Nothing Special   »   [go: up one dir, main page]

Skip to main content
Log in

Well-posedness and qualitative properties of a dynamical model for the ideal free distribution

  • Published:
Journal of Mathematical Biology Aims and scope Submit manuscript

Abstract

Understanding the spatial distribution of populations in heterogeneous environments is an important problem in ecology. In the case of a population of organisms that can sense the quality of their environment and move to increase their fitness, one theoretical description of the expected distribution of the population is the ideal free distribution, where individuals locate themselves to optimize fitness. A model for a dynamical process that allows a population to achieve an ideal free distribution was proposed by the Cosner (Theor Popul Biol 67:101–108, 2005). The model is based on a reaction–diffusion–advection equation with nonlinear diffusion which is similar to a porous medium equation with additional advection and population growth terms. We establish that the model is well-posed, show that solutions stabilize, determine the stationary states, discuss their stability, and describe the biological interpretation of the results.

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

Access this article

Subscribe and save

Springer+ Basic
$34.99 /Month
  • Get 10 units per month
  • Download Article/Chapter or eBook
  • 1 Unit = 1 Article or 1 Chapter
  • Cancel anytime
Subscribe now

Buy Now

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

  • Aronson DG (1986) The porous medium equation. Nonlinear diffusion problems, Lect. 2nd 1985 Sess. C.I.M.E., Montecatini Terme/Italy 1985. Lect Notes Math 1224:1–46

  • Aronson D, Crandall MG, Peletier LA (1982) Stabilization of solutions of a degenerate nonlinear diffusion problem. Nonlinear Anal TMA 6(10):1001–1022

    Article  MATH  MathSciNet  Google Scholar 

  • Averill I, Lou Y, Munther D (2012) On several conjectures from evolution of dispersal. J Biol Dyn 6:117–130

    Article  Google Scholar 

  • Battin J (2004) When good animals love bad habitats: ecological traps and the conservation of animal populations. Conserv Biol 18:1482–1491

    Article  Google Scholar 

  • Belgacem F, Cosner C (1995) The effects of dispersal along environmental gradients on the dynamics of populations in heterogeneous environments. Can Appl Math Q 3:379–397

    MATH  MathSciNet  Google Scholar 

  • Cantrell RS, Cosner C, DeAngelis D, Padron V (2007) The ideal free distribution as an evolutionarily stable strategy. J Biol Dyn 1:249–271

    Article  MATH  MathSciNet  Google Scholar 

  • Cantrell RS, Cosner C, Lou Y (2008) Approximating the ideal free distribution via reaction–diffusion–advection equations. J Differ Equ 245:3687–3703

    Article  MATH  MathSciNet  Google Scholar 

  • Cantrell RS, Cosner C, Lou Y (2010) Evolution of dispersal in heterogeneous landscapes. In: Cantrell RS, Cosner C, Ruan S (eds) Spatial ecology. Chapman Hall/CRC, Boca Raton, pp 213–227

  • Cantrell RS, Cosner C, Lou Y (2010) Evolution of dispersal and ideal free distribution. Math. Biosci. Eng. 7:17–36

    Article  MATH  MathSciNet  Google Scholar 

  • Cantrell RS, Cosner C, Lou Y (2012) Evolutionary stability of ideal free dispersal strategies in patchy environments. J Math Biol 65:943–965

    Article  MATH  MathSciNet  Google Scholar 

  • Cantrell RS, Cosner C, Lou Y, Xie C (2013) Random dispersal versus fitness dependent dispersal. J Differ Equ 254:2905–2941

    Google Scholar 

  • Cosner C (2005) A dynamic model for the ideal-free distribution as a partial differential equation. Theor Popul Biol 67:101–108

    Article  MATH  Google Scholar 

  • Delgado M, Suárez A (2000) On the existence of dead cores for degenerate Lotka–Volterra models. Proc R Soc Edinburgh A 130:743–766

    Article  MATH  Google Scholar 

  • Fretwell SD (1972) Populations in a seasonal environment. Princeton University Press, Princeton

    Google Scholar 

  • Fretwell SD, Lucas HL Jr (1970) On territorial behavior and other factors influencing habitat selection in birds. Theoretical development. Acta Biotheretica 19:16–36

    Article  Google Scholar 

  • Haugen TO, Winfield IJ, Vøllestad LA, Fletcher JM, James JB, Stenseth NC (2006) The ideal free pike: 50 years of fitness-maximizing dispersal in Windermere. Proc R Soc B 273:2917–2924

    Article  Google Scholar 

  • Krivan V, Cressman R, Schneider C (2008) The ideal free distribution: a review and synthesis of the game theoretic perspective. Theor Popul Biol 73:403–425

    Article  MATH  Google Scholar 

  • Ladyzenskaja OA, Solonnikov VA, Uralceva NN (1968) Linear and quasi-linear equations of parabolic type. AMS, Providence

    Google Scholar 

  • Lions PL (1984) Structure of the set of steady-state solutions and asymptotic behaviour of semilinear heat equations. J Differ Equ 53:362–386

    Article  MATH  Google Scholar 

  • McCall A (1990) Dynamic geography of marine fish populations Washington sea grant. University of Washington Press, Seattle

  • Milinski M (1979) An evolutionarily stable feeding strategy in sticklebacks. Zeitschrift für Tierpsychologie 51:36–40

    Article  Google Scholar 

  • Oksanen T, Power M, Oksanen L (1995) Ideal habitat selection and consumer resource dynamics. Am Nat 146:565–585

    Article  Google Scholar 

  • Porzio MM, Vespri V (2003) Holder estimates for solutions of some doubly nonlinear degenerate parabolic equations. J Differ Equ 103:146–178

    Article  MathSciNet  Google Scholar 

  • Rowell JT (2009) The limitation of species range: a consequence of searching along resource gradients. Theor Popul Biol 75:216–227

    Article  MATH  Google Scholar 

  • Rowell JT (2010) Tactical population movements and distributions for ideally motivated competitors. Am Nat 176:638–650

    Article  Google Scholar 

  • Sebastián-González E, Botella F, Sempere RA, Sánchez-Zapata JA (2011) An empirical demonstration of the ideal free distribution: little grebes Tachybaptus rucollis breeding in intensive agricultural landscapes. Ibis 152:643–650

    Article  Google Scholar 

  • Simader CG (1972) On Dirichlet’s boundary value problem. An \(L^p\)-theory based on a generalization of Garding’s inequality. Lect Notes Math 268:1–246

    Article  MathSciNet  Google Scholar 

  • Tomi F (1969) Über semilineare elliptische Differentialgleichungen zweiter Ordnung. Math Z 111:350–366

    Article  MATH  MathSciNet  Google Scholar 

Download references

Acknowledgments

The research of CC was partially supported by NSF Grant 1118623.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Chris Cosner.

Appendices

Appendix A: Well-posedness. Proof of Theorem 2.1

In view of well-known results on the lack of regularity in degenerate diffusion problems of type (1.3) (cf. e.g. Aronson 1986), it seems adequate to study (1.3) in the framework of weak solutions. For definiteness we explicitly formulate the solution concept pursued below as follows. Here and in the following section, there is actually no need to require \(u_0\) to be continuous, whence we shall suppose here that \(u_0\) belongs to \(L^\infty (\Omega )\) only. Further relaxations are of course possible, including solutions possibly unbounded near \(t=0\); but since our focus below will rather be on the large time behavior of solutions, we shall not pursue the question of optimal initial regularity any further.

Definition 4.1

Let \(T>0\), and suppose that \(u_0 \in L^\infty (\Omega )\) is nonnegative. Then by a weak solution of (1.3) in \(\Omega \times (0,T)\) we mean a nonnegative function

$$\begin{aligned} u \in C^0([0,T];L^1(\Omega )) \cap L^2((0,T);W^{1,2}(\Omega )) \cap L^\infty (\Omega \times (0,T)) \end{aligned}$$
(4.1)

with the property that for each \(t\in (0,T]\), the identity

$$\begin{aligned} -\int _0^t \int _\Omega u \varphi _t + \int _\Omega u(\cdot ,t)\varphi (\cdot ,t)&= \int _\Omega u_0 \varphi (\cdot ,0) - \int _0^t \int _\Omega u\nabla (u-m) \cdot \nabla \varphi \nonumber \\&+ \int _0^t \int _\Omega u(m-u)\varphi \end{aligned}$$
(4.2)

holds for any nonnegative \(\varphi \in C^1(\bar{\Omega }\times [0,t])\).

A weak subsolution (resp. supersolution) of (1.3) in \(\Omega \times (0,T)\) is a function from the above class which satisfies (4.2) with ‘\(=\)’ replaced by ‘\(\le \)’ (resp. ‘\(\ge \)’). Finally, a measurable function \(u:\Omega \times (0,\infty ) \rightarrow \mathbb {R}\) is said to be a global weak (sub-/super-) solution of (1.3) if \(u\) is a weak (sub-/super-)solution of (1.3) in \(\Omega \times (0,T)\) for any \(T>0\).

1.1 A comparison principle. Uniqueness

We first make sure that in the framework of weak (sub- and super-)solutions, (1.3) allows for a comparison principle. In the verification thereof we follow the lines of a celebrated approach introduced in Aronson et al. (1982) for equations of the form \(u_t=\Delta \phi (u) + f(u)\), which involves solutions of a problem dual to (1.3) in an appropriate sense. Though the basic idea in Lemma 4.2 will thus be straightforward, it turns out that the presence of the convective term in (1.3) requires some additional efforts; to mention one particular technical detail in this respect, we note that unlike in the original argument, it seems that the auxiliary function \(\psi _\delta \) introduced below (cf. 4.5) fails to possess an appropriate second-order regularity property, e.g. boundedness of some weighted space–time \(L^2\) norm of \(\Delta \psi _\delta \), guaranteeing suitable decay of error terms stemming from regularization of the degeneracy (cf. the last term in 4.10).

The comparison principle provided by the following lemma will be essential to most of our results below. Since we could not find a version thereof in the literature which is sufficient for our purpose, we therefore include a full proof here.

Lemma 4.2

Let \(T>0\), and suppose that \(u\) resp. \(v\) are a weak sub- resp. supersolution of (1.3) in \(\Omega \times (0,T)\), corresponding to the nonnegative initial data \(u_0 \in L^\infty (\Omega )\) resp. \(v_0 \in L^\infty (\Omega )\). Furthermore, assume that both \(u\) and \(v\) belong to \(C^{\theta ,\frac{\theta }{2}}(\bar{\Omega }\times (0,T])\) for some \(\theta \in (0,1)\), and that

$$\begin{aligned} \int _0^T \int _\Omega \chi _{\{u>0\}} u^{p-1} |\nabla u|^2 + \int _0^T \int _\Omega \chi _{\{v>0\}} v^{p-1} |\nabla v|^2 < \infty \end{aligned}$$
(4.3)

for some \(p\in (0,1)\). Then

$$\begin{aligned} \int _\Omega (u(\cdot ,t)-v(\cdot ,t))_+ \le e^{M_0 t} \cdot \int _\Omega (u_0-v_0)_+ \quad \text{ for } \text{ all } t\in (0,T) \end{aligned}$$
(4.4)

with \(M_0=\Vert m_+\Vert _{L^\infty (\Omega )}\) as in Theorem 2.1. In particular, if \(u_0 \le v_0\) in \(\Omega \) then \(u \le v\) in \(\Omega \times (0,T)\).

Proof

Since (4.1) requires \(u\) and \(v\) to be continuous in \([0,T]\) as \(L^1(\Omega )\)-valued functions, replacing \(t\) by \(t+\tau \) and taking \(\tau \searrow 0\) if necessary we may restrict ourselves to proving (4.4) under the stronger assumption that \(u\) and \(v\) belong to \(C^{\theta ,\frac{\theta }{2}}(\bar{\Omega }\times [0,T])\).

