\cdot \nabla \psi [{\bar{\varrho }}\big )\nabla \psi [{\bar{\varrho }}] +|\mathbf{m}|^{2(\gamma -1)}\mathbf{m}=0, \&x\in \Omega ,\\&\mathbf{m}(x,t)=\mathbf{0}, \qquad&x\in \partial \Omega ,\\&\mathbf{m}(x,0)=\mathbf{m}_0(x), \qquad&x\in \Omega \end{aligned} \right. \end{aligned}$$
(5.8)

admits a unique solution \({\bar{\mathbf{m}}}{=}\mathbf{m}[\psi [{\bar{\varrho }}]]\) belonging to \(C([0,T];L^\infty (\Omega ))\cap C^{\theta ,\frac{\theta }{2}}(\Omega _T)\).

Thus, for any \({\bar{\varrho }}\in E_R\) we can introduce a mapping \(\Psi \) by setting \(\Psi ({\bar{\varrho }})=\varrho \), where \(\varrho \) is a solution of the initial-boundary value problem

$$\begin{aligned} \left\{ \begin{aligned}&\partial _t\varrho =\text {div}\left( (I+{\bar{\mathbf{m}}}\otimes {\bar{\mathbf{m}}})(\nabla \varrho +\varrho \nabla {\bar{\psi }})\right) +S\varrho , \quad&x\in \Omega ,\\&\varrho (x,t)=0, \qquad \quad&x\in \partial \Omega ,\\&\varrho (x,0)=\varrho _0(x), \qquad&x\in \Omega . \end{aligned} \right. \end{aligned}$$
(5.9)

Notice that the local existence of solution to (5.9) can be showed by utilizing the linear parabolic theory, where \({\bar{\psi }}=\psi [{\bar{\varrho }}](x,t)\) is determined by (5.6) and \({\bar{\mathbf{m}}}=\mathbf{m}[\psi [{\bar{\varrho }}]]\) is determined by (5.8). Moreover, \(\varrho \) satisfies

$$\begin{aligned} \Vert \varrho \Vert _{C([0,T]\times L^2(\Omega ))}+\Vert \varrho \Vert _{L^2([0,T];H^1(\Omega ))}\le C_1\Vert \rho _0\Vert _{L^2(\Omega )}, \end{aligned}$$
(5.10)

see Theorem 2.1 and Theorem 4.1 in Chapter 3 of [22] for details. Moreover, we see from [22, Theorem 3.7.1] that there exists a positive constant \(C=C(R)\) such that

$$\begin{aligned} \Vert \varrho \Vert _{L^\infty ([0,T]\times \Omega )}\le C\Vert j_\varepsilon *\varrho _0\Vert _{L^\infty ([0,T]\times \Omega )}. \end{aligned}$$
(5.11)

Invoking (5.10), it is clear that \(\Psi \) maps \(E_R\) into itself when T is suitable small.

Next, we shall certify that \(\Psi \) is contractive and thus has a fixed point in \(E_R\) provided that T is small enough. To this end, for any given \({\bar{\varrho }}^i\in E_R\), let \({\bar{\psi }}^i:=\psi [{\bar{\varrho }}^i]\) and \({\bar{\mathbf{m}}}^i:=\mathbf{m}[{\bar{\psi }}^i]\) be the solution of system (5.6) and (5.8) with \({\bar{\mathbf{m}}}={\bar{\mathbf{m}}}^i\) and \({\bar{\psi }}={\bar{\psi }}^i\), \(i=1,2\), respectively. For simplicity, we also set

$$\begin{aligned} \delta {\bar{\mathbf{m}}}:= & {} {\bar{\mathbf{m}}}^1-{\bar{\mathbf{m}}}^2, \qquad \delta {\bar{\psi }}:={\bar{\psi }}^1-{\bar{\psi }}^2, \qquad \delta \varrho =\varrho ^1-\varrho ^2 \qquad \mathrm {and}\qquad \\ \delta {\bar{\varrho }}:= & {} {\bar{\varrho }}^1-{\bar{\varrho }}^2, \end{aligned}$$

where \(\varrho ^i:=\Psi ({\bar{\varrho }}^i)\), \(i=1,2\). The key of the proof is to estimate \(\Vert \delta \varrho (t)\Vert _{L^2(\Omega )}\). To achieve this, we notice that \(\delta \varrho \) satisfies the initial-boundary value problem

