1 Introduction

How strong does a degrading term need to be in order to rule out chemotactic collapse in a Keller–Segel system? Or, phrased differently, when is chemotactic aggregation stronger than even superlinear dampening?

On the one hand, in the absence of any degrading terms the minimal Keller–Segel model, proposed in the 1970s [20] to model chemotaxis, that is, the directed movement of, for instance, cells or bacteria toward a chemical signal, and given by

$$\begin{aligned} {\left\{ \begin{array}{ll} u_t = \Delta u - \nabla \cdot (u \nabla v), \\ v_t = \Delta v - v + u, \end{array}\right. } \end{aligned}$$
(1.1)

admits solutions blowing up in finite time.

While all solutions are global and bounded in one dimension [31], blow-up does occur in two- [12, 14, 32] and higher-dimensional [46] domains. For a broader introduction to (1.1) and similar systems, we refer to the surveys [1, 26].

On the other hand, adding logistic terms to (1.1), for instance, in order to model population dynamics [13, 33] or tumor invasion processes [3], leads to the system

$$\begin{aligned} {\left\{ \begin{array}{ll} u_t = \Delta u - \nabla \cdot (u \nabla v) + \lambda u - \mu u^{1+\kappa }, \\ v_t = \Delta v - v + u, \end{array}\right. } \end{aligned}$$

where \(\lambda , \mu > 0\) and \(\kappa = 1\) are given parameters. Here, all solutions are global and bounded in two [30] and, provided that \(\mu \ge \mu _0\) for some \(\mu _0 > 0\) depending on the space dimension, also in higher dimensions [44]. (See also [38] for the corresponding parabolic–elliptic system.) Moreover, without any restriction on \(\mu > 0\), in all space dimensions global weak solutions have been constructed, which in three-dimensional convex domains additionally become eventually smooth if \(\lambda \) is small enough [24].

In order to better understand the relative strengths of the possibly explosion-enhancing cross-diffusive chemotaxis effect and the damping force of the logistic terms, also systems with weaker damping (e.g., \(\kappa <1\)) have been investigated with respect to the existence of classical, weak or generalized solutions (see, e.g., [29, 40, 41, 53, 54]).

For a more complete answer, however, it seems indispensable to also search for the opposite case, that of blow-up: What happens for \(\kappa \in (0, 1)\) (or for \(\kappa = 1\) and small \(\mu > 0\))? For which values of \(\kappa \) can solutions blowing up in finite time be constructed?

Even beyond solutions that grow on smaller timescales in case of slow diffusion [23, 47], some partial results in this direction are available: The first blow-up result for a chemotaxis system with superlinear degradation apparently goes back to [45]. Following a simplification introduced in [19] by Jäger and Luckhaus, there it was shown that for the system

$$\begin{aligned} {\left\{ \begin{array}{ll} u_t = \Delta u - \nabla \cdot (u \nabla v) + \lambda u - \mu u^{1+\kappa }, \\ 0 = \Delta v - \overline{M}(t) + u, \quad \overline{M}(t) = \frac{1}{|\Omega |} \int _\Omega u \end{array}\right. } \end{aligned}$$
(1.2)

in a ball in \(\mathbb {R}^n\), \(n \ge 5\), finite-time blow-up is possible provided that \(\lambda , \mu > 0\) and \(\kappa < \frac{1}{2} + \frac{1}{2n-2}\). Moreover, chemotactic collapse may even happen in the physically (most) relevant space dimension three. In [49], it was shown that the parabolic–elliptic system

$$\begin{aligned} {\left\{ \begin{array}{ll} u_t = \Delta u - \nabla \cdot (u \nabla v) + \lambda u - \mu u^{1+\kappa }, \\ 0 = \Delta v - v + u \end{array}\right. } \end{aligned}$$
(1.3)

in a ball in \(\mathbb {R}^n\), \(n \ge 3\), admits solutions blowing up in finite time provided \(\lambda , \mu > 0\) and

$$\begin{aligned} \kappa < \kappa _{\mathrm {pe}}(n) {:}{=}{\left\{ \begin{array}{ll} \frac{1}{6}, &{} n \in \{3, 4\}, \\ \frac{1}{2n-2}, &{} n > 4. \end{array}\right. } \end{aligned}$$

This result has also recently been transferred to a setting where two species are attracted by the same signal in [39], where only limited adaptations were necessary to retrieve the same values for \(\kappa _{\mathrm {pe}}(n)\), and to a system with a weaker sensitivity function, [36].