Then performing a variant of a well-established procedure presented in Aronson et al. (1982), given \(t_0 \in (0,T), \delta \in (0,1)\) and \(\chi \in S:=\{\tilde{\chi }\in C_0^\infty (\Omega ) \ | \ 0 \le \tilde{\chi }\le 1 \}\), we let \(\psi _\delta \) denote the solution of the problem

$$\begin{aligned} \!\!\left\{ \begin{array}{l} \psi _{\delta t} = \frac{1}{2}(u\!+\!v\!+\!\delta )\Delta \psi _\delta \!+\! \nabla m \cdot \nabla \psi _\delta + m\psi _\delta - (u+v)\psi _\delta , \quad x\in \Omega , \ t\in (0,t_0), \\ \frac{\partial \psi _\delta }{\partial \nu }=0, \quad x\in \partial \Omega , \ t\in (0,t_0), \\ \psi _\delta (x,0)=\chi (x), \quad x\in \Omega . \end{array} \right. \nonumber \\ \end{aligned}$$
(4.5)

Since \(u+v\) is nonnegative and Hölder continuous in \(\bar{\Omega }\times [0,T]\), standard theory for linear non-degenerate parabolic equations (Ladyzenskaja et al. 1968) asserts that (4.5) indeed possesses a classical solution \(\psi _\delta \in C^{2+\theta ,1+\frac{\theta }{2}}(\bar{\Omega }\times [0,t_0])\) which, according to the classical comparison principle, satisfies

$$\begin{aligned} 0 \le \psi _\delta (x,t) \le e^{M_0 t} \quad \text{ for } \text{ all } (x,t) \in \Omega \times (0,t_0). \end{aligned}$$
(4.6)

Let us first make sure that there exists \(c_1>0\) such that with \(p\in (0,1)\) as in (4.3) we have

$$\begin{aligned} \int _0^{t_0} \int _\Omega |\nabla \psi _\delta |^2 \le c_1 \delta ^{-p-1} \quad \text{ for } \text{ all } \delta \in (0,1). \end{aligned}$$
(4.7)

To see this, we multiply (4.5) by \(\psi _\delta \) and integrate by parts over \(\Omega \) to obtain

$$\begin{aligned} \frac{1}{2} \int _\Omega \psi _\delta ^2(\cdot ,t_0) \!&= \! \frac{1}{2} \int _\Omega \chi ^2 \!-\! \frac{1}{2} \int _0^{t_0} \int _\Omega (u+v+\delta ) |\nabla \psi _\delta |^2 - \frac{1}{2} \int _0^{t_0} \int _\Omega \psi _\delta \nabla (u+v) \cdot \nabla \psi _\delta \nonumber \\&+ \int _0^{t_0} \int _\Omega \psi _\delta \nabla m \cdot \nabla \psi _\delta + \int _0^{t_0} \int _\Omega m \psi _\delta ^2 - \int _0^{t_0} \int _\Omega (u+v) \psi _\delta ^2 \nonumber \\&=: I_1+\cdots +I_6 \quad \text{ for } \text{ all } t\in (0,t_0). \end{aligned}$$
(4.8)

Clearly, \(I_1 \le \frac{|\Omega |}{2}\), \(I_6 \le 0\) and

$$\begin{aligned} I_5 \le c_2:=M_0 e^{2M_0 t} \cdot |\Omega | T, \end{aligned}$$

and another integration by parts combined with (4.6) shows that

$$\begin{aligned} I_4&= - \frac{1}{2} \int _0^{t_0} \int _\Omega \psi _\delta ^2 \Delta m + \frac{1}{2} \int _0^{t_0} \int _{\partial \Omega } \psi _\delta ^2 \cdot \frac{\partial m}{\partial \nu } \\&\le c_3:=\frac{M_2}{2} e^{2M_0 T} \cdot |\Omega | T + \frac{M_1}{2} e^{2M_0T} \cdot |\partial \Omega |_{n-1} T, \end{aligned}$$

where \(M_2:=\Vert (-\Delta m)_+\Vert _{L^\infty (\Omega )}, M_1:=\Vert \nabla m\Vert _{L^\infty (\Omega )}\) and \(|\partial \Omega |_{n-1}\) denotes the \((n-1)\)-dimensional Lebesgue measure of \(\partial \Omega \).

As for \(I_3\), we invoke Young’s inequality to estimate

$$\begin{aligned} |I_3| \le \frac{1}{4} \int _0^{t_0} \int _\Omega (u+v+\delta ) |\nabla \psi _\delta |^2 + \frac{1}{4} \int _0^{t_0} \int _\Omega \psi _\delta ^2 \cdot \frac{|\nabla (u+v)|^2}{u+v+\delta }, \end{aligned}$$

where according to (4.6),

$$\begin{aligned} \frac{1}{4} \int _0^{t_0} \int _\Omega \psi _\delta ^2 \cdot \frac{|\nabla (u+v)|^2}{u+v+\delta }&\le \frac{e^{2M_0 T}}{4} \int _0^{t_0} \int _\Omega \frac{|\nabla (u+v)|^2}{u+v+\delta } \\&\le \frac{e^{2M_0 T}}{2} \cdot \left\{ \int _0^{t_0} \int _\Omega \frac{|\nabla u|^2}{u+\delta } + \int _0^{t_0} \int _\Omega \frac{|\nabla v|^2}{v+\delta } \right\} . \end{aligned}$$

Here we apply the elementary inequality

$$\begin{aligned} \xi ^{1-p} \le \eta \xi + c(p) \eta ^{-\frac{1-p}{p}}, \end{aligned}$$

valid for all \(\xi \ge 0\) and \(\eta >0\) with \(c(p):=(1-p)^\frac{1-p}{p}-(1-p)^\frac{1}{p}>0\), to \(\xi :=u\) and \(\eta :=(\frac{c(p)}{\delta })^p\) to see that

$$\begin{aligned} u+\delta&\ge \frac{1}{\eta } \cdot \left( u^{1-p}-c(p) \cdot \eta ^{-\frac{1-p}{p}} \right) + \delta \\&= \left( \frac{\delta }{c(p)}\right) ^p \cdot u^{1-p} \quad \text{ in } \Omega \times (0,T). \end{aligned}$$

Therefore, since \(\nabla u=0\) a.e. in \(\{u=0\}\) (Simader 1972),

$$\begin{aligned} \int _0^{t_0} \int _\Omega \frac{|\nabla u|^2}{u+\delta } \le \left( \frac{c(p)}{\delta }\right) ^p \cdot \int _0^{t_0} \int _\Omega \chi _{\{u>0\}} u^{p-1} |\nabla u|^2, \end{aligned}$$

and proceeding similarly for \(v\) we conclude using (4.3) that (4.8) implies

$$\begin{aligned} \frac{1}{4} \int _0^{t_0} \int _\Omega (u+v+\delta ) |\nabla \psi _\delta |^2 \le c_4 \cdot (1+\delta ^{-p}) \end{aligned}$$

with some \(c_4\,{>}\,0\) independent of \(\delta \,{\in }\, (0,1)\), \(t_0\,{\in }\, (0,T)\) and \(\chi \,{\in }\, S\). Since \(u+v+\delta \ge \delta \), this yields (4.7).

We now choose \(\varphi _\delta (x,t):=\psi _\delta (x,t_0-t)\), \((x,t)\in \bar{\Omega }\times [0,t_0]\), as a test function in (4.2).

This first gives

$$\begin{aligned}&-\int _0^{t_0} \int _\Omega (u-v) \cdot \varphi _{\delta t} + \int _\Omega (u(\cdot ,t_0)-v(\cdot ,t_0)) \cdot \varphi _\delta (\cdot ,t_0)\nonumber \\&\quad \le \int _\Omega (u_0-v_0) \cdot \varphi _\delta (\cdot ,0) \nonumber \\&\qquad - \int _0^{t_0} \int _\Omega (u\nabla u - v\nabla v) \cdot \nabla \varphi _\delta \nonumber \\&\qquad + \int _0^{t_0} \int _\Omega (u-v) \nabla m \cdot \nabla \varphi _\delta \nonumber \\&\qquad + \int _0^{t_0} \int _\Omega (u-v)m\varphi _\delta \nonumber \\&\qquad - \int _0^{t_0} \int _\Omega (u^2-v^2) \varphi _\delta , \end{aligned}$$
(4.9)

where (4.6) implies that

$$\begin{aligned} \int _\Omega (u_0-v_0) \cdot \varphi _\delta (\cdot ,0)= \int _\Omega (u_0-v_0)\psi _\delta (\cdot ,t_0) \le e^{M_0 t_0} \int _\Omega (u_0-v_0)_+ . \end{aligned}$$

Since \(\frac{\partial \varphi _\delta }{\partial \nu }=0\) on \(\partial \Omega \), an integration by parts shows that

$$\begin{aligned} - \int _0^{t_0} \int _\Omega (u\nabla u-v\nabla v) \cdot \nabla \varphi _\delta =\frac{1}{2} \int _0^{t_0} \int _\Omega (u^2-v^2) \Delta \varphi _\delta , \end{aligned}$$

whence from (4.9) we obtain

$$\begin{aligned}&\int _\Omega (u(\cdot ,t_0)-v(\cdot ,t_0)) \cdot \varphi _\delta (\cdot ,t_0) \le e^{M_0 t_0} \int _\Omega (u_0-v_0)_++ \int _0^{t_0} \int _\Omega (u-v) \\&\quad \times \cdot \left\{ \varphi _{\delta t} + \frac{1}{2}(u+v)\Delta \varphi _\delta + \nabla m \cdot \nabla \varphi _\delta + m\varphi _\delta - (u+v)\varphi _\delta \right\} . \end{aligned}$$

According to (4.5), we thus infer that

$$\begin{aligned} \int _\Omega (u(\cdot ,t_0)-v(\cdot ,t_0)) \cdot \chi \le e^{M_0 t_0} \int _\Omega (u_0-v_0)_+ -\frac{\delta }{2} \int _0^{t_0} \int _\Omega (u-v)\Delta \varphi _\delta ,\qquad \qquad \end{aligned}$$
(4.10)

where we once again integrate by parts and estimate using the Cauchy-Schwarz inequality to see that

$$\begin{aligned} -\frac{\delta }{2} \int _0^{t_0} \int _\Omega (u-v)\Delta \varphi _\delta \!&= \! \frac{\delta }{2} \int _0^{t_0} \int _\Omega \nabla (u-v) \cdot \nabla \varphi _\delta \\ \!&\le \! \frac{\delta }{2} ( \Vert \nabla u\Vert _{L^2(\Omega \times (0,T))} {+}\Vert \nabla v\Vert _{L^2(\Omega \times (0,T))}) \cdot \left( \int _0^{t_0} \int _\Omega |\nabla \varphi _\delta |^2 \right) ^\frac{1}{2}\!. \end{aligned}$$

Thanks to (4.7), however, we know that

$$\begin{aligned} \delta \cdot \left( \int _0^{t_0} \int _\Omega |\nabla \psi _\delta |^2 \right) ^\frac{1}{2} \le c_1^\frac{1}{2} \cdot \delta ^\frac{1-p}{2} \rightarrow 0 \quad \text{ as } \delta \searrow 0, \end{aligned}$$

whence from (4.10) we conclude that

$$\begin{aligned} \int _\Omega (u(\cdot ,t_0)-v(\cdot ,t_0)) \cdot \chi \le e^{M_0 t_0} \int _\Omega (u_0-v_0)_+. \end{aligned}$$

Since \(\chi \in S\) and \(t_0 \in (0,T)\) were arbitrary, this yields (4.4). \(\square \)

Fortunately, the integrability condition (4.3) is automatically fulfilled by any weak solution of (1.3).

Lemma 4.3

Suppose that \(u_0 \in L^\infty (\Omega )\) is nonnegative, and that \(u\) is a weak solution of (1.3) in \(\Omega \times (0,T)\) for some \(T>0\). Then for all \(p\in (0,1)\) there exists \(C(p)>0\) such that

$$\begin{aligned} \int _0^T \int _\Omega \chi _{\{u>0\}} u^{p-1} |\nabla u|^2 \le C(p). \end{aligned}$$
(4.11)

Proof

We first note that \(u\) is continuous in \(\bar{\Omega }\times (0,T]\) according to (Porzio and Vespri (2003), Theorem 1.3), and hence the corresponding positivity set \(\mathcal{P} := \{(x,t) \in \bar{\Omega }\times (0,T] \ | \ u(x,t)>0\}\) is relatively open in \(\bar{\Omega }\times (0,T]\). Therefore, standard parabolic regularity results (Ladyzenskaja et al. 1968) assert that \(u\) even belongs to \(C^{2,1}(\mathcal{P})\) and satisfies the PDE and the boundary conditions in (1.3) classically in \(\mathcal{P}\), because \(m \in C^{2+\kappa }(\bar{\Omega })\).

Based on this information, we plan to study the evolution of a regularized version of the functional \(\int _\Omega u^p(\cdot ,t)\) by performing a suitable truncation near \(u=0\). Due to the lack of both convexity and differentiability near zero of \(0 < \xi \mapsto \xi ^p\), however, appropriate modification of straightforward cut-off procedures appears to be in order.

To achieve this, for \(\delta \in (0,1)\) we let \(\chi _\delta \in C^\infty ([0,\infty )\) be nondecreasing and such that \(\chi _\delta \equiv 0\) on \([0,\frac{\delta }{2}]\) and \(\chi _\delta \equiv 1\) on \([\delta ,\infty )\), and such that \(\chi _\delta \nearrow 1\) on \((0,\infty )\) as \(\delta \searrow 0\). We next let

$$\begin{aligned} \psi _\delta (s):=\chi _\delta (s) \cdot s^{p-1} + \left( \frac{\delta }{2}\right) ^{p-1} (1-\chi _\delta (s)), \quad s\ge 0, \end{aligned}$$

for \(\delta \in (0,1)\), and observe that \(\psi _\delta \) is nonnegative and smooth on \([0,\infty )\) with

$$\begin{aligned} \psi _\delta '(s)&= (p-1) \chi _\delta (s) \cdot s^{p-2} + \chi _\delta '(s) \cdot \left\{ s^{p-1}-\left( \frac{\delta }{2}\right) ^{p-1} \right\} \nonumber \\&\le (p-1) \chi _\delta (s) \cdot s^{p-2} \le 0 \quad \text{ for } \text{ all } s\ge 0 \end{aligned}$$
(4.12)

thanks to the above properties of \(\chi _\delta \) and the fact that \(p<1\). The primitive

$$\begin{aligned} \Psi _\delta (s):=\int _0^\delta \psi _\delta (\sigma )d\sigma , \quad s\ge 0, \end{aligned}$$

clearly inherits nonnegativity from \(\psi _\delta \), and moreover we have the upper estimate

$$\begin{aligned} \Psi _\delta (s)&= \int _0^s \chi _\delta (\sigma ) \sigma ^{p-1} d\sigma + \left( \frac{\delta }{2}\right) ^{p-1} \cdot \int _0^s (1-\chi _\delta (\sigma ))d\sigma \nonumber \\&\le \int _0^s \sigma ^{p-1}d\sigma + \left( \frac{\delta }{2}\right) ^{p-1} \cdot \delta = \frac{1}{p} s^p + \frac{\delta ^p}{2^{p-1}} \quad \text{ for } \text{ all } s\ge 0.\qquad \qquad \quad \end{aligned}$$
(4.13)