$$\begin{aligned} \left\{ \begin{aligned}&\partial _t\delta \varrho =\Delta \delta \varrho +\text {div}\big (({\bar{\mathbf{m}}}^1\otimes {\bar{\mathbf{m}}}^1) \nabla \varrho ^1-({\bar{\mathbf{m}}}^2\otimes {\bar{\mathbf{m}}}^2)\nabla \varrho ^2\big )\\&+\text {div}\big (\varrho ^1\nabla {\bar{\psi }}^1-\varrho ^2\nabla {\bar{\psi }}^2\big ) +\text {div}\big (({\bar{\mathbf{m}}}^1\otimes {\bar{\mathbf{m}}}^1) \nabla {\bar{\psi }}^1\varrho ^1-({\bar{\mathbf{m}}}^2\otimes {\bar{\mathbf{m}}}^2) \nabla {\bar{\psi }}^2\varrho ^2\big )+\delta \varrho S,&x\in \Omega , \\&\delta \varrho =0, \quad&x\in \partial \Omega ,\\&\delta \varrho (x,0)=0, \quad&x\in \Omega . \end{aligned} \right. \nonumber \\ \end{aligned}$$
(5.12)

First of all, multiplication of (5.12) by \(\delta \varrho \) and integration by parts give us

$$\begin{aligned} \frac{1}{2}\frac{d}{dt}\int _{\Omega }|\delta \varrho |^2dx +\int _\Omega |\nabla \delta \varrho |^2dx=&-\int _\Omega \big (({\bar{\mathbf{m}}}^1\otimes {\bar{\mathbf{m}}}^1)\nabla \varrho ^1-({\bar{\mathbf{m}}}^2 \otimes {\bar{\mathbf{m}}}^2)\nabla \varrho ^2\big ) \cdot \nabla \delta \varrho dx \nonumber \\&- \int _\Omega \big (\varrho ^1\nabla {\bar{\psi }}^1-\varrho ^2\nabla {\bar{\psi }}^2\big ) \cdot \nabla \delta \varrho dx \nonumber \\&-\int _\Omega \big (({\bar{\mathbf{m}}}^1\otimes {\bar{\mathbf{m}}}^1) \nabla {\bar{\psi }}^1\varrho ^1-({\bar{\mathbf{m}}}^2\otimes {\bar{\mathbf{m}}}^2) \nabla {\bar{\psi }}^2\varrho ^2\big )\cdot \nabla \delta \varrho dx\nonumber \\&+\int _\Omega |\delta \varrho |^2Sdx \nonumber \\ =:&\;\Pi _1+\Pi _2+\Pi _3+\Pi _4. \end{aligned}$$
(5.13)

For \(\Pi _1\), a careful calculation shows that

$$\begin{aligned} \Pi _1 =&-\int _\Omega \big [({\bar{\mathbf{m}}}^1\otimes {\bar{\mathbf{m}}}^1) \nabla \delta \varrho +\big (\delta {\bar{\mathbf{m}}}\otimes {\bar{\mathbf{m}}}^1\big ) \nabla \varrho ^2+\big ({\bar{\mathbf{m}}}^2\otimes (\delta {\bar{\mathbf{m}}})\big ) \nabla \varrho ^2\big ]\cdot \nabla \delta \varrho dx\nonumber \\ \le&-\int _\Omega \big ({\bar{\mathbf{m}}}^1\cdot \nabla \delta \varrho \big )^2dx+\Vert \delta {\bar{\mathbf{m}}}\Vert _{L^\infty (\Omega )} \Vert \nabla \varrho ^2\Vert _{L^2(\Omega )}\Vert \nabla \delta \varrho \Vert _{L^2(\Omega )}\nonumber \\&\big (\Vert {\bar{\mathbf{m}}}^1\Vert _{L^\infty (\Omega )} +\Vert {\bar{\mathbf{m}}}^2\Vert _{L^\infty (\Omega )}\big )\nonumber \\ \le&-\Vert {\bar{\mathbf{m}}}^1\cdot \nabla \delta \varrho \Vert _{L^2(\Omega )}^2 +\frac{1}{8}\Vert \nabla \delta \varrho \Vert _{L^2(\Omega )}^2 +C\Vert \delta {\bar{\mathbf{m}}}\Vert _{L^\infty (\Omega )}^2\Vert \nabla \varrho ^2 \Vert _{L^2(\Omega )}^2\nonumber \\&\big (\Vert {\bar{\mathbf{m}}}^1\Vert _{L^\infty (\Omega )}^2 +\Vert {\bar{\mathbf{m}}}^2\Vert _{L^\infty (\Omega )}^2\big ). \end{aligned}$$
(5.14)