In two-dimensional domains, however, known results seem to be limited to the case of space-dependent functions \(\mu \). That is, if one replaces the first equation in (1.2) with \(u_t = \Delta u - \nabla \cdot (u \nabla v) + \lambda u - \mu _1 |x|^\alpha u^{1+\kappa }\), then solutions blowing up in finite-time have been constructed if again the domain is a ball, \(\lambda , \mu _1, \alpha > 0\) and \(\kappa < \frac{\alpha }{2}\). Phrased differently, given any \(\kappa > 0\), there exist blow-up solutions even in 2D— provided \(\alpha \) is large enough [10].

There is another effect that can hinder blow-up and is often included in the model, be it for reasons of biological modelling, for example, of tumor cells, cf. [35] or [11, 21], or from a purely mathematical motivation: nonlinear, porous-medium type diffusion (i.e., the replacement of \(\Delta u\) by, e.g., \(\nabla \cdot ((u+1)^{m-1}\nabla u)\)). If sufficiently strong, it can prevent blow-up even on its own (see [15, 17, 34, 37] for boundedness results in case of \(m>2-\frac{2}{n}\)) or at least in combination with logistic sources, [27, 42, 55,56,57].

However, for the regime of slightly weaker diffusion, the occurrence of blow-up may still be possible. And indeed, in the absence of logistics (\(\kappa =0\), \(\mu =\lambda \)), \(m<\frac{2n-2}{n}\) for \(\Omega \) being a ball in \({\mathbb {R}}^n\) allows for some unbounded solutions (i.e., blow-up after either finite or infinite time, see [15, 43]), with finite-time blow-up having been detected in [6]. As to blow-up for different combinations of diffusion and sensitivity terms, refer to [4, 5, 50] or to [16, 18] for the case of degenerate diffusion.

Parabolic–elliptic analogues were investigated in [7, 25, 52], revealing blow-up after finite or infinite time for different parameter ranges.

If logistics and nonlinear diffusion both are incorporated into the model, at least in space dimensions \(n\ge 5\) sometimes blow-up is possible: In [28], it was shown that the techniques of [45] can be applied for diffusion rates \(m\in [1,\frac{2n-4}{n})\) and dampening exponents \(\kappa \in (0,\frac{mn}{2(n-1)})\), resulting in finite-time blow-up for some radial solutions of the system

$$\begin{aligned} {\left\{ \begin{array}{ll} u_t = \nabla \cdot ((u+1)^{m-1} \nabla u - (u \nabla v)) + \lambda (|x|) u - \mu (|x|) u^{1+\kappa }, &{} \hbox { in}\ \Omega \times (0, T_{\max }) \\ 0 = \Delta v - \overline{M}(t) + u, \quad \overline{M}(t) {:}{=}\frac{1}{|\Omega |} \int _\Omega u, &{} \hbox { in}\ \Omega \times (0, T_{\max }) \\ \partial _\nu u = \partial _\nu v = 0, &{} \hbox { on}\ \partial \Omega \times (0, T_{\max })\\ u(\cdot , 0) = u_0, &{} \hbox { in}\ \Omega , \\ \end{array}\right. } \end{aligned}$$
(JL)

in which the signal’s equation is simplified analogously to the system from the famous work [19] by Jäger and Luckhaus. In [28, Remark 1.2], it was conjectured that blow-up should occur for \(m<\frac{2n-2}{n}\) but the answer was left open to further research. For an earlier extension of the same methods from [45] to systems with nonlinear diffusion and logistics, combined with superlinear sensitivity functions, see [59].

The recent advances of [49] in the linear-diffusion case with logistic raise some hope that also in a nonlinear setting, the discovery of blow-up is also possible in the slightly less simplified parabolic–elliptic system

$$\begin{aligned} {\left\{ \begin{array}{ll} u_t = \nabla \cdot ((u+1)^{m-1} \nabla u - (u \nabla v)) + \lambda (|x|) u - \mu (|x|) u^{1+\kappa }, &{} \hbox { in}\ \Omega \times (0, T_{\max }) \\ 0 = \Delta v - v + u, &{} \hbox { in}\ \Omega \times (0, T_{\max }) \\ \partial _\nu u = \partial _\nu v = 0, &{} \hbox { on}\ \partial \Omega \times (0, T_{\max })\\ u(\cdot , 0) = u_0, &{} \hbox { in}\ \Omega \\ \end{array}\right. } \end{aligned}$$
(PE)

and, more importantly, even in the physically more relevant case of \(n=3\), for instance.