Now since \(u\in C^{2,1}(\mathcal{P})\), the smoothness of \(\Psi _\delta \) implies that \(\Psi _\delta (u)\) belongs to \(C^{2,1}(\bar{\Omega }\times (0,T])\). We thus may differentiate and integrate by parts to obtain for fixed \(\tau \in (0,T)\) the identity

$$\begin{aligned} \int _\Omega \Psi _\delta (u(\cdot ,T)) - \int _\Omega \Psi _\delta (u(\cdot ,\tau ))&= \int _\tau ^T \int _\Omega \psi _\delta (u)u_t \\&= - \int _\tau ^T \int _\Omega u\psi _\delta '(u) \nabla u \cdot \nabla (u-m)\\&+ \int _\tau ^T \int _\Omega u\psi _\delta (u) (m-u), \end{aligned}$$

from which we infer that

$$\begin{aligned} - \int _\tau ^T \int _\Omega u\psi _\delta '(u) |\nabla u|^2&= \int _\Omega \Psi _\delta (u(\cdot ,T)) - \int _\Omega \Psi _\delta (u(\cdot ,\tau )) - \int _\tau ^T \int _\Omega u\psi _\delta '(u) \nabla u \cdot \nabla m\nonumber \\&- \int _\tau ^T \int _\Omega mu\psi _\delta (u) + \int _\tau ^T \int _\Omega u^2 \psi _\delta (u) \nonumber \\&=: I_1+\cdots +I_5 \end{aligned}$$
(4.14)

for any \(\delta \in (0,1)\) and \(\tau \in (0,T)\). Here, by nonnegativity of \(\Psi _\delta \) we have

$$\begin{aligned} I_2 \le 0, \end{aligned}$$
(4.15)

whereas (4.13) entails that

$$\begin{aligned} I_1 \le c_1(\delta ):=\frac{1}{p} \int _\Omega u^p(\cdot ,T) + \frac{\delta ^p|\Omega |}{2^{p-1}}. \end{aligned}$$
(4.16)

Moreover, using that \(s(1-\chi _\delta (s)) \le \delta \) and hence

$$\begin{aligned} s\psi _\delta (s) \le \chi _\delta (s) \cdot s^p + \left( \frac{\delta }{2}\right) ^{p-1} \cdot s(1-\chi _\delta (s)) \le s^p + \frac{\delta ^p}{2^{p-1}} \quad \text{ for } \text{ all } s\ge 0,\nonumber \\ \end{aligned}$$
(4.17)

we can estimate

$$\begin{aligned} I_4 \le c_4(\delta ):=|\Omega | T \cdot \Vert m\Vert _{L^\infty (\Omega )} \cdot \left\{ \Vert u\Vert ^p_{L^\infty (\Omega \times (0,T))} + \frac{\delta ^p}{2^{p-1}} \right\} . \end{aligned}$$
(4.18)

Likewise, we obtain

$$\begin{aligned} I_5 \le c_5(\delta ) := |\Omega | T \cdot \left\{ \Vert u\Vert _{L^\infty (\Omega \times (0,T))}^{p+1} +\frac{\delta ^p}{2^{p-1}} \Vert u\Vert _{L^\infty (\Omega \times (0,T))} \right\} . \end{aligned}$$
(4.19)

As for \(I_3\), we once more integrate by parts to find that

$$\begin{aligned} I_3 = - \int _\tau ^T \int _\Omega \nabla \rho _\delta (u) \cdot \nabla m = \int _\tau ^T \int _\Omega \rho _\delta (u) \Delta m - \int _\tau ^T \int _{\partial \Omega } \rho _\delta (u) \cdot \frac{\partial m}{\partial \nu } \end{aligned}$$

holds with

$$\begin{aligned} \rho _\delta (s):=\int _0^s \sigma \psi _\delta '(\sigma )d\sigma \equiv s\psi _\delta (s) - \Psi _\delta (s), \quad s\ge 0. \end{aligned}$$

Since by (4.13) and (4.17) we have

$$\begin{aligned} |\rho _\delta (s)| \le s\psi _\delta (s)+\Psi _\delta (s) \le \left( 1+\frac{1}{p}\right) s^p + \left( \frac{\delta }{2}\right) ^p \quad \text{ for } \text{ all } s\ge 0, \end{aligned}$$

again using that \(m\in C^2(\bar{\Omega })\) we conclude that

$$\begin{aligned} |I_3| \le c_3(\delta )&:= T \cdot \left\{ \left( 1+\frac{1}{p}\right) \cdot \Vert u\Vert _{L^\infty (\Omega \times (0,T))}\right. \nonumber \\&\left. + \left( \frac{\delta }{2}\right) ^p \right\} \cdot \{ |\Omega | \cdot \Vert \Delta m\Vert _{L^\infty (\Omega )} + |\partial \Omega |_{n-1} \cdot \Vert \nabla m\Vert _{L^\infty (\Omega )}\}.\nonumber \\ \end{aligned}$$
(4.20)

Collecting (4.15), (4.16), (4.18), (4.19) and (4.20), from (4.14) we conclude that

$$\begin{aligned} -\int _\tau ^T \int _\Omega u\psi _\delta '(u)|\nabla u|^2 \le c_6 := \sup _{\delta \in (0,1)} \left\{ c_1(\delta )+c_3(\delta )+c_4(\delta ) +c_5(\delta )\right\} \end{aligned}$$

for all \(\delta \in (0,1)\) and \(\tau \in (0,T)\), where we observe that \(c_6\) is finite due to the evident fact that \(c_1,c_3,c_4\) and \(c_5\) are all increasing in \(\delta \). In light of (4.12), this implies that

$$\begin{aligned} (1-p) \int _\tau ^T \int _\Omega \chi _\delta (u) u^{p-1} |\nabla u|^2 \le c_6 \end{aligned}$$

for all \(\delta \in (0,1)\) and \(\tau \in (0,T)\), so that twice applying the monotone convergence theorem in taking \(\delta \searrow 0\) and then \(\eta \searrow 0\) yields the claim. \(\square \)

As a consequence, we may apply Lemma 4.2 in a standard way so as to obtain first that the nonlinear semigroup generated by (1.3) is quasi-contractive in \(L^1(\Omega )\).

Corollary 4.4

Suppose that \(u_0\in L^\infty (\Omega )\) and \(v_0\in L^\infty (\Omega )\) are nonnegative, and that \(u\) and \(v\) are weak solutions of (1.3) in \(\Omega \times (0,T)\) for some \(T>0\) with initial data \(u_0\) and \(v_0\), respectively. Then with \(M_0\) as in Theorem 2.1, we have

$$\begin{aligned} \Vert u(\cdot ,t)-v(\cdot ,t)\Vert _{L^1(\Omega )} \le e^{M_0 t} \Vert u_0-v_0\Vert _{L^1(\Omega )} \quad \text{ for } \text{ all } t\in (0,T). \end{aligned}$$
(4.21)

Proof

If \(u\) and \(v\) are two weak solutions, both \(u\) and \(v\) are Hölder continuous in \(\bar{\Omega }\times (0,T]\) by Porzio and Vespri (2003). In view of Lemma 4.3, \(u\) and \(v\) satisfy the regularity requirements from Lemma 4.2, whence twice applying the latter we easily obtain (4.21). \(\square \)

In particular, this immediately implies that weak solutions are unique.

Corollary 4.5

Let \(u_0\in L^\infty (\Omega )\) be nonnegative and \(T>0\). Then the problem (1.3) possesses at most one weak solution in \(\Omega \times (0,T)\).

1.2 Global existence and approximation of solutions

Our next goal is to construct a solution of (1.3) by performing a suitable approximation procedure. To this end, we consider the regularized versions of (1.3) as given by

$$\begin{aligned} \left\{ \begin{array}{l} u_{\varepsilon t} = \nabla \cdot ((u_\varepsilon +\varepsilon ) \nabla u_\varepsilon - u_\varepsilon \nabla m) + u_\varepsilon ( m-u_\varepsilon ), \quad x\in \Omega , \ t>0, \\ (u_\varepsilon +\varepsilon )\frac{\partial u_\varepsilon }{\partial \nu }=u_\varepsilon \frac{\partial m}{\partial \nu }, \quad x\in \partial \Omega , \ t>0, \\ u_\varepsilon (x,0)=u_0(x), \quad x\in \Omega , \end{array} \right. \qquad \qquad \qquad \end{aligned}$$
(4.22)

for \(\varepsilon \in (0,1)\). According to Cantrell et al. (2008), for each \(\varepsilon \in (0,1)\) this problem possesses a global classical solution \(u_\varepsilon \in C^{2,1}(\bar{\Omega }\times (0,\infty )) \cap C^0(\bar{\Omega }\times [0,\infty ))\) which is nonnegative in \(\bar{\Omega }\times [0,\infty )\) and thus, as a consequence of the strong maximum principle, even strictly positive in \(\bar{\Omega }\times (0,\infty )\).

We shall next derive \(\varepsilon \)-independent estimates for the solutions of (4.22). We first invoke an argument based on the classical maximum principle to establish a pointwise upper bound for \(u_\varepsilon \).

Lemma 4.6

There exists \(C>0\) such that for all \(\varepsilon \in (0,1)\) we have

$$\begin{aligned} u_\varepsilon (x,t) \le C \quad \text{ for } \text{ all } x\in \Omega \text{ and } t>0. \end{aligned}$$
(4.23)

Proof

We let

$$\begin{aligned} \rho _\varepsilon (s):=s+\varepsilon \ln s \quad \text{ for } s>0 \text{ and } \varepsilon \in (0,1). \end{aligned}$$

Then \(\rho _\varepsilon '>0\) on \((0,\infty )\) and \(\rho _\varepsilon ((0,\infty ))=\mathbb {R}\). The inverse function \(\rho _\varepsilon ^{-1}:\mathbb {R}\rightarrow (0,\infty )\) is thus increasing on \(\mathbb {R}\) and satisfies \(\rho _\varepsilon ^{-1}(1)=1\) and

$$\begin{aligned} \rho _\varepsilon ^{-1}(\sigma ) \ge \frac{\sigma }{2} \quad \text{ for } \text{ all } \sigma >0, \end{aligned}$$

because since \(\ln s \le s\) for \(s>0\) we have

$$\begin{aligned} \rho _\varepsilon (s) \le s+\varepsilon s \le 2s \quad \text{ for } \text{ all } s>0. \end{aligned}$$
(4.24)

Likewise, the inequality \(\rho _\varepsilon (s) \ge s\) for \(s>1\) entails that

$$\begin{aligned} \rho _\varepsilon ^{-1}(\sigma ) \le \sigma \quad \text{ for } \text{ all } \sigma >1. \end{aligned}$$
(4.25)

In view of (4.24), it is possible to fix \(A>1\) large enough such that

$$\begin{aligned} A \ge 2\Vert u_0\Vert _{L^\infty (\Omega )} \quad \text{ and } \quad \rho _\varepsilon ^{-1}(A) \ge \Vert m\Vert _{L^\infty (\Omega )} \quad \text{ for } \text{ all } \varepsilon \in (0,1).\qquad \quad \end{aligned}$$
(4.26)

With this value of \(A\) being fixed henceforth, given \(\varepsilon \in (0,1)\) we now define

$$\begin{aligned} v(x,t):=\rho _\varepsilon (u_\varepsilon (x,t)) - m(x) \quad \text{ and } \quad \overline{v}(x,t):=A \end{aligned}$$

for \(x\in \bar{\Omega }\) and \(t\ge 0\). Then by (4.22) and the fact that \(u_\varepsilon \) is positive,

$$\begin{aligned} v_t&= \rho _\varepsilon '(u_\varepsilon ) u_{\varepsilon t} \\&= \left( 1+\frac{\varepsilon }{u_\varepsilon } \right) \left( \nabla \cdot \left\{ (u_\varepsilon +\varepsilon )\nabla u_\varepsilon - u_\varepsilon \nabla m \right\} +u_\varepsilon (m-u_\varepsilon ) \right) \\&= \left( 1+\frac{\varepsilon }{u_\varepsilon }\right) \left( \nabla \cdot \left\{ u_\varepsilon \nabla \cdot \left( \rho _\varepsilon (u_\varepsilon )-m\right) \right\} +u_\varepsilon (m-u_\varepsilon ) \right) \\&= \left( 1+\frac{\varepsilon }{u_\varepsilon }\right) ( \nabla \cdot (u_\varepsilon \nabla v) + \rho _\varepsilon ^{-1}(v+m) \cdot \{ m-\rho _\varepsilon ^{-1}(v+m)\}) \quad \text{ in } \Omega \times (0,\infty ). \end{aligned}$$

Moreover,

$$\begin{aligned} \frac{\partial v}{\partial \nu }&= \rho _\varepsilon '(u_\varepsilon ) \frac{\partial u_\varepsilon }{\partial \nu } - \frac{\partial m}{\partial \nu } \\&= \frac{1}{u_\varepsilon } \cdot \left\{ (u_\varepsilon +\varepsilon )\frac{\partial u_\varepsilon }{\partial \nu } - u_\varepsilon \frac{\partial m}{\partial \nu }\right\} \\&= 0 \quad \text{ on } \partial \Omega \end{aligned}$$

and

$$\begin{aligned} v(x,0) \le \rho _\varepsilon (u_0(x)) \le 2\Vert u_0\Vert _{L^\infty (\Omega )} \quad \text{ for } \text{ all } x\in \Omega \end{aligned}$$

according to (4.24).