Similarly, \(\Pi _2\) can be controlled as follows:

$$\begin{aligned} \Pi _2&=-\int _\Omega \delta \varrho \nabla {\bar{\psi }}^1\cdot \nabla \delta \varrho dx-\int _\Omega \varrho ^2\nabla \delta {\bar{\psi }}\cdot \nabla \delta \varrho dx\nonumber \\&=-\frac{1}{2}\int _\Omega {\bar{\rho }}^1(\delta \varrho )^2dx -\int _\Omega \varrho ^2\nabla \delta {\bar{\psi }}\cdot \nabla \delta \varrho dx\nonumber \\&\le -\frac{1}{2}\int _\Omega {\bar{\rho }}^1(\delta \varrho )^2dx +\Vert \rho ^2\Vert _{L^3(\Omega )}\Vert \nabla \delta {\bar{\psi }}\Vert _{L^6(\Omega )} \Vert \nabla \delta \varrho \Vert _{L^2(\Omega )}. \end{aligned}$$

Moreover, employing Sobolev’s embedding \(H^1(\Omega )\hookrightarrow L^6(\Omega )\) and Poincaré’s inequality, we have

$$\begin{aligned} \Pi _2\le -\frac{1}{2}\int _\Omega {\bar{\rho }}^1(\delta \varrho )^2dx +C\Vert \rho ^2\Vert _{L^3(\Omega )}\Vert \delta {\bar{\psi }}\Vert _{H^2(\Omega )} \Vert \nabla \delta \varrho \Vert _{L^2(\Omega )}. \end{aligned}$$
(5.15)

Here we need to estimate the factor \(\Vert \delta {\bar{\psi }}\Vert _{H^2(\Omega )}\). Note that \({\bar{\psi }}^i\in H^2(\Omega )\cap H^1_0(\Omega )\) is a solution of (5.6) and \(\delta {\bar{\psi }}\) is a solution to the boundary-value problem

$$\begin{aligned} \left\{ \begin{array}{ll} -\Delta \delta \psi =\delta {\bar{\varrho }}, \quad \quad &{} x\in \Omega ,\\ \delta \psi =0, \quad \quad &{} x\in \partial \Omega . \end{array} \right. \end{aligned}$$

Hence, it demonstrates that \(\delta {\bar{\psi }}\in H^2(\Omega )\cap H^1_0(\Omega )\) and

$$\begin{aligned} \Vert \delta {\bar{\psi }}\Vert _{H^2(\Omega )}\le C\Vert \delta {\bar{\varrho }}\Vert _{L^2(\Omega )}, \end{aligned}$$
(5.16)

see Theorem 6.3.4 in [9]. Substitute (5.16) into (5.15), to find:

$$\begin{aligned} \Pi _2&\le -\frac{1}{2}\int _\Omega {\bar{\rho }}^1(\delta \varrho )^2dx +C\Vert \rho ^2\Vert _{L^3(\Omega )}\Vert \delta {\bar{\varrho }}\Vert _{L^2(\Omega )} \Vert \nabla \delta \varrho \Vert _{L^2(\Omega )}\nonumber \\&\le \frac{1}{2}\Vert {\bar{\rho }}^1\Vert _{L^2(\Omega )}\Vert \delta \varrho \Vert _{L^4(\Omega )}^2+ \frac{1}{8}\Vert \nabla \delta \rho \Vert _{L^2(\Omega )}^2 +C\Vert \rho ^2\Vert _{L^3(\Omega )}^2\Vert \delta {\bar{\varrho }}\Vert _{L^2(\Omega )}^2. \end{aligned}$$
(5.17)

Next, by employing (5.7) and (5.16) we can estimate the term \(\Pi _3\) as follows:

$$\begin{aligned} \Pi _3&=-\int _\Omega \big ((\delta {\bar{\mathbf{m}}}\otimes {\bar{\mathbf{m}}}^1) \nabla {\bar{\psi }}^2\varrho ^2+({\bar{\mathbf{m}}}^2 \otimes \delta {\bar{\mathbf{m}}})\nabla {\bar{\psi }}^2\varrho ^2\big )\cdot \nabla \delta \varrho dx\nonumber \\&\quad -\int _\Omega ({\bar{\mathbf{m}}}^1\otimes {\bar{\mathbf{m}}}^1) \big (\nabla \delta {\bar{\psi }}\varrho ^1+\nabla {\bar{\psi }}^2\delta \varrho \big ) \cdot \nabla \delta \varrho dx\nonumber \\&\le \Vert \delta {\bar{\mathbf{m}}}\Vert _{L^\infty (\Omega )}\Vert \nabla {\bar{\psi }}^2 \Vert _{L^6(\Omega )}\Vert \varrho ^2\Vert _{L^3(\Omega )} \Vert \nabla \delta \varrho \Vert _{L^2(\Omega )}\big (\Vert {\bar{\mathbf{m}}}^1\Vert _{L^\infty (\Omega )} +\Vert {\bar{\mathbf{m}}}^2\Vert _{L^\infty (\Omega )}\big ) \nonumber \\&\quad +\Vert {\bar{\mathbf{m}}}^1\Vert _{L^\infty (\Omega )} \Vert {\bar{\mathbf{m}}}^1\cdot \nabla \delta \varrho \Vert _{L^2(\Omega )} \big (\Vert \nabla \delta {\bar{\psi }}\Vert _{L^6(\Omega )}\Vert \varrho ^1\Vert _{L^3(\Omega )} +\Vert \nabla {\bar{\psi }}^2\Vert _{L^6(\Omega )}\Vert \delta \rho \Vert _{L^3(\Omega )}\big ) \nonumber \\&\le C\Vert \delta {\bar{\mathbf{m}}}\Vert _{L^\infty (\Omega )}\Vert {\bar{\varrho }}^2 \Vert _{L^2(\Omega )}\Vert \varrho ^2\Vert _{L^3(\Omega )} \Vert \nabla \delta \varrho \Vert _{L^2(\Omega )}\big (\Vert {\bar{\mathbf{m}}}^1 \Vert _{L^\infty (\Omega )}+\Vert {\bar{\mathbf{m}}}^2\Vert _{L^\infty (\Omega )}\big ) \nonumber \\&\quad +C\Vert {\bar{\mathbf{m}}}^1\Vert _{L^\infty (\Omega )}\Vert {\bar{\mathbf{m}}}^1\cdot \nabla \delta \varrho \Vert _{L^2(\Omega )}\big (\Vert \delta {\bar{\varrho }} \Vert _{L^2(\Omega )}\Vert \varrho ^1\Vert _{L^3(\Omega )}+\Vert {\bar{\varrho }}^2 \Vert _{L^2(\Omega )}\Vert \delta \rho \Vert _{L^3(\Omega )}\big ) \nonumber \\&\le \frac{1}{8}\Vert \nabla \delta \varrho \Vert _{L^2(\Omega )}^2 +C\Vert \delta {\bar{\mathbf{m}}}\Vert _{L^\infty (\Omega )}^2\Vert {\bar{\varrho }}^2 \Vert _{L^2(\Omega )}^2\Vert \varrho ^2\Vert _{L^3(\Omega )}\big (\Vert {\bar{\mathbf{m}}}^1 \Vert _{L^\infty (\Omega )}^2+\Vert {\bar{\mathbf{m}}}^2\Vert _{L^\infty (\Omega )}^2\big ) \nonumber \\&\quad +\frac{1}{8}\Vert {\bar{\mathbf{m}}}^1\cdot \nabla \delta \varrho \Vert _{L^2(\Omega )}^2 +C\Vert {\bar{\mathbf{m}}}^1\Vert _{L^\infty (\Omega )}^2\big (\Vert \delta {\bar{\varrho }} \Vert _{L^2(\Omega )}^2\Vert \varrho ^1\Vert _{L^3(\Omega )}^2+\Vert {\bar{\varrho }}^2 \Vert _{L^2(\Omega )}^2\Vert \delta \rho \Vert _{L^3(\Omega )}^2\big ). \end{aligned}$$
(5.18)

Substituting (5.14), (5.17) and (5.18) into (5.13), we obtain that there exists a positive constant \(C_2=C_2(R)\) such that

$$\begin{aligned}&\frac{d}{dt}\Vert \delta \varrho \Vert _{L^2(\Omega )}^2 +\Vert \nabla \delta \varrho \Vert _{L^2(\Omega )}^2\nonumber \\&\le C_2\Big (\Vert \delta {\bar{\mathbf{m}}}\Vert _{L^\infty (\Omega )}^2 \Vert \nabla \varrho ^2\Vert _{L^2(\Omega )}^2 +\Vert \delta \varrho \Vert _{L^4(\Omega )}^2{+}\big (\Vert \rho ^1\Vert _{L^3(\Omega )}^2 +\Vert \rho ^2\Vert _{L^3(\Omega )}^2\big )\Vert \delta {\bar{\varrho }}\Vert _{L^2(\Omega )}^2 \nonumber \\&\quad +\Vert \delta {\bar{\mathbf{m}}}\Vert _{L^\infty (\Omega )}^2\Vert \varrho ^2\Vert _{L^2(\Omega )} +\Vert \delta \rho \Vert _{L^3(\Omega )}^2\big (1+\Vert S\Vert _{L^3(\Omega )}\big )\Big ). \end{aligned}$$
(5.19)