On the other hand, \(\overline{v}\) satisfies \(\frac{\partial \overline{v}}{\partial \nu }=0\) on \(\partial \Omega \) and

$$\begin{aligned} \overline{v}(x,0)=A \ge 2\Vert u_0\Vert _{L^\infty (\Omega )} \quad \text{ for } \text{ all } x\in \Omega \end{aligned}$$

thanks to (4.26), and furthermore we have

$$\begin{aligned} I&:= \overline{v}_t - \left( 1+\frac{\varepsilon }{u_\varepsilon }\right) \, \nabla \cdot (u_\varepsilon \nabla \overline{v}) - \left( 1+\frac{\varepsilon }{u_\varepsilon }\right) \cdot \rho _\varepsilon ^{-1}(\overline{v}+m) \cdot \{m-\rho _\varepsilon ^ {-1} (\overline{v}+m)\} \\&= - \left( 1+\frac{\varepsilon }{u_\varepsilon }\right) \cdot \rho _\varepsilon ^{-1}(A+m) \cdot \{ m- \rho _\varepsilon ^{-1} (A+m)\} \quad \text{ in } \Omega \times (0,\infty ). \end{aligned}$$

Here we use (4.26) along with the monotonicity of \(\rho _\varepsilon ^{-1}\) and the nonnegativity of \(m\) to estimate

$$\begin{aligned} \rho _\varepsilon ^{-1} (A+m(x)) \ge \rho _\varepsilon ^{-1}(A) \ge m(x) \ge 0 \quad \text{ for } \text{ all } x\in \Omega \end{aligned}$$

to obtain that

$$\begin{aligned} I \ge 0 \quad \text{ in } \Omega \times (0,\infty ). \end{aligned}$$

Therefore, the comparison principle ensures that \(v\le \overline{v}\) in \(\Omega \times (0,\infty )\) and thus, again since \(\rho _\varepsilon ^{-1}\) increases, that

$$\begin{aligned} u_\varepsilon =\rho _\varepsilon ^{-1}(v+m) \le \rho _\varepsilon ^{-1}(\overline{v}+m) \le \rho _\varepsilon ^{-1}(A+\Vert m\Vert _{L^\infty (\Omega )}) \quad \text{ in } \Omega \times (0,\infty ). \end{aligned}$$

In light of (4.25) and the fact that \(A>1\), this proves that (4.23) is valid with \(C:=A+\Vert m\Vert _{L^\infty (\Omega )}\). \(\square \)

We next use the uniformity of the upper bound in Lemma 4.6 along with well-established regularity theory for degenerate parabolic equations to conclude the following.

Lemma 4.7

There exists \(\theta >0\) with the property that for all \(\tau \in (0,1)\) one can find \(C(\tau )>0\) such that

$$\begin{aligned} \Vert u_\varepsilon \Vert _{C^{\theta ,\frac{\theta }{2}}(\bar{\Omega }\times [t,t+1])} \le C(\tau ) \quad \text{ for } \text{ all }\quad t\ge \tau \text{ and } \varepsilon \in (0,1). \end{aligned}$$
(4.27)

In particular,

$$\begin{aligned} \Vert u_\varepsilon (\cdot ,t)\Vert _{C^{\theta }(\bar{\Omega })} \le C(\tau ) \quad \text{ for } \text{ all }\quad t\ge \tau \text{ and } \varepsilon \in (0,1) \end{aligned}$$
(4.28)

and

$$\begin{aligned} (u_\varepsilon )_{\varepsilon }\in (0,1)\, \hbox {is relatively compact in}\, C^0_{loc}(\bar{\Omega }\times (0,\infty )). \end{aligned}$$
(4.29)

If moreover \(u_0 \in C^1(\bar{\Omega })\) then \(C(\tau )\) can be chosen to be independent of \(\tau \in (0,1)\), and then \((u_\varepsilon )_{\varepsilon \in (0,1)}\) is relatively compact even in \(C^0_{loc}(\bar{\Omega }\times [0,\infty ))\).

Proof

For \(\varepsilon \in (0,1)\), the function \(v_\varepsilon :=u_\varepsilon -\varepsilon \) is a classical solution of the problem

$$\begin{aligned} \left\{ \begin{array}{l} v_{\varepsilon t} = \nabla \cdot \left( v_\varepsilon \nabla v_\varepsilon - (v_\varepsilon +\varepsilon ) \nabla m \right) + (v_\varepsilon +\varepsilon ) ( m-v_\varepsilon -\varepsilon ), \quad x\in \Omega , \ t>0, \\ v_\varepsilon \frac{\partial v_\varepsilon }{\partial \nu }=(v_\varepsilon +\varepsilon ) \frac{\partial m}{\partial \nu }, \quad x\in \partial \Omega , \ t>0, \\ v_\varepsilon (x,0)=v_{0\varepsilon }(x):=u_0(x)+\varepsilon , \quad x\in \Omega , \end{array} \right. \end{aligned}$$

with porous medium-type diffusive term. Since \((v_\varepsilon )_{\varepsilon \in (0,1)}\) is bounded in \(L^\infty (\Omega \times (0,\infty ))\) by Lemma 4.6, the estimate (4.27) therefore directly follows from the interior Hölder regularity result in (Porzio and Vespri (2003), Theorem 1.3). Whereas (4.28) and (4.29) are immediate consequences of (4.27) and the Arzelà–Ascoli theorem, the claimed independence of \(C(\tau )\) of \(\tau \) is asserted by the corresponding boundary regularity statement ((Porzio and Vespri (2003), Remark 1.4)). \(\square \)

In the limit procedure \(\varepsilon \searrow 0\) to be carried out in Lemma 4.9 below, we shall moreover need a convenient compactness property of \((\nabla u_\varepsilon )_{\varepsilon \in (0,1)}\).

Lemma 4.8

There exists \(C>0\) such that for all \(\varepsilon \in (0,1)\) we have

$$\begin{aligned} \int _t^{t+1} \int _\Omega |\nabla u_\varepsilon |^2 \le C \quad \text{ for } \text{ all } t\ge 0. \end{aligned}$$
(4.30)

Proof

Recalling that \(u_\varepsilon \) is positive, we may multiply (4.22) by \(\ln u_\varepsilon \) and integrate by parts over \(\Omega \) to obtain

$$\begin{aligned} \frac{d}{dt} \int _\Omega (u_\varepsilon \ln u_\varepsilon - u_\varepsilon ) \!=\! - \int _\Omega \frac{u_\varepsilon +\varepsilon }{u_\varepsilon } |\nabla u_\varepsilon |^2 \!+\! \int _\Omega \nabla u_\varepsilon \cdot \nabla m + \int _\Omega u_\varepsilon \ln u_\varepsilon \cdot (m-u_\varepsilon )\nonumber \\ \end{aligned}$$
(4.31)

for all \(t>0\), where clearly

$$\begin{aligned} \int _\Omega \frac{u_\varepsilon +\varepsilon }{u_\varepsilon } |\nabla u_\varepsilon |^2 \ge \int _\Omega |\nabla u_\varepsilon |^2 \end{aligned}$$

and, by Young’s inequality,

$$\begin{aligned} \left| \int _\Omega \nabla u_\varepsilon \cdot \nabla m \right| \le \frac{1}{2} \int _\Omega |\nabla u_\varepsilon |^2 + \frac{1}{2} \int _\Omega |\nabla m|^2. \end{aligned}$$

Since moreover Lemma 4.6 ensures that for some \(c_1>0\) and \(c_2>0\) we have

$$\begin{aligned} \left| \int _\Omega u_\varepsilon \ln u_\varepsilon \cdot (m-u_\varepsilon ) \right|&\le c_2 \quad \text{ and } \quad \left| \int _\Omega (u_\varepsilon \ln u_\varepsilon - u_\varepsilon ) \right| \le c_2 \quad \text{ for } \text{ all } t>0 \text{ and } \\&\varepsilon \in (0,1), \end{aligned}$$

(4.31) entails that

$$\begin{aligned} \frac{1}{2} \int _t^{t+1} \int _\Omega |\nabla u_\varepsilon |^2 \le 2c_2 + \frac{1}{2} \int _\Omega |\nabla m|^2 + c_1 \end{aligned}$$

for any \(t\ge 0\) and \(\varepsilon \in (0,1)\). \(\square \)

We can now use the boundedness and compactness properties collected above in a standard way to make sure that indeed (1.3) is globally weakly solvable.

Lemma 4.9

There exists a nonnegative function

$$\begin{aligned} u\in L^\infty (\Omega \times (0,\infty )) \, \cap \, C^0(\bar{\Omega }\times [0,\infty )) \, \cap \, L^2_{loc}([0,\infty );W^{1,2}(\Omega ))\qquad \qquad \end{aligned}$$
(4.32)

with the property that for each \(T>0\),

$$\begin{aligned} \begin{array}{lll} u_\varepsilon &{}\rightarrow &{} u \quad \text{ in } C^0(\bar{\Omega }\times [0,T]) \quad \text{ and } \\ \nabla u_\varepsilon &{}\rightharpoonup &{} \nabla u \quad \text{ in } L^2(\Omega \times (0,T)) \end{array} \end{aligned}$$
(4.33)

as \(\varepsilon \searrow 0\). Moreover, \(u\) is a global weak solution of (1.3) in the sense of Definition 4.1.

Proof

In view of Lemmas 4.7 and 4.8 it is evident upon standard extraction arguments that by passing to an appropriate sequence \((\varepsilon _k)_{k\in \mathbb {N}}\) such that \(\varepsilon _k\searrow 0\) as \(k\rightarrow \infty \), we can achieve that for each \(T>0\) and \(\tau \in (0,T)\) we have

$$\begin{aligned} \begin{array}{lll} u_{\varepsilon _k} &{}\rightarrow &{} u \quad \text{ in } C^0(\bar{\Omega }\times [\tau ,T]) \quad \text{ and } \\ \nabla u_{\varepsilon _k} &{}\rightharpoonup &{} \nabla u \quad \text{ in } L^2(\Omega \times (0,T)) \end{array} \end{aligned}$$
(4.34)

with some \(u \in C^0(\bar{\Omega }\times (0,\infty )) \cap L^2_{loc}([0,\infty );W^{1,2}(\Omega ))\). Clearly, \(u\) is nonnegative, and moreover Lemma 4.6 warrants that \(u\) is bounded. It remains to verify that along some subsequence of \((\varepsilon _k)_{k\in \mathbb {N}}\) this convergence in fact is uniform in \(\Omega \times (0,1)\). Indeed, once this has been shown, (4.32) and hence the regularity requirements in (4.1) are satisfied, and the validity of (4.2) can readily be checked on testing (4.22) by \(\varphi \in C^1(\bar{\Omega }\times [0,t])\) and using (4.33) in a straightforward manner; finally, it will then follow from the uniqueness of weak solutions and a standard argument that the convergence statements in (4.33) actually hold along the entire net \(\varepsilon \searrow 0\).

To assert the claimed convergence property, given \(\eta >0\) we fix \(\overline{u}_0 \in C^1(\bar{\Omega })\) such that \(u_0 \le \overline{u}_0 \le u_0+\frac{\eta }{4}\). From the classical comparison principle we then obtain that the corresponding solutions of (4.22) satisfy \(\overline{u}_\varepsilon \ge u_\varepsilon \) in \(\Omega \times (0,\infty )\). Moreover, another application of Lemma 4.7 shows that since \(\overline{u}_0 \in C^1(\bar{\Omega })\), along a subsequence – still denoted by \((\varepsilon _k)_{k\in \mathbb {N}}\) – we have \(\overline{u}_\varepsilon \rightarrow \overline{u}\) in \(C^0_{loc}(\bar{\Omega }\times [0,\infty ))\) with a certain limit function \(\overline{u}\). Now since \(\overline{u}\) is continuous at \(t=0\), there exists \(\tau _\eta \in (0,1)\) such that \(\overline{u}(\cdot ,t) \le \overline{u}_0+\frac{\eta }{8}\) for all \(t\in (0,\tau _\eta )\), and since \(\overline{u}_\varepsilon \rightarrow \overline{u}\) uniformly in \(\Omega \times (0,\tau _\eta )\) as \(\varepsilon =\varepsilon _k\searrow 0\), this implies that \(\overline{u}_\varepsilon (\cdot ,t) \le \overline{u}_0+\frac{\eta }{4}\) for all \(t\in (0,\tau _\eta )\), provided that \(\varepsilon \in (\varepsilon _k)_{k\in \mathbb {N}}\) is suitably small. Thus, for any such \(\varepsilon \) we find that

$$\begin{aligned} u_\varepsilon (\cdot ,t) \le \overline{u}_\varepsilon (\cdot ,t) \le u_0+\frac{\eta }{2} \quad \text{ for } \text{ all } t\in (0,\tau _\eta ), \end{aligned}$$
(4.35)

and by a similar approximation from below we obtain that on passing to a further subsequence and diminishing \(\tau _\eta \) if necessary, also

$$\begin{aligned} u_\varepsilon (\cdot ,t) \ge u_0-\frac{\eta }{2} \quad \text{ for } \text{ all } t\in (0,\tau _\eta ) \end{aligned}$$
(4.36)

holds for all sufficiently small \(\varepsilon \in (\varepsilon _k)_{k\in \mathbb {N}}\). Combining (4.35) and (4.36) shows that

$$\begin{aligned} |u_\varepsilon -u_{\varepsilon '}| \le \eta \quad \text{ in } \Omega \times (0,\tau _\eta ) \end{aligned}$$

for all small \(\varepsilon ,\varepsilon ' \in (\varepsilon _k)_{k\in \mathbb {N}}\). Together with (4.34), this proves that \((u_{\varepsilon _k})_{k\in \mathbb {N}}\) forms a Cauchy sequence in \(C^0(\bar{\Omega }\times [0,1])\), as desired. \(\square \)

Proof of Theorem 2.1

The existence statement is part of the result provided by Lemma 4.9. Uniqueness and the estimate (2.1) are precisely asserted by Corollaries 4.5 and 4.4. \(\square \)

Appendix B: Weak steady states. Proof of Proposition 2.2

The natural counterpart of Definition 4.1 is the following.

Definition 5.1

A stationary weak (sub-/super-)solution of (1.3) is a nonnegative function \(w \in L^\infty (\Omega ) \cap W^{1,2}(\Omega )\) satisfying

$$\begin{aligned} \int _\Omega w \nabla (w-m) \cdot \nabla \psi = (\le , \ge ) \, \int _\Omega w(m-w) \psi \end{aligned}$$
(5.1)

for all nonnegative \(\psi \in C^1(\bar{\Omega })\).

We proceed to identify a family of stationary weak sub- and supersolutions of (1.3).

Lemma 5.2

Suppose that \(\Omega _0 \subset \Omega \) is open, and that

$$\begin{aligned} m(x)+a \ge 0 \quad \text{ for } \text{ all } x\in \Omega _0 \quad \text{ and } \quad m(x)+a=0 \quad \text{ for } \text{ all } x\in \partial \Omega _0 \cap \Omega \qquad \end{aligned}$$
(5.2)

with some \(a\in \mathbb {R}\). Then

$$\begin{aligned} w(x):= \left\{ \begin{array}{l@{\quad }l} m(x)+a \quad &{} \text{ if } x\in \Omega _0, \\ 0 &{} \text{ if } x\in \Omega \!\setminus \! \Omega _0, \end{array} \right. \end{aligned}$$
(5.3)

defines a stationary weak subsolution of (1.3) if \(a \le 0\), and a stationary weak supersolution of (1.3) if \(a\ge 0\).

Proof

We first note that (5.2) asserts that \(w\) is nonnegative and continuous in \(\bar{\Omega }\) and hence also belongs to \(W^{1,2}(\Omega )\), because \(m\in C^1(\bar{\Omega })\). We thus only need to check the respective integral inequalities in (5.1).

In fact, if \(a\ge 0\) then for any nonnegative \(\psi \in C^1(\bar{\Omega })\) we have

$$\begin{aligned} \int _\Omega w\nabla (w-m) \cdot \nabla \psi = \int _{\Omega _0} (m+a) \nabla a \cdot \nabla \psi =0, \end{aligned}$$

because \(a=const.\), and

$$\begin{aligned} \int _\Omega w(m-w) \psi = - a \int _{\Omega _0} (m+a)\psi \le 0, \end{aligned}$$

for \(m+a \ge 0\) in \(\Omega _0\) by (5.2). This shows that \(w\) is a stationary weak supersolution in this case, whereas the subsolution property in the case \(a\le 0\) can be seen similarly. \(\square \)

This already proves one part of Proposition 2.2:

Corollary 5.3

Let \(\sigma =(\sigma _i)_{i\in \mathbb {N}} \subset \{0,1\}\). Then the function \(m_\sigma \) defined in (1) is a stationary weak solution of (1.3).

Proof

We only need to apply Lemma 5.2 to the set

$$\begin{aligned} \Omega _0:=\bigcup \limits _{i\in \mathbb {N}, \ \sigma _i=1} G_i, \end{aligned}$$

which clearly satisfies (5.2) with \(a=0\). \(\square \)

In fact, the above corollary already reveals all conceivable steady states. In order to verify this, and thus to complete the proof of Proposition 2.2, we provide an elementary preparation.

Lemma 5.4

Suppose that \(w\in C^0(\bar{\Omega })\) is nonnegative and such that

$$\begin{aligned} w \cdot (w-m)^2 \equiv 0 \quad \text{ in } \Omega , \end{aligned}$$
(5.4)

and let \(G\subset \{m>0\}\) be open and connected. Then either \(w\equiv 0\) in \(G\) or \(w\equiv m\) in \(G\).

Proof

By continuity of \(w\) and \(m\), the sets

$$\begin{aligned} G_1:= \{ x\in G |\ w(x)>0\} \quad \text{ and } \quad G_2:=\{x\in G \Vert \ w(x)\ne m(x)\} \end{aligned}$$

are both open, and we first claim that \(G_1 \cap G_2 = \emptyset \). Indeed, this results from the fact that if \(x\in G_1\) then \(w(x)>0\), and hence (5.4) enforces that \(w(x)=m(x)\), that is, \(x\not \in G_2\).

Moreover, (5.4) along with the positivity of \(m\) inside \(G\) implies that \(G=G_1 \cup G_2\): To see this, we pick any \(x\in G\) and then obtain from (5.4) that either \(w(x)=0\) or \(w(x)=m(x)\). In the former case, clearly \(w(x)\ne m(x)\) and thus \(x\in G_2\), whereas in the latter we have \(w(x)>0\) and therefore \(x\in G_1\).

Now since \(G\) is connected, the above properties imply that either \(G_1=G\) or \(G_2=G\), again in view of (5.4) meaning that either \(w\equiv m\) in \(G\) or \(w\equiv 0\) in \(G\). \(\square \)

We are now in the position to fully characterize the set of all weak equilibria of (1.3).

Proof

of Proposition 2.2.    In view of Corollary 5.3, we only need to make sure that any stationary weak solution \(w\) of (1.3) is of the form in (2.2) with some \((\sigma _i)_{i\in \mathbb {N}} \subset \{0,1\}\). To see this, we first observe that since \(\Omega \times (0,\infty ) \ni (x,t) \mapsto w(x)\) is a global weak solution of (1.3), it follows from (Porzio and Vespri (2003), Theorem 1.3) that \(w\) is continuous in \(\bar{\Omega }\). Since (5.1) clearly extends so as to remain valid for each \(\psi \in W^{1,2}(\Omega )\), we may choose \(\psi :=w-m\) as a test function there to obtain

$$\begin{aligned} \int _\Omega w|\nabla (w-m)|^2 + \int _\Omega w(w-m)^2 =0. \end{aligned}$$

Since \(w\) is nonnegative, the claim therefore easily results from Lemma 5.4. \(\square \)

Appendix C: Large time behavior. Proof of Theorem 2.3

We next examine the large time behavior of the solution \(u\) gained above. Our analysis in this direction will be based on the integral inequality (6.2) satisfied by the approximate solutions \(u_\varepsilon \). This inequality can be viewed as an approximate version of the energy identity

$$\begin{aligned} \frac{1}{2} \frac{d}{dt} \int _\Omega (u-m)^2 = - \int _\Omega u | \nabla (u-m)|^2 - \int _\Omega u(u-m)^2, \quad t>0, \end{aligned}$$
(6.1)

formally associated with (1.3).

Lemma 6.1

Suppose that \(u_0\in C^0(\bar{\Omega }\) be nonnegative, and let \(\varepsilon \in (0,1)\). Then given any \(\eta \in (0,1)\), the solution \(u_\varepsilon \) of (4.22) satisfies the inequality

$$\begin{aligned}&\frac{d}{dt} \int _\Omega \left\{ \frac{1}{2} (u_\varepsilon -m)^2 + \varepsilon (u_\varepsilon \ln u_\varepsilon - u_\varepsilon ) \right\} + \int _\Omega u_\varepsilon \left| \nabla (u_\varepsilon +\varepsilon \ln u_\varepsilon - m ) \right| ^2\nonumber \\&\quad + (1-\eta ) \int _\Omega u_\varepsilon (u_\varepsilon -m)^2 \le \frac{\varepsilon ^2}{4\eta } \int _\Omega u_\varepsilon \ln ^2 u_\varepsilon \end{aligned}$$
(6.2)

for all \(t>0\).

Proof

We test (4.22) against \(u_\varepsilon +\varepsilon \ln u_\varepsilon - m\) to obtain

$$\begin{aligned}&\frac{d}{dt} \int _\Omega \left\{ \frac{1}{2} (u_\varepsilon -m)^2 + \varepsilon (u_\varepsilon \ln u_\varepsilon - u_\varepsilon )\right\} \nonumber \\&\qquad = \int _\Omega (u_\varepsilon +\varepsilon \ln u_\varepsilon -m) \cdot u_{\varepsilon t} \nonumber \\&\qquad = - \int _\Omega \nabla (u_\varepsilon +\varepsilon \ln u_\varepsilon -m) \cdot \left( (u_\varepsilon +\varepsilon )\nabla u_\varepsilon - u_\varepsilon \nabla m\right) \nonumber \\&\quad \qquad + \int _\Omega (u_\varepsilon +\varepsilon \ln u_\varepsilon -m) \cdot u_\varepsilon (m-u_\varepsilon ) \nonumber \\&\qquad =: I_1+I_2 \end{aligned}$$
(6.3)

for \(t>0\). On simple rearrangements, we see that

$$\begin{aligned} I_1=-\int _\Omega u_\varepsilon \left| \nabla (u_\varepsilon +\varepsilon \ln u_\varepsilon -m) \right| ^2 \end{aligned}$$

and

$$\begin{aligned} I_2= - \int _\Omega u_\varepsilon (u_\varepsilon -m)^2 + \varepsilon \int _\Omega u_\varepsilon \ln u_\varepsilon \cdot (m-u_\varepsilon ), \end{aligned}$$

where by Young’s inequality

$$\begin{aligned} \varepsilon \int _\Omega u_\varepsilon \ln u_\varepsilon \cdot (m-u_\varepsilon ) \le \eta \int _\Omega u_\varepsilon (u_\varepsilon -m)^2 + \frac{\varepsilon ^2}{4\eta } \int _\Omega u_\varepsilon \ln ^2 u_\varepsilon , \end{aligned}$$

so that (6.2) is a consequence of (6.3). \(\square \)

We now take \(\varepsilon \searrow 0\) here to conclude that \(u\) at least satisfies a weakened version of (6.1).

Lemma 6.2

Assume that \(u_0 \in C^0(\bar{\Omega })\) is nonnegative, and let \(t_0\ge 0\). Then for the solution \(u\) of (1.3) we have

$$\begin{aligned}&\frac{1}{2} \int _\Omega (u(x,t)-m(x))^2 dx + \int _{t_0}^t \int _\Omega u(u-m)^2\nonumber \\&\qquad \le \frac{1}{2} \int _\Omega (u(x,t_0)-m(x))^2 dx \quad \text{ for } \text{ all } t>t_0. \end{aligned}$$
(6.4)

In particular,

$$\begin{aligned} t \mapsto \int _\Omega (u(x,t)-m(x))^2 dx \quad \text{ is } \text{ nonincreasing } \text{ on } [0,\infty ) \end{aligned}$$
(6.5)

and

$$\begin{aligned} \int _0^\infty \int _\Omega u(u-m)^2 < \infty . \end{aligned}$$
(6.6)

Proof

For fixed \(t_0>0\) and \(t>t_0\) and arbitrary \(\eta \in (0,1)\) we may integrate (6.2) over the time interval \((t_0,t)\) to see upon dropping a nonnegative term on the left that

$$\begin{aligned}&\frac{1}{2} \int _\Omega (u_\varepsilon (x,t)-m(x))^2 dx +(1-\eta ) \int _{t_0}^t \int _\Omega u_\varepsilon (u_\varepsilon -m)^2 \nonumber \\&\quad \le \frac{1}{2} \int _\Omega ( u_\varepsilon (x,t_0)-m(x))^2 dx \nonumber \\&\qquad -\,\varepsilon \int _\Omega ( u_\varepsilon (x,t)\ln u_\varepsilon (x,t)-u_\varepsilon (x,t)) dx\nonumber \\&\qquad +\, \varepsilon \int _\Omega ( u_\varepsilon (x,t_0)\ln u_\varepsilon (x,t_0)-u_\varepsilon (x,t_0)) dx \nonumber \\&\qquad +\, \frac{\varepsilon ^2}{4\eta } \int _{t_0}^t \int _\Omega u_\varepsilon \ln ^2 u_\varepsilon \end{aligned}$$
(6.7)

holds for all \(\varepsilon \in (0,1)\). Now according to Lemma 4.6 and the nonnegativity of \(u_\varepsilon \) we know that there exist \(c_1>0\) and \(c_2>0\) such that

$$\begin{aligned} |u_\varepsilon \ln u_\varepsilon - u_\varepsilon | \le c_1 \quad \text{ and } \quad |u_\varepsilon \ln ^2 u_\varepsilon | \le c_2 \quad \text{ in } \Omega \times (0,\infty ) \end{aligned}$$

for all \(\varepsilon \in (0,1)\), whence it follows that

$$\begin{aligned}&-\, \varepsilon \int _\Omega ( u_\varepsilon (x,t)\ln u_\varepsilon (x,t)\!-\!u_\varepsilon (x,t)) dx \!+\! \varepsilon \int _\Omega ( u_\varepsilon (x,t_0)\ln u_\varepsilon (x,t_0)-u_\varepsilon (x,t_0)) dx\nonumber \\&\quad +\, \frac{\varepsilon ^2}{4\eta } \int _{t_0}^t \int _\Omega u_\varepsilon \ln ^2 u_\varepsilon \le 2c_1 |\Omega | \varepsilon + \frac{(t-t_0) |\Omega |}{4\eta } \varepsilon ^2 \nonumber \\&\quad \rightarrow 0 \quad \text{ as } \varepsilon \searrow 0. \end{aligned}$$
(6.8)

Moreover, using the uniform convergence \(u_{\varepsilon } \rightarrow u\) in \(\bar{\Omega }\times [t_0,t]\) asserted by Lemma 4.9 we find that

$$\begin{aligned} \int _\Omega (u_\varepsilon (x,s)-m(x))^2 dx \rightarrow \int _\Omega (u(x,s)-m(x))^2 dx \quad \text{ for } \text{ all } s\in [t_0,t] \end{aligned}$$

and

$$\begin{aligned} \int _{t_0}^t \int _\Omega u_\varepsilon (u_\varepsilon -m)^2 \rightarrow \int _{t_0}^t \int _\Omega u(u-m)^2 \end{aligned}$$

as \(\varepsilon \searrow 0\). In view of (6.8), we thus conclude from (6.7) in the limit \(\varepsilon =\searrow 0\) that

$$\begin{aligned} \frac{1}{2} \int _\Omega (u(x,t)-m(x))^2 dx + (1-\eta ) \int _{t_0}^t \int _\Omega u(u-m)^2 \le \frac{1}{2} \int _\Omega (u(x,t_0)-m(x))^2 dx \end{aligned}$$

for each \(\eta \in (0,1)\), which clearly implies (6.4). \(\square \)

Remark

In fact, in accordance with (6.2) it is possible to include a nonnegative space–time integral involving \(\nabla u\) on the left-hand side of (6.4), and thus obtain an energy inequality more closely related to (6.1). The additional information thereby obtained with regard to the total energy dissipation is not needed in the sequel, however.