It follows from Sobolev’s embedding and Poincaré’s inequality that

$$\begin{aligned} \Vert \delta \rho \Vert _{L^3(\Omega )}\le&\Vert \delta \rho \Vert _{L^2(\Omega )}^{\frac{1}{2}} \Vert \delta \varrho \Vert _{L^6(\Omega )}^{\frac{1}{2}}\le C\Vert \delta \rho \Vert _{L^2(\Omega )}^{\frac{1}{2}}\Vert \delta \rho \Vert _{H^1(\Omega )}^{\frac{1}{2}}\le C\Vert \delta \rho \Vert _{L^2(\Omega )}^{\frac{1}{2}} \Vert \nabla \delta \rho \Vert _{L^2(\Omega )}^{\frac{1}{2}};\\ \Vert \delta \rho \Vert _{L^4(\Omega )}\le&\Vert \delta \rho \Vert _{L^2(\Omega )}^{\frac{1}{4}} \Vert \delta \varrho \Vert _{L^6(\Omega )}^{\frac{3}{4}}\le C\Vert \delta \rho \Vert _{L^2(\Omega )}^{\frac{1}{4}}\Vert \delta \rho \Vert _{H^1(\Omega )}^{\frac{3}{4}} \le C\Vert \delta \rho \Vert _{L^2(\Omega )}^{\frac{1}{4}}\Vert \nabla \delta \rho \Vert _{L^2(\Omega )}^{\frac{3}{4}}. \end{aligned}$$

Invoking this, we get from (5.19) and Young’s inequality that there is a positive constant C such that

$$\begin{aligned}&\frac{d}{dt}\Vert \delta \varrho \Vert _{L^2(\Omega )}^2+\Vert \nabla \delta \varrho \Vert _{L^2(\Omega )}^2\\&\quad \le C\Big (\Vert \delta {\bar{\mathbf{m}}}\Vert _{L^\infty (\Omega )}^2 \Vert \nabla \varrho ^2\Vert _{L^2(\Omega )}^2\\&\qquad +\Vert \delta \rho \Vert _{L^2(\Omega )}^{\frac{1}{2}}\Vert \nabla \delta \rho \Vert _{L^2(\Omega )}^{\frac{3}{2}} +\big (\Vert \rho ^1\Vert _{L^3(\Omega )}^2+\Vert \rho ^2\Vert _{L^3(\Omega )}^2\big ) \Vert \delta {\bar{\varrho }}\Vert _{L^2(\Omega )}^2\\&\qquad +\Vert \delta {\bar{\mathbf{m}}}\Vert _{L^\infty (\Omega )}^2\Vert \varrho ^2\Vert _{L^2(\Omega )} +\Vert \delta \rho \Vert _{L^2(\Omega )}\Vert \nabla \delta \rho \Vert _{L^2(\Omega )} \big (1+\Vert S\Vert _{L^3(\Omega )}\big )\Big )\\&\quad \le \frac{1}{2}\Vert \nabla \delta \rho \Vert _{L^2(\Omega )}^2+ C\Big (\Vert \delta {\bar{\mathbf{m}}}\Vert _{L^\infty (\Omega )}^2 \Vert \nabla \varrho ^2\Vert _{L^2(\Omega )}^2+\big (\Vert \rho ^1\Vert _{L^3(\Omega )}^2 +\Vert \rho ^2\Vert _{L^3(\Omega )}^2\big )\Vert \delta {\bar{\varrho }}\Vert _{L^2(\Omega )}^2\\&\qquad +\Vert \delta {\bar{\mathbf{m}}}\Vert _{L^\infty (\Omega )}^2\Vert \varrho ^2\Vert _{L^2(\Omega )} +\Vert \delta \rho \Vert _{L^2(\Omega )}^2\big (1+\Vert S\Vert _{L^3(\Omega )}^2\big )\Big ). \end{aligned}$$

By employing Hölder’s inequality and (5.11), it arrives at

$$\begin{aligned} \frac{d}{dt}\Vert \delta \varrho \Vert _{L^2(\Omega )}^2&\le C\Big (\Vert \delta {\bar{\mathbf{m}}}\Vert _{L^\infty (\Omega )}^2\Vert \nabla \varrho ^2\Vert _{L^2(\Omega )}^2 +\big (\Vert \rho ^1\Vert _{L^3(\Omega )}^2+\Vert \rho ^2\Vert _{L^3(\Omega )}^2\big ) \Vert \delta {\bar{\varrho }}\Vert _{L^2(\Omega )}^2\nonumber \\&\quad +\Vert \delta {\bar{\mathbf{m}}}\Vert _{L^\infty (\Omega )}^2\Vert \varrho ^2\Vert _{L^2(\Omega )} +\Vert \delta \rho \Vert _{L^2(\Omega )}^2\big (1+\Vert S\Vert _{L^3(\Omega )}^2\big )\Big )\nonumber \\&\le C\Big (\Vert \delta {\bar{\mathbf{m}}}\Vert _{L^\infty (\Omega )}^2 \big (1+\Vert \nabla \varrho ^2\Vert _{L^2(\Omega )}^2\big ) +\Vert \delta {\bar{\varrho }}\Vert _{L^2(\Omega )}^2\nonumber \\&\quad +\Vert \delta \rho \Vert _{L^2(\Omega )}^2\big (1+\Vert S\Vert _{L^3(\Omega )}^2\big )\Big ). \end{aligned}$$
(5.20)

Here we need to estimate the factor \(\Vert \delta {\bar{\mathbf{m}}}\Vert _{L^\infty (\Omega )}\).