As a first – though yet rather weak – information on the asymptotics of \(u\), the inequality (6.6) says that there exists some \((t_k)_{k\in \mathbb {N}} \subset (0,\infty )\) such that \(t_k \rightarrow \infty \) and

$$\begin{aligned} \int _\Omega u(x,t_k) \cdot (u(x,t_k)-m(x))^2 dx \rightarrow 0 \end{aligned}$$

as \(k\rightarrow \infty \). This will be sharpened in the following lemma so as to cover convergence along the entire net \(t\rightarrow \infty \).

Lemma 6.3

For any choice of \(0 \le u_0 \in C^0(\bar{\Omega })\), the solution \(u\) of (1.3) satisfies

$$\begin{aligned} \int _\Omega u(x,t) \cdot (u(x,t)-m(x))^2 dx \rightarrow 0 \quad \text{ as } t\rightarrow \infty . \end{aligned}$$
(6.9)

Proof

Let us set

$$\begin{aligned} I(t):=\int _\Omega u(x,t) \cdot (u(x,t)-m(x))^2 dx \quad \text{ for } t\ge 0. \end{aligned}$$

Then assuming that the claim be false, we would obtain \((t_k)_{k\in \mathbb {N}} \subset (1,\infty )\) and \(\delta >0\) such that \(t_k\rightarrow \infty \) as \(k\rightarrow \infty \) and

$$\begin{aligned} I(t_k) \ge \delta \quad \text{ for } \text{ all } k\in \mathbb {N}, \end{aligned}$$
(6.10)

where we may assume without loss of generality that \(t_{k+1} \ge t_k+1\) for all \(k\in \mathbb {N}\). Now since \(u\) is uniformly continuous in \(\bar{\Omega }\times [1,\infty )\) according to Lemma 4.7, it follows that \(I\) is uniformly continuous in \([1,\infty )\). It is therefore possible to find \(\tau \in (0,1)\) such that

$$\begin{aligned} |I(t)-I(t_k)| \le \frac{\delta }{2} \quad \text{ for } \text{ all } t\in (t_k,t_k+\tau ) \hbox { and each } k\in \mathbb {N}, \end{aligned}$$

which implies that

$$\begin{aligned} \int _\Omega u(x,t) \cdot (u(x,t)-m(x))^2 dx \ge \frac{\delta }{2} \quad \text{ for } \text{ all } t\in (t_k,t_k+\tau ) \text{ and } \text{ each } k\in \mathbb {N}. \end{aligned}$$

Consequently, an integration in time shows that

$$\begin{aligned} \int _0^{t_k+1} \int _\Omega u(u\!-\!m)^2 \ge \sum _{j=1}^k \int _{t_j}^{t_j+\tau } \int _\Omega u(x,t) ( u(x,t)\!-\!m(x))^2 dxdt \ge k \cdot \frac{\delta }{2} \quad \text{ for } \text{ all } k\in \mathbb {N}\end{aligned}$$

and thereby contradicts Lemma 6.2. \(\square \)

Now thanks to the fact that \((u(\cdot ,t))_{t>1}\) is equicontinuous by Lemma 4.7, one particular outcome of Lemma 6.3 is that for each individual \(x\in \bar{\Omega }\) we either have \(u(x,t)\rightarrow 0\) or \(u(x,t)\rightarrow m(x)\) as \(t\rightarrow \infty \). The main concern of the following lemma is to make sure that within this alternative the selection is in fact independent of \(x \in G\) for each component \(G\) of \(\{m>0\}\).

Lemma 6.4

Assume that \(u_0 \in C^0(\bar{\Omega })\) is nonnegative, and let \(u\) denote the solution of (1.3). Moreover, suppose \((t_k)_{k\in \mathbb {N}} \subset (1,\infty )\) is such that \(t_k\rightarrow \infty \) and

$$\begin{aligned} u(\cdot ,t_k) \rightarrow u_\infty \quad \text{ in } C^0(\bar{\Omega }) \end{aligned}$$
(6.11)

as \(k\rightarrow \infty \) with some \(u_\infty \in C^0(\bar{\Omega })\). Then for each component \(G\) of \(\{m>0\}\),

$$\begin{aligned} \text{ either } \ u_\infty \equiv 0\, \hbox {in}\, G \,\quad or \quad u_\infty \equiv m\, \hbox {in}\, G. \end{aligned}$$
(6.12)

Proof

According to Lemma 6.3, \(u_\infty \) satisfies the identity \(u_\infty \cdot (u_\infty -m)^2 \equiv 0\) in \(\Omega \). Therefore, Lemma 5.4 states that either \(u_\infty \equiv 0\) in \(G\) or \(u_\infty \equiv m\) in \(G\). \(\square \)

As a crucial step towards the proof of Theorem 2.3, we next make sure that for all sufficiently small nontrivial initial data which vanish outside some component \(G\) of \(\{m>0\}\), always the latter alternative in (6.12) occurs. The proof of this will essentially rely on the monotonicity property in Lemma 6.2, which rules out the possibility that such a solution, being nontrivial and lying below \(m\) in \(G\), decays to zero in the large time limit. Solutions satisfying such special initial conditions will play an important role as comparison functions in the proof of Lemma 6.6.

Lemma 6.5

Let \(G\) be a component of \(\{m>0\}\), and let \(u_0 \in C^0(\bar{\Omega })\) be nonnegative and such that \(u_0 \not \equiv 0\) and

$$\begin{aligned} u_0 \le m \cdot \chi _G \quad \text{ in } \Omega . \end{aligned}$$
(6.13)

Then the solution \(u\) of (1.3) satisfies

$$\begin{aligned} u(\cdot ,t) \rightarrow m \cdot \chi _G \quad \text{ in } C^0(\bar{\Omega }) \quad \text{ as } t\rightarrow \infty . \end{aligned}$$
(6.14)

Proof

Since \(m \cdot \chi _G\) is a stationary weak solution of (1.3) by Proposition 2.2, (6.13) together with the comparison principle in Lemma 4.2 implies that

$$\begin{aligned} u(\cdot ,t) \le m\cdot \chi _G \quad \text{ in } \Omega \quad \text{ for } \text{ all } t\ge 0. \end{aligned}$$
(6.15)

Now suppose that (6.14) be false. Since \((u(\cdot ,t))_{t>1}\) is relatively compact in \(C^0(\bar{\Omega })\) by Lemma 4.7, we can then find a sequence \((t_k)_{k\in \mathbb {N}} \subset (1,\infty )\) fulfilling \(t_k\rightarrow \infty \) and \(u(\cdot ,t_k) \rightarrow u_\infty \) in \(C^0(\bar{\Omega })\) as \(k\rightarrow \infty \) with some nonnegative \(u_\infty \in C^0(\bar{\Omega })\) such that \(u_\infty \not \equiv m \cdot \chi _G\). But from (6.15) we know that \(u_\infty \le m\cdot \chi _G\) in \(\Omega \), so that in light of Lemma 6.4 we must have \(u_\infty \equiv 0\) in \(\Omega \). This means that \(u(\cdot ,t_k) \rightarrow 0\) in \(C^0(\bar{\Omega })\), whence in particular

$$\begin{aligned} \int _\Omega (u(\cdot ,t_k)-m)^2 \rightarrow \int _\Omega m^2 \quad \text{ as } k\rightarrow \infty . \end{aligned}$$
(6.16)

However, when combined with the downward monotonicity of \(t\mapsto \int _\Omega (u(\cdot ,t)-m)^2\) as asserted by Lemma 6.2, (6.16) says that we should have

$$\begin{aligned} \int _\Omega (u_0-m)^2 \ge \int _\Omega m^2. \end{aligned}$$

This implies that

$$\begin{aligned} 2 \int _\Omega u_0 m \le \int _\Omega u_0^2 \end{aligned}$$

and hence

$$\begin{aligned} 2 \int _\Omega u_0 m \le \int _\Omega u_0 m, \end{aligned}$$

because \(u_0 \le m \cdot \chi _G\). Since \(u_0 \cdot m\) is nontrivial and nonnegative according to our assumption, this is impossible, so that actually (6.14) must be valid. \(\square \)

Based on the above lemma, we can proceed to prove the main part of Theorem 2.3 by means of a comparison argument.

Lemma 6.6

Let \(u_0\in C^0(\bar{\Omega })\) be nonnegative. Then for each component \(G\) of \(\{m>0\}\), the solution \(u\) of (1.3) has the property that

$$\begin{aligned} \text{ either } \quad u\equiv 0\,\hbox {in}\,G \times [0,\infty ) \quad \hbox {or} \quad u(\cdot ,t) \rightarrow m\,\hbox {in}\,C^0(\bar{G}) \hbox {as} t\rightarrow \infty . \end{aligned}$$

Proof

Suppose that \(u\not \equiv 0\) in \(G \times [0,\infty )\). Then there exists \(t_0 \ge 0\) such that \(u(\cdot ,t_0) \not \equiv 0\) in \(G\). Since \(u(\cdot ,t_0)\) is continuous, we can therefore pick some nontrivial nonnegative \(\psi \in C_0^\infty (\Omega )\) such that \(\mathrm{supp} \, \psi \subset G\) and \(\psi \le u(\cdot ,t_0)\) in \(\Omega \) , where we can also achieve that \(\psi \le m\) in \(G\), because \(m\) is positive in \(G\).

We now let \(\underline{u}\) denote the weak solution of

$$\begin{aligned} \left\{ \begin{array}{l} \underline{u}_t = \nabla \cdot \left( \underline{u}\nabla (\underline{u}-m) \right) + \underline{u}(m-\underline{u}), \quad x \in \Omega , \ t>t_0, \\ \underline{u}\cdot \frac{\partial }{\partial \nu } (\underline{u}-m)=0, \quad x\in \partial \Omega , \ t>t_0, \\ \underline{u}(x,t_0) = \psi (x), \quad x\in \Omega , \end{array} \right. \end{aligned}$$

which by uniqueness can be obtained by letting \(\underline{u}(x,t):=\tilde{u}(x,t-t_0)\), \((x,t) \in \bar{\Omega }\times [t_0,\infty )\), where \(\tilde{u}\) solves (1.3) with initial data \(\psi \).

According to Lemma 6.5, since \(\psi \le m \cdot \chi _G\) in \(\Omega \) and \(\psi \not \equiv 0\), we have

$$\begin{aligned} \underline{u}(\cdot ,t)\rightarrow m \cdot \chi _G \quad \text{ in } C^0(\bar{\Omega }) \quad \text{ as } t\rightarrow \infty . \end{aligned}$$
(6.17)

On the other hand, a comparison argument involving Lemma 4.2 asserts that \(u\ge \underline{u}\) in \(\Omega \times (t_0,\infty )\). In conjunction with (6.17) this shows that

$$\begin{aligned} \liminf _{t\rightarrow \infty } u(x,t) \ge m(x) \quad \text{ for } \text{ all } x\in G, \end{aligned}$$

so that an application of Lemma 6.4 warrants that indeed \(u(x,t) \rightarrow m(x)\) as \(t\rightarrow \infty \), uniformly with respect to \(x\in G\). \(\square \)

Now our main result on the asymptotics in (1.3) actually reduces to a corollary:

Proof of Theorem 2.3

The claimed solution behavior in \(\{m>0\}\) is precisely described by Lemma 6.6. Outside \(\{m>0\}\), however, Lemma 6.3 and Lemma 4.7 easily imply that \(u(\cdot ,t) \rightarrow 0\) in \(C^0(\{m \le 0\})\) as \(t\rightarrow \infty \). \(\square \)

Appendix D: Decay to zero. Proof of Proposition 2.4

Clearly, Theorem 2.3 implies that if \(m\le 0\) throughout \(\Omega \) then any solution \(u\) of (1.3) will satisfy \(u(\cdot ,t) \rightarrow 0\) in \(C^0(\bar{\Omega })\) as \(t\rightarrow \infty \). In the case when \(\{m>0\}\) is not empty, however, the question whether such nontrivial but decaying solutions exist appears to be nontrivial. The following lemma provides our main step in identifying a class of situation where such phenomena occur.

Lemma 7.1

Suppose that \(\Omega _1 \subset \Omega \) is open and such that

$$\begin{aligned} \sup _{x\in \partial \Omega _1 \cap \Omega } m(x) \le 0 \quad \text{ and } \quad \sup _{x\in \partial \Omega _1 \cap \Omega } m(x) < \sup _{x\in \Omega _1} m(x). \end{aligned}$$
(7.1)

Then there exists a nontrivial nonnegative \(w \in C^0(\bar{\Omega })\) with \(w \equiv 0\) in \(\Omega \setminus \Omega _1\) such that whenever \(u_0 \in C^0(\bar{\Omega })\) is nonnegative fulfilling \(u_0 \le w\) in \(\Omega \), the solution \(u\) of (1.3) satisfies

$$\begin{aligned} u \equiv 0 \quad \text{ in } (\Omega {\setminus } \Omega _1) \times [0,\infty ). \end{aligned}$$
(7.2)

Proof

Fixing \(x_1 \in \bar{\Omega }_1\) such that \(m(x_1)=M:=\sup _{x\in \Omega _1} m(x)\), we know from (7.1) that \(x_1\) is an interior point of \(\Omega _1\), and that moreover for some nonnegative \(a\) with \(-a<M\) we have

$$\begin{aligned} m(x) \le -a \quad \text{ for } \text{ all } x\in \partial \Omega _1 \cap \Omega . \end{aligned}$$

We now let \(\Omega _0\) denote the component of \(\{m>-a\}\) which contains \(x_1\), and then evidently have \(m+a=0\) throughout \(\partial \Omega _0 \cap \Omega \). Since \(a\ge 0\), Lemma 5.2 says that the nonnegative function \(w\) defined in (5.3) is a stationary weak supersolution of (1.3), and the fact that \(-a<M\) asserts that \(w(x_1)\ne 0\), hence \(w\not \equiv 0\). As \(w\equiv 0\) in \(\Omega {\setminus } \Omega _0 \supset \Omega {\setminus } \Omega _1\), the claim thus results upon an application of the comparison principle in Lemma 4.2. \(\square \)

Now combined with Theorem 2.3, the above indeed yields a family of such decaying solutions under the assumption that \(m\) has the property (2.7).

Proof of Proposition 2.4

Observing that (2.7) is stronger than (7.1), we can apply Lemma 7.1 to fix \(w \in C^0(\bar{\Omega })\) with the properties listed there. we then take any nonnegative \(u_0 \in C^0(\bar{\Omega })\) such that \(u_0 \le w\) in \(\Omega \) and let \(u\) denote the solution of (1.3). Then (7.2) says that we trivially have \(u(\cdot ,t) \rightarrow 0\) in \(C^0(\bar{\Omega }{\setminus } \Omega _1)\) as \(t\rightarrow \infty \). Inside \(\Omega _1\), however, we know that \(m\le 0\) by (2.7), and hence Theorem 2.3 guarantees that \(u(\cdot ,t) \rightarrow 0\) also in \(C^0(\bar{\Omega }_1)\) as \(t\rightarrow \infty \). \(\square \)

Appendix E: Stability and attractivity properties of equilibria

1.1 Stability

The proof of Proposition 2.5 will be an immediate consequence of the following more general result.

Proposition 8.1

Let \(w\) be a stationary weak solution of (1.3). Then \(w\) is stable from below in both \(X_1\) and \(X_\infty \). Moreover, \(m_+\) is stable from above in \(X_1\), while if \(w\ne m_+\) then \(w\) is unstable from above in both \(X_1\) and \(X_\infty \).

Proof

We first note that any weak solution \(u\) of (1.3) satisfies

$$\begin{aligned} \int _\Omega u(\cdot ,t) = \int _\Omega u_0 + \int _0^t \int _\Omega u(m-u) \quad \text{ for } \text{ all } t>0, \end{aligned}$$
(8.1)

which can easily be seen on choosing \(\varphi \equiv 1\) as a test function in (4.2).

Now let \(w\) be a weak steady state of (1.3), so that \(w\equiv m\) in \(\{w>0\}\) according to Proposition 2.2. Moreover, let \(u_0 \in C^0(\bar{\Omega })\) be such that \(0\le u_0 \le w\) in \(\Omega \). Then thanks to the comparison principle in Lemma 4.2 we know that \(u(\cdot ,t)\le w\) in \(\Omega \), whence in particular \(u(\cdot ,t) \le m\) in \(\{w>0\}\) and \(u(\cdot ,t)\equiv 0\) in \(\Omega {\setminus } \{w>0\}\) for all \(t>0\). Therefore (8.1) implies

$$\begin{aligned} \int _\Omega u(\cdot ,t) - \int _\Omega u_0 = \int _0^t \int _{\{w>0\}} u(m-u) \ge 0. \end{aligned}$$

Accordingly,

$$\begin{aligned} \Vert w\!-\!u(\cdot ,t)\Vert _{L^1(\Omega )} \!=\! \int _\Omega w \!-\! \int _\Omega u(\cdot ,t) \le \int _\Omega w \!-\! \int _\Omega u_0 \!=\! \Vert w\!-\!u_0\Vert _{L^1(\Omega )} \quad \! \text{ for } \text{ all } \, t\!>\!0,\nonumber \\ \end{aligned}$$
(8.2)

which entails that \(w\) is stable from below in \(X_1\).

Next, given \(u_0\in C^0(\bar{\Omega })\) such that \(0 \le u_0 \le w\) in \(\Omega \), we let \(\delta :=\Vert w-u_0\Vert _{L^\infty (\Omega )}\), so that \(u_0 \ge (w-\delta )_+\) in \(\Omega \). Since \((w-\delta )_+\) is a stationary weak subsolution of (1.3) by Lemma 5.2, the comparison principle in Lemma 4.2 says that \(w\ge u(\cdot ,t) \ge (w-\delta )_+\) in \(\Omega \) and thus also

$$\begin{aligned} \Vert w-u(\cdot ,t)\Vert _{L^\infty (\Omega )} \le \delta =\Vert w-u_0\Vert _{L^\infty (\Omega )} \quad \text{ for } \text{ all } t>0, \end{aligned}$$
(8.3)

implying stability from below in \(X_\infty \).

The stability of \(w=m_+\) from above in \(X_1\) can be seen similarly: Namely, if \(u_0 \in C^0(\bar{\Omega })\) is such that \(u_0 \ge m_+\) in \(\Omega \) then \(u(\cdot ,t)\ge m_+ \ge m\) by comparison and hence

$$\begin{aligned} \int _\Omega u(\cdot ,t) - \int _\Omega u_0 = \int _0^t \int _\Omega u(m-u) \le 0, \end{aligned}$$

implying that

$$\begin{aligned} \Vert u(\cdot ,t)-m_+\Vert _{L^1(\Omega )} \le \Vert u_0-m_+\Vert _{L^1(\Omega )} \quad \text{ for } \text{ all } t>0. \end{aligned}$$

Finally, to see that any weak steady state \(w\ne m_+\) is unstable from above in both \(X_1\) and \(X_\infty \), it is evidently sufficient to observe that from each \(\delta >0\), the solution \(u\) of (1.3) emanating from \(u_0:=w+\delta \) satisfies \(u(\cdot ,t)\rightarrow m_+ \not \equiv w\) in \(C^0(\bar{\Omega })\) as \(t\rightarrow \infty \). \(\square \)

Our proof of the fact that \(m_+\) is stable from above also with respect to the norm in \(L^\infty (\Omega )\) is more involved, and it will be postponed to the separate Lemma 8.4 below. Before, we provide two preparatory lemmata which address approximate versions of the steady-state problem associated with (1.3).

Lemma 8.2

There exists a sequence \((w_k)_{k\in \mathbb {N}} \subset C^2(\bar{\Omega })\) of positive classical solutions to the boundary-value problems

$$\begin{aligned} \left\{ \begin{array}{l} \nabla \cdot \left( w_k \nabla (w_k-m)\right) + w_k(m-w_k) + \frac{1}{k}=0, \quad x\in \Omega , \\ w_k \frac{\partial }{\partial \nu } (w_k-m) = \frac{1}{k}, \quad x\in \partial \Omega , \end{array} \right. \end{aligned}$$
(8.4)

which satisfy

$$\begin{aligned} m_+(x) < w_k(x) \le w_{k-1}(x) \quad \text{ for } \text{ all } x\in \bar{\Omega }\end{aligned}$$
(8.5)

whenever \(k\ge 2\).

Proof

Given \(k\in \mathbb {N}\), we let

$$\begin{aligned} \eta _k:=\min \left\{ 1, \frac{1}{k\cdot (\Vert \Delta m\Vert _{L^\infty (\Omega )} + \Vert m\Vert _{L^\infty (\Omega )} +1)}, \frac{1}{k \Vert \nabla m\Vert _{L^\infty (\Omega )}} \right\} \qquad \quad \end{aligned}$$
(8.6)

and can then fix a positive function \(\phi _k\in C^\infty (\mathbb {R})\) such that \(\phi _k(s)=s\) for all \(s\ge \eta _k\) and \(\phi _k'>0\) on \(\mathbb {R}\). Therefore, the boundary-value problem

$$\begin{aligned} \left\{ \begin{array}{l} \mathcal{E}_k w \equiv - \nabla \cdot \left( \phi _k(w)\nabla (w-m)\right) - w(m-w) - \frac{1}{k}=0, \quad x\in \Omega , \\ \mathcal{B}_k w \equiv \phi _k(w) \frac{\partial }{\partial \nu } (w-m)-\frac{1}{k}=0, \quad x\in \partial \Omega , \end{array} \right. \qquad \qquad \end{aligned}$$
(8.7)

is uniformly elliptic, and we claim that for each \(k\in \mathbb {N}\) it possesses a classical solution \(w_k\in C^2(\bar{\Omega })\) such that

$$\begin{aligned} m_+(x) + \eta _k \le w_k(x) \le w_{k-1}(x) \quad \text{ for } \text{ all } x\in \Omega , \end{aligned}$$
(8.8)

with a suitably function \(w_0 \in C^2(\bar{\Omega })\). To construct the latter, we first fix any \(\psi \in C^2(\bar{\Omega })\) such that \(\frac{\partial \psi }{\partial \nu } \le -1\) on \(\partial \Omega \), and then choose a positive constant \(A\) large enough fulfilling

$$\begin{aligned} A \ge \max \{1+\psi ,1+\psi -m\} \quad \text{ in } \Omega \end{aligned}$$
(8.9)

and

$$\begin{aligned} A^2 + (\Delta \psi -2\psi +m) \cdot A + \left\{ \nabla \cdot \left( (m-\psi )\nabla \psi \right) - (m-\psi )\psi -1 \right\} \ge 0 \quad \text{ in } \Omega \nonumber \\ \end{aligned}$$
(8.10)

as well as

$$\begin{aligned} A-(m-\psi )\frac{\partial \psi }{\partial \nu } -1 \ge 0 \quad \text{ on } \partial \Omega , \end{aligned}$$
(8.11)

which is clearly possible because \(m\in C^1(\bar{\Omega })\) and \(\psi \in C^2(\bar{\Omega })\).

Upon these choices, we let \(w_0\in C^2(\bar{\Omega })\) be defined by

$$\begin{aligned} w_0(x):=A+m(x)-\psi (x), \quad x\in \Omega , \end{aligned}$$

and then obtain from (8.9) that

$$\begin{aligned} w_0 \ge \max \{m+1,1\} \equiv m_+ +1 \quad \text{ in } \Omega . \end{aligned}$$
(8.12)

In particular, \(w_0\ge 1\) and hence \(\phi _1(w_0)\equiv w_0\), so that

$$\begin{aligned} \mathcal{E}_1 w_0&= - \nabla \cdot \left( (A+m-\psi )\nabla (A-\psi )\right) + (A+m-\psi )(A-\psi ) - 1 \\&= \nabla \cdot \left( (m-\psi )\nabla \psi \right) + A\Delta \psi + A^2 -2A\psi +Am-(m-\psi )\psi -1 \\&\ge 0 \quad \text{ in } \Omega \end{aligned}$$

according to (8.10). Moreover, (8.11) implies that

$$\begin{aligned} \mathcal{B}_1 w_0&= -A\frac{\partial \psi }{\partial \nu }-(m-\psi ) \frac{\partial \psi }{\partial \nu }-1 \\&\ge A -(m-\psi )\frac{\partial \psi }{\partial \nu }-1 \\&\ge 0 \quad \text{ on } \partial \Omega , \end{aligned}$$

because \(\frac{\partial \psi }{\partial \nu }\le -1\) on \(\partial \Omega \).

We next make sure that for all \(k\in \mathbb {N}\), the functions defined by

$$\begin{aligned} \hat{w}_k(x):=\eta _k, \quad x\in \Omega , \quad \text{ and } \quad \tilde{w}_k(x):=m(x)+\eta _k, \quad x\in \Omega , \end{aligned}$$

both are subsolutions with regard to (8.7). Indeed, using (8.6) we compute

$$\begin{aligned} \mathcal{E}_k \hat{w}_k&= - \nabla \cdot \left( \eta _k \nabla (\eta _k-m)\right) - \eta _k (m-\eta _k) - \frac{1}{k} \\&= \eta _k (\Delta m - m + \eta _k) - \frac{1}{k} \\&\le \eta _k (\Vert \Delta m\Vert _{L^\infty (\Omega )} + \Vert m\Vert _{L^\infty (\Omega )}+1) - \frac{1}{k} \\&\le 0 \quad \text{ in } \Omega \end{aligned}$$

and

$$\begin{aligned} \mathcal{B}_k \hat{w}_k = - \eta _k \frac{\partial m}{\partial \nu } - \frac{1}{k} \le \eta _k \Vert \nabla m\Vert _{L^\infty (\Omega )} - \frac{1}{k} \le 0 \quad \text{ on } \partial \Omega . \end{aligned}$$

Similarly,

$$\begin{aligned} \mathcal{E}_k \tilde{w}_k = (m+\eta _k)\eta _k - \frac{1}{k} \le (\Vert m\Vert _{L^\infty (\Omega )}+1)\eta _k - \frac{1}{k} \le 0 \quad \text{ in } \Omega \end{aligned}$$

by (8.6) and, obviously,

$$\begin{aligned} \mathcal{B}_k \tilde{w}_k = -\frac{1}{k} \le 0 \quad \text{ on } \partial \Omega . \end{aligned}$$

It thus follows that \(\underline{w}_k:=\min \{\hat{w}_k,\tilde{w}_k\} \equiv m_+ +\eta _k\) is a subsolution of (8.7) in the Nagumo sense, which satisfies \(\underline{w}_k \le w_0\) in \(\Omega \) thanks to (8.12). Consequently, standard monotonicity methods for elliptic boundary-value problems (Tomi 1969) can be applied to first assert the existence of a classical solution \(w_1 \in C^2(\bar{\Omega })\) of (8.7) complying with (8.8).