On the other hand, \(\delta {\bar{\mathbf{m}}}\) satisfies the initial-boundary value problem

$$\begin{aligned} \left\{ \begin{aligned}&\partial _t\delta {\bar{\mathbf{m}}}-\Delta \delta {\bar{\mathbf{m}}} +\big (|{\bar{\mathbf{m}}}^1|^{2(\gamma -1)}{\bar{\mathbf{m}}}^1-|{\bar{\mathbf{m}}}^2|^{2(\gamma -1)} {\bar{\mathbf{m}}}^2\big )\\&=\big (({\bar{\mathbf{m}}}^1\cdot \nabla {\bar{\psi }}^1)\nabla {\bar{\psi }}^1 -({\bar{\mathbf{m}}}^2\cdot \nabla {\bar{\psi }}^2)\nabla {\bar{\psi }}^2\big ),\&(x,t)\in \Omega _T,\\&\delta {\bar{\mathbf{m}}}(x,t)=\mathbf{0},\&(x,t)\in \Sigma _T, \\&\delta {\bar{\mathbf{m}}}(x,0)=\mathbf{0},\&x\in \Omega , \end{aligned} \right. \end{aligned}$$
(5.21)

where \({\bar{\psi }}^i\in L^\infty \big ([0,T];H^1_0(\Omega )\cap H^2(\Omega )\big )\) is a solution of (5.6) and \({\bar{\mathbf{m}}}^i\in L^\infty ([0,T]\times \Omega )\cap C^{\theta ,\frac{\theta }{2}}(\Omega _T)\) is a solution of (5.8). Similar to the proof of Lemma 2.4, due to \({\bar{\mathbf{m}}}^i\in L^\infty ([0,T]\times \Omega )\), we obtain from (5.2), (5.5) and (5.7) that

$$\begin{aligned} \Vert \delta {\bar{\mathbf{m}}}\Vert _{L^\infty (\Omega )}&\le C \int _0^t(t-s)^{-\frac{n}{6}}\Vert ({\bar{\mathbf{m}}}^1\cdot \nabla {\bar{\psi }}^1)\nabla {\bar{\psi }}^1 -({\bar{\mathbf{m}}}^2\cdot \nabla {\bar{\psi }}^2) \nabla {\bar{\psi }}^2\Vert _{L^3(\Omega )}ds\\&\quad +C\int _0^t\big \Vert |{\bar{\mathbf{m}}}^1|^{2(\gamma -1)} {\bar{\mathbf{m}}}^1-|{\bar{\mathbf{m}}}^2|^{2(\gamma -1)}{\bar{\mathbf{m}}}^2 \big \Vert _{L^\infty (\Omega )}ds\\&\le C \int _0^t(t-s)^{-\frac{n}{6}}\Vert (\delta {\bar{\mathbf{m}}} \cdot \nabla {\bar{\psi }}^1)\nabla {\bar{\psi }}^1\\&\quad +({\bar{\mathbf{m}}}^2\cdot \nabla \delta {\bar{\psi }})\nabla {\bar{\psi }}^1 +({\bar{\mathbf{m}}}^2\cdot \nabla {\bar{\psi }}^2) \nabla \delta {\bar{\psi }}\Vert _{L^3(\Omega )}ds\\&\quad +C\int _0^t\big (\Vert {\bar{\mathbf{m}}}^1\Vert _{L^\infty (\Omega )}^{2(\gamma -1)} +\Vert {\bar{\mathbf{m}}}^2\Vert _{L^\infty (\Omega )}^{2(\gamma -1)}\big ) \Vert \delta {\bar{\mathbf{m}}}\big \Vert _{L^\infty (\Omega )}ds \\&\le C\int _0^t(t-s)^{-\frac{n}{6}}\Vert \delta {\bar{\mathbf{m}}} \Vert _{L^\infty (\Omega )}\Vert \nabla {\bar{\psi }}^1\Vert _{L^6(\Omega )}^2ds\\&\quad +C\int _0^t(t-s)^{-\frac{n}{6}}\Vert {\bar{\mathbf{m}}}^2 \Vert _{L^\infty (\Omega )}\Vert \nabla \delta {\bar{\psi }}\Vert _{L^6(\Omega )} \big (\Vert \nabla {\bar{\psi }}^1\Vert _{L^6(\Omega )}+\Vert \nabla {\bar{\psi }}^2\Vert _{L^6(\Omega )}\big ) ds\\&\quad +C\int _0^t\big (\Vert {\bar{\mathbf{m}}}^1\Vert _{L^\infty (\Omega )}^{2(\gamma -1)} +\Vert {\bar{\mathbf{m}}}^2\Vert _{L^\infty (\Omega )}^{2(\gamma -1)}\big ) \Vert \delta {\bar{\mathbf{m}}}\big \Vert _{L^\infty (\Omega )}ds \\&\le C\int _0^t(t-s)^{-\frac{n}{6}}\Vert \delta {\bar{\mathbf{m}}} \Vert _{L^\infty (\Omega )}\Vert {\bar{\psi }}^1\Vert _{H^2(\Omega )}^2ds \\&\quad +C\int _0^t(t-s)^{-\frac{n}{6}}\Vert {\bar{\mathbf{m}}}^2\Vert _{L^\infty (\Omega )} \Vert \delta {\bar{\psi }}\Vert _{H^2(\Omega )}\big (\Vert {\bar{\psi }}^1\Vert _{H^2(\Omega )} +\Vert {\bar{\psi }}^2\Vert _{H^2(\Omega )}\big ) ds \\&\quad +C\int _0^t\big (\Vert {\bar{\mathbf{m}}}^1\Vert _{L^\infty (\Omega )}^{2(\gamma -1)} +\Vert {\bar{\mathbf{m}}}^2\Vert _{L^\infty (\Omega )}^{2(\gamma -1)}\big ) \Vert \delta {\bar{\mathbf{m}}}\big \Vert _{L^\infty (\Omega )}ds \\&\le C\int _0^t(t-s)^{-\frac{n}{6}}\Vert \delta {\bar{\mathbf{m}}}\Vert _{L^\infty (\Omega )}ds+ CT^{1-\frac{n}{6}}\Vert \delta {\bar{\psi }}\Vert _{L^\infty ([0,T];H^2(\Omega ))}\\&\quad +C\int _0^t\Vert \delta {\bar{\mathbf{m}}}\big \Vert _{L^\infty (\Omega )}ds. \end{aligned}$$

We can pick T suitable small such that \((t-s)^{-\frac{n}{6}}\ge t^{-\frac{n}{6}}\ge T^{-\frac{n}{6}}>1\). Hence there exist two positive constant \(C_4\) and \(C_5\) such that

$$\begin{aligned} \Vert \delta {\bar{\mathbf{m}}}\Vert _{L^\infty (\Omega )} \le C_4T^{1-\frac{n}{6}}\Vert \delta {\bar{\psi }}\Vert _{L^\infty ([0,T];H^2(\Omega ))} +C_5\int _0^t(t-s)^{-\frac{n}{6}}\Vert \delta {\bar{\mathbf{m}}}\big \Vert _{L^\infty (\Omega )}ds. \end{aligned}$$
(5.22)

Thus, a singular Gronwall’s inequality, see Lemma 2.3, gives

$$\begin{aligned} \Vert \delta {\bar{\mathbf{m}}}\Vert _{L^\infty (\Omega )}\le C_4T^{1-\frac{n}{6}}\Vert \delta {\bar{\psi }}\Vert _{L^\infty ([0,T];H^2(\Omega ))} \big (1+C_5C_6T^{1-\frac{n}{6}}e^{(1+\epsilon )\mu T}\big ):=M_0 \end{aligned}$$
(5.23)

for some \(\epsilon >0\) and \(C_6>0\), where \(\mu =\big (C_5\Gamma (1-\frac{n}{6})\big )^{\frac{6}{6-n}}\).

Then returning to (5.20), we conclude from (5.10) and (5.23) that

$$\begin{aligned} \Vert \delta \varrho \Vert _{L^2(\Omega )}^2&\le CM_0^2\int _0^t\Big (1+\Vert \nabla \varrho ^2\Vert _{L^2(\Omega )}^2\Big )ds\nonumber \\&\quad +C\int _0^t\Big (\big (1+\Vert S\Vert _{L^3(\Omega )}^2\big ) \Vert \delta \varrho \Vert _{L^2(\Omega )}^2+\Vert \delta {\bar{\varrho }}\Vert _{L^2(\Omega )}^2\Big )ds \nonumber \\&\le CM_0^2(1+T)+C\Vert \delta {\bar{\varrho }}\Vert _{L^\infty ([0,T]; L^2(\Omega ))}^2T\nonumber \\&\quad +C\big (1+\Vert S\Vert _{L^3(\Omega )}^2\big )\int _0^t \Vert \delta \varrho \Vert _{L^2(\Omega )}^2ds. \end{aligned}$$
(5.24)