Next, in order to recursively construct \(w_l\) for \(k\ge 2\), we suppose that for some \(k\ge 2\) we already have found \(w_{k-1}\in C^2(\bar{\Omega })\) fulfilling \(\mathcal{E}_{k-1} w_{k-1}=0\) in \(\Omega \), \(\mathcal{B}_{k-1} w_{k-1}=0\) on \(\partial \Omega \) and \(w_{k-1} \ge m_+ +\eta _{k-1}\) in \(\Omega \). Then since \(\eta _{k-1} \ge \eta _k\) by (8.6), this implies that \(w_{k-1} \ge m_+ + \eta _k=\underline{w}_k\), and moreover we have

$$\begin{aligned} \mathcal{E}_k w_{k-1}&= - \nabla \cdot \left( w_{k-1} \nabla (w_{k-1}-m)\right) - w_{k-1}(m-w_{k-1}) - \frac{1}{k-1}\\&= \frac{1}{k}- \frac{1}{k-1}>0 \quad \text{ in } \Omega \end{aligned}$$

as well as

$$\begin{aligned} \mathcal{B}_k w_{k-1} = w_{k-1} \frac{\partial }{\partial \nu } (w_{k-1}-m) - \frac{1}{k-1} = \frac{1}{k}-\frac{1}{k-1} \quad \text{ on } \partial \Omega . \end{aligned}$$

Therefore, \(\underline{w}_k\) and \(w_{k-1}\) form a properly ordered pair of sub- and supersolutions for (8.7) on the sense of Nagumo, whence again we infer the existence of a solution lying in between, that is, of a classical solution \(w_k\in C^2(\bar{\Omega })\) fulfilling (8.8). \(\square \)

Lemma 8.3

The solutions \(w_k\) of (8.4) constructed in Lemma 8.2 satisfy

$$\begin{aligned} w_k\rightarrow m_+ \quad \text{ in } C^0(\bar{\Omega }) \quad \text{ as } k\rightarrow \infty . \end{aligned}$$
(8.13)

Proof

Since \(w_k>0\) in \(\bar{\Omega }\) by (8.5), we may test (8.4) by \(\ln w_k\) to find that

$$\begin{aligned} \int _\Omega |\nabla w_k|^2 = \int _\Omega \nabla w_k\cdot \nabla m + \int _\Omega w_k(m-w_k) \ln w_k + \frac{1}{k} \int _\Omega \ln w_k + \frac{1}{k} \int _{\partial \Omega } \ln w_k.\nonumber \\ \end{aligned}$$
(8.14)

Now from (8.5) we obtain that there exists \(c_1>0\), independent of \(k\in \mathbb {N}\), such that \(w_k \le c_1\) in \(\Omega \). Since \(\int _\Omega \nabla w_k \cdot \nabla m \le \frac{1}{2} \int _\Omega |\nabla w_k|^2 + \frac{1}{2} \int _\Omega |\nabla m|^2\) by Young’s inequality, it thus follows from (8.14) that \((w_k)_{k\in \mathbb {N}}\) is bounded in \(W^{1,2}(\Omega )\). In light of the ordering in (8.5), this means that there exists \(w\in W^{1,2}(\Omega ) \cap L^\infty (\Omega )\) such that \(w_k\searrow w\) in \(\bar{\Omega }\) and \(w_k \rightharpoonup w\) in \(W^{1,2}(\Omega )\) as \(k\rightarrow \infty \), where the former combined with Beppo Levi’s theorem implies that also \(w_k \rightarrow w\) in \(L^p(\Omega )\) as \(k\rightarrow \infty \) for any \(p\in (1,\infty )\). Given any \(\psi \in C^1(\bar{\Omega })\), in the identity

$$\begin{aligned} \int _\Omega w_k \nabla (w_k-m) \cdot \nabla \psi = \int _\Omega w_k(m-w_k)\psi + \frac{1}{k} \int _\Omega \psi + \frac{1}{k} \int _{\partial \Omega }\psi , \end{aligned}$$

gained from (8.4) on testing by \(\psi \), we may thus let \(k\rightarrow \infty \) in each integral separately to conclude that \(w\) is a weak solution of the limit problem (1.3) in the sense of Definition 5.1. However, since \(w\ge m_+\) in \(\Omega \) by (8.5), in view of Proposition 2.2 this already means that \(w\equiv m_+\) in \(\Omega \). In particular, \(w\) is continuous in \(\bar{\Omega }\), and hence Dini’s theorem asserts that the monotone convergence \(w_k\rightarrow w\) is even uniform in \(\bar{\Omega }\). \(\square \)

Lemma 8.4

The stationary weak solution \(m_+\) of (1.3) is stable from above in \(X_\infty \).

Proof

Given \(\delta >0\), from Lemma 8.3 we obtain some large \(k_\delta \in \mathbb {N}\) such that the solution \(w_{k_\delta }\) of (8.4) constructed in Lemma 8.2 there satisfies

$$\begin{aligned} w_{k_\delta } \le m_+ +\delta \quad \text{ in } \Omega . \end{aligned}$$
(8.15)

Since \(w_{k_\delta }>m_+\) in \(\Omega \) by (8.5), the number

$$\begin{aligned} \eta _\delta := \inf _{x\in \Omega } (w_{k_\delta }(x)-m_+(x)) \end{aligned}$$
(8.16)

is positive. Now let \(u_0 \in C^0(\bar{\Omega })\) be nonnegative and such that \(u_0 \le m_+ + \eta _\delta \) in \(\Omega \). Then \(u_0 \le w_{k_\delta }\) in \(\Omega \) by (8.16), and since \(w_{k_\delta }\) clearly is a stationary weak supersolution of (1.3), by comparison we conclude that the solution \(u\) of (1.3) satisfies \(u(\cdot ,t)\le w_{k_\delta }\) in \(\Omega \) and hence, according to (8.15), \(u(\cdot ,t) \le m_+ + \delta \) in \(\Omega \) for all \(t>0\). \(\square \)

Summarizing, we obtain the announced result on stability properties of all weak steady states of (1.3).

Proof of Proposition 2.5

The claim is part of the statements of Proposition 8.1 and Lemma 8.4. \(\square \)

1.2 Domains of attraction

We now investigate the attractivity properties of the weak steady states of (1.3) in more detail.

Proof of Proposition 2.6

Let \(\sigma =(\sigma _i)_{i\in \mathbb {N}} \subset \{0,1\}\), and suppose that \(u_0 \in C^0(\bar{\Omega })\) is nonnegative and such that \(u_0 \le m_\sigma \) and \(u_0 \not \equiv 0\) in each component \(G_i\) of \(\{m>0\}\) with \(\sigma _i=1\). Then according to the comparison principle in Lemma 4.2 we have \(u(\cdot ,t) \le m_\sigma \) in \(\Omega \) for all \(t>0\). Thus it follows that we have \(u\not \equiv 0\) in \(G_i \times [0,\infty )\) for some \(i\in \mathbb {N}\) if and only if \(\sigma _i=1\). Therefore Theorem 2.3 says that \(u(\cdot ,t) \rightarrow m_\sigma \) in \(X\) as \(t\rightarrow \infty \). \(\square \)

The proof of Proposition 2.7 is also immediate.

Proof of Proposition 2.7

Since \(X_+:=\{\psi \in X \ | \ \psi >0 \text{ in } \bar{\Omega }\}\) is dense in \(X\), (2.8) implies that \(\mathcal{D}(m_+)\) contains the dense set \(X_+\) and thus itself is dense in \(X\). On the other hand, if \(w\ne m_+\) then again using (2.8) it can easily be checked that any \(u_0\in X_+\) does not belong to the closure of \(\mathcal{D}(w)\). \(\square \)

Some more efforts are necessary in the study of openness.

Proof of Proposition 2.8

For definiteness, in giving details we fix \(X=X_\infty \); the case \(X=X_1\) can be addressed using minor modifications.

First, since for each \(\delta >0\) the function \(w+\delta \) belongs to \(\mathcal{D}(m_+)\) by (2.8), it follows that \(\mathcal{D}(w)\) is not open in \(X\) whenever \(w\ne m_+\).

Next, assume that \(w=m_+\) and that \(\{m>0\}\) has infinitely many components, whence in the notation of Proposition 2.2 we have \(G_0 \ne \emptyset \) for infinitely many \(i\in \mathbb {N}\). Then if \(\mathcal{D}(m_+)\) were open in \(X\), there would exist \(\delta >0\) such that

$$\begin{aligned} \{ u_0\in X \Vert \ \Vert u_0-m_+\Vert _{L^\infty (\Omega )} < \delta \} \, \subset \, \mathcal{D}(m_+). \end{aligned}$$
(8.17)

Now it is easy to check that by continuity of \(m\) in \(\bar{\Omega }\) we must necessarily have \(\Vert m\Vert _{L^\infty (G_i)} \rightarrow 0\) as \(i\rightarrow \infty \). It is therefore possible to fix some large \(i\in \mathbb {N}\) such that \(G_i \ne \emptyset \) and \(\Vert m\Vert _{L^\infty (G_i)} \le \frac{\delta }{2}\). Then

$$\begin{aligned} u_0 := m_+ \cdot \chi _{\Omega {\setminus } G_i} \end{aligned}$$

clearly coincides with one of the weak steady states identified in Proposition 2.2, and hence the corresponding solution \(u\) of (1.3) satisfies \(u(\cdot ,t)\equiv u_0\) for all \(t>0\). Since \(G_i \ne \emptyset \), however, we have \(u_0 \not \equiv m_+\) in \(\Omega \), so that \(u_0\) does not belong to \(\mathcal{D}(m_+)\). This contradiction to (8.17) shows that \(\mathcal{D}(m_+=\) is not open in \(X\) in this case.

It remains to be verified that \(\mathcal{D}(m_+)\) is open in \(X\) if \(\{m>0\}\) consists of at most finitely many components only, where in the case \(\{m>0\}=\emptyset \) this is obvious, because then Theorem 2.3 says that \(\mathcal{D}(m_+)=X\). Thus, still referring to the notation from Proposition 2.2, we may assume that for some \(n\in \mathbb {N}, \{m>0\}\) precisely consists of the nonempty and mutually disjoint components \(G_1,\ldots ,G_n\). We then fix and arbitrary \(u_0\in \mathcal{D}(m_+)\) and let \(u\) denote the solution of (1.3) evolving from \(u_0\). In order to construct a suitable neighborhood of \(u_0\) which is fully contained in \(\mathcal{D}(m_+)\), we introduce the positive numbers

$$\begin{aligned} \mu _i:=\int _{G_i} m, \quad i\in \{1,\ldots ,n\}, \quad \text{ and } \quad \mu :=\min _{i\in \{1,\ldots ,n\}} \mu _i. \end{aligned}$$

Then since \(u_0 \in \mathcal{D}(m_+)\) implies that \(u(\cdot ,t) \rightarrow m_+\) with respect to the norm in \(L^1(\Omega )\), we can find \(t_0>0\) such that

$$\begin{aligned} \Vert u(\cdot ,t_0)-m_+\Vert _{L^1(\Omega )} \le \frac{\mu }{2}. \end{aligned}$$

In particular, this means that

$$\begin{aligned} \int _{G_i} u(\cdot ,t_0) \ge \int _{G_i} m - \Vert u(\cdot ,t_0)-m_+\Vert _{L^1(\Omega )} \ge \mu _i - \frac{\mu }{2} \ge \frac{\mu }{2} \quad \text{ for } \text{ all } i\in \{1,\ldots ,n\}.\nonumber \\ \end{aligned}$$
(8.18)

Now let

$$\begin{aligned} \delta :=\frac{\mu }{4|\Omega |} e^{-M_0 t_0} \end{aligned}$$

with \(M_0:=\Vert m_+\Vert _{L^\infty (\Omega )}\) as in Lemma 4.2. Then for any \(\tilde{u}_0 \in X\) fulfilling \(\Vert \tilde{u}_0-u_0\Vert _{L^\infty (\Omega )} < \delta \), according to Corollary 4.4 the corresponding solution \(\tilde{u}\) of (1.3) satisfies

$$\begin{aligned} \Vert \tilde{u}(\cdot ,t_0)-u(\cdot ,t_0)\Vert _{L^1(\Omega )}&\le e^{M_0 t_0} \Vert \tilde{u}_0-u_0\Vert _{L^1(\Omega )} \\&\le e^{M_0 t_0} \cdot |\Omega | \cdot \Vert \tilde{u}_0-u_0\Vert _{L^\infty (\Omega )} \\&\le \frac{\mu }{4}. \end{aligned}$$

Therefore, using (8.18) we find that

$$\begin{aligned} \int _{G_i} \tilde{u}(\cdot ,t_0) \ge \int _{G_i} u(\cdot ,t_0) - \Vert \tilde{u}(\cdot ,t_0)-u(\cdot ,t_0)\Vert _{L^1(\Omega )} \ge \frac{\mu }{2} - \frac{\mu }{4} >0, \end{aligned}$$

which entails that \(\tilde{u}(\cdot ,t_0) \not \equiv 0\) in \(G_i\) for each \(i\in \{1,\ldots ,n\}\). In light of Theorem 2.3, this means that as \(t\rightarrow \infty \) we have \(\tilde{u}(\cdot ,t)\rightarrow m\) in \(C^0(\bar{G}_i)\) for all \(i\in \{1,\ldots ,n\}\) and thus implies that \(\tilde{u}(\cdot ,t) \rightarrow m_+\) in \(C^0(\bar{\Omega })\), because \(\{m>0\}=\bigcup _{i=1}^n G_i\). Accordingly, we conclude that indeed \(\tilde{u}_0 \in \mathcal{D}(m_+)\) for any such \(\tilde{u}_0\), and that hence \(\mathcal{D}(m_+)\) is open in \(X\). \(\square \)

Rights and permissions

Reprints and permissions

About this article

Cite this article

Cosner, C., Winkler, M. Well-posedness and qualitative properties of a dynamical model for the ideal free distribution. J. Math. Biol. 69, 1343–1382 (2014). https://doi.org/10.1007/s00285-013-0733-z

Download citation

  • Received:

  • Revised:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00285-013-0733-z

Keywords

Mathematics Subject Classification (2010)

Navigation