Thus, Gronwall’s inequality gives that there are two positive constants \(C_7\) and \(C_8\) such that

$$\begin{aligned} \Vert \delta \varrho \Vert _{L^2(\Omega )}^2\le C_7 \Big (M_0^2(1+T)+\Vert \delta {\bar{\varrho }}\Vert _{L^\infty ([0,T];L^2(\Omega ))}^2T\Big ) e^{C_8\big (1+\Vert S\Vert _{L^3(\Omega )}^2\big )T}. \end{aligned}$$
(5.25)

Moreover, we can infer from (5.16), (5.23) and (5.25) that there exists a positive constant \(C_9\) such that

$$\begin{aligned} \Vert \delta \varrho \Vert _{L^2(\Omega )}^2&\le C_9\Big [T^{2-\frac{n}{3}}\big (1+T^{1-\frac{n}{6}}e^{(1+\epsilon )\mu T}\big )^2(1+T)+T\Big ]\nonumber \\&\qquad e^{C_8\big (1+\Vert S\Vert _{L^3(\Omega )}^2\big )T} \Vert \delta {\bar{\varrho }}\Vert _{L^\infty ([0,T];L^2(\Omega ))}^2. \end{aligned}$$
(5.26)

We can pick T suitable small such that

$$\begin{aligned} \Big [T^{2-\frac{n}{3}}\big (1+C_8T^{1-\frac{n}{6}}e^{(1+\epsilon )\mu T}\big )^2(1+T)+T\Big ]e^{C\big (1+\Vert S\Vert _{L^3(\Omega )}^2\big )T}<\frac{1}{C_9}, \end{aligned}$$

and, thus, get from (5.26) that there exists \(\kappa \in (0,1)\) such that

$$\begin{aligned} \Vert \delta \varrho \Vert _{L^2(\Omega )}\le \kappa \Vert \delta {\bar{\varrho }}\Vert _{L^\infty ([0,T];L^2(\Omega ))}, \end{aligned}$$

which implies that \(\Psi \) is a contraction mapping from \(E_R\) to \(E_R\) and thus has a fixed point \(\varrho \) in \(E_R\). Hence, \((\varrho ,\psi ,\mathbf{m})\) is a solution to the initial-boundary value problem (1.1)–(1.2) on [0, T].

To show the higher regularity, we notice from elliptic equation \(-\Delta \psi =\varrho \) with homogeneous Dirichlet boundary condition that \(\psi \in L^\infty ((0,T);W^{2,p}(\Omega ))\) for any \(p>1\) by using the inequality (5.11) and the \(L^p\) regularity theory [10, 23]. Next, by applying [22, Theorem 3.10.1] to the Dirichlet problem associated with the linear parabolic equation

$$\begin{aligned} \varrho _t=\partial _i(a_{ij}\varrho _{x_i}+b_i\varrho )+S\varrho , \end{aligned}$$

we get that there exists \(\theta _1\in (0,1)\) such that \(\varrho \in C^{\theta _1,\frac{\theta _1}{2}}(\Omega _T)\), due to \(\delta _{ij}+m_im_j=a_{ij}\in C^{\theta ,\frac{\theta }{2}}(\Omega _T)\) and \(a_{ij}\partial _i\psi =b_i\in L^\infty (\Omega _T)\). Hence, using the Schauder theory [10, 23] for the elliptic equation we get \(\psi \in C^{2+\theta _1,\frac{\theta _1}{2}}(\Omega _T)\), which also implies from Lemma 2.4 that \(\mathbf{m}\in C^{2+\theta _2,1+\frac{\theta _2}{2}}(\Omega _T)\) for some \(\theta _2\in (0,1)\) by utilizing the standard Schauder theory for the parabolic equation. In the sequel, by employing the standard Schauder theory for parabolic equation again, we infer that \(\varrho \in C^{2+\theta _3,1+\frac{\theta _3}{2}}(\Omega _T)\) for some \(\theta _3\in (0,1)\), due to \(\mathbf{m}\in C^{2+\theta _2,1+\frac{\theta _2}{2}}(\Omega _T)\) and \(\psi \in C^{2+\theta _1,\frac{\theta _1}{2}}(\Omega _T)\), provided that \(S\in C^\alpha (\Omega )\) with \(\alpha \in (0,1)\), which leads to \(\psi \in C^{2+\theta _3,1+\frac{\theta _3}{2}}(\Omega _T)\). This implies the assertions in Lemma 3.1. \(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Li, B. Global Existence and Decay Estimates of Solutions of a Parabolic–Elliptic–Parabolic System for Ion Transport Networks. Results Math 75, 45 (2020). https://doi.org/10.1007/s00025-020-1172-y

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s00025-020-1172-y

Mathematics Subject Classification

Keywords

Navigation