Main results The present article is dedicated to this question. Aiming for blow-up, we study (JL) and (PE) in a ball \(\Omega \subset \mathbb {R}^n\), \(n \ge 3\), for sufficiently smooth nonnegative functions \(\lambda , \mu \) and a parameter \(\kappa \ge 0\). We refer to the introduction of [10] for a motivation for logistic source terms with spatial dependence.

We extend the methods of [49] to nonlinear diffusion and show that they are applicable in (JL) as well as in (PE). At the same time, we would like to offer a different perspective on these, seeing them as a bridge connecting pointwise upper estimates of solutions to the occurrence of explosions. We therefore give our main result in the following form:

Theorem 1.1

Let \(n \ge 3\), \(\Omega {:}{=}B_R(0)\), \(R > 0\), \(M_0 > 0\), \(M_1\in (0,M_0)\), \(\alpha \ge 0\), \(\mu _1 > 0\), \(p \ge n\), \(T > 0\) as well as \(K > 0\) and suppose that \(\lambda , \mu \) are such that

$$\begin{aligned} 0 \le \lambda \in C^{1+\beta }([0, R]) \quad \text {and} \quad 0 \le \mu \in C^\beta ([0, R]) \cap C^{1+\beta }((0, R]) \qquad \text {for some }\beta \in (0, 1) \end{aligned}$$
(1.4)

and comply with

$$\begin{aligned} \mu (r) \le \mu _1 r^\alpha \qquad \text {for all }r \in [0, R]. \end{aligned}$$
(1.5)

Assume moreover that \(\kappa \ge 0\) and \(m >0\) satisfy

$$\begin{aligned} m<1+\frac{n-2}{p} \end{aligned}$$

and

$$\begin{aligned} 0 \le \kappa&< \frac{\alpha }{p} + \min \left\{ \frac{n}{2p}, \frac{n - 2}{p} - (m-1)_+ \right\}&\qquad \text {if } m\ge \frac{2}{p}\; \end{aligned}$$
(1.6)
$$\begin{aligned} \text { or }\quad 0\le \kappa&<\frac{\alpha }{p}+\min \left\{ \frac{n}{2p},\frac{n-1}{p}-\frac{m}{2}\right\}&\qquad \text {if } m\in \left( 0,\frac{2}{p}\right) . \end{aligned}$$
(1.7)

Then, we can find \(r_1 \in (0, R)\) with the following property: If

$$\begin{aligned} (u, v)\in \left( C^0(\overline{\Omega }\times [0,T_{\max }))\times C^{2,1}(\overline{\Omega }\times (0,T_{\max }))\right) ^2 \end{aligned}$$

is a classical solution to (JL) or (PE) in \({\overline{\Omega }}\times [0, T_{\max })\) for some \(T_{\max }\in (0, \infty ]\) with

$$\begin{aligned} u_0 \in \bigcup _{\beta \in (0, 1)} {C^{\beta }(\overline{\Omega })}\quad \text {being nonnegative, radially symmetric and radially decreasing} \end{aligned}$$
(1.8)

and

$$\begin{aligned} \int \limits _{\Omega } u_0 = M_0 \quad \text {but} \quad \int \limits _{B_{r_1}(0)} u_0 \ge M_1 \end{aligned}$$
(1.9)

as well as

$$\begin{aligned} \sup _{t \in (0, \min \{T, T_{\max }\})} u(x, t) \le K |x|^{-p} \qquad \text {for all }x \in \Omega , \end{aligned}$$
(1.10)

then (uv) blows up in finite time in the sense that \(T_{\max }< \infty \) and

$$\begin{aligned} \limsup _{t \nearrow T_{\max }} \Vert u(\cdot , t)\Vert _{{L^{\infty }(\Omega )}} = \infty . \end{aligned}$$
(1.11)

If upper estimates as in (1.10) are known, this theorem shows that finite-time blow-up is possible in (PE) and (JL); that is, there are initial data such that (1.11) holds with some \(T_{\max }<\infty \). This results in the following:

Theorem 1.2

Let \(n \ge 3\), \(\Omega {:}{=}B_R(0)\), \(R > 0\), \(M_0 > 0\), \(\alpha \ge 0\), \(\mu _1 > 0\), \(p \ge n\) and

$$\begin{aligned} m < \frac{2n-2}{n}. \end{aligned}$$

Assume moreover that \(\lambda , \mu \) satisfy (1.4) and (1.5).

(i) Suppose additionally \(-\lambda ', \mu ' \ge 0\) in (0, R). If

$$\begin{aligned} 0 \le \kappa&< \frac{\alpha }{n} + \min \left\{ \frac{1}{2}, \frac{n - 2}{n} - (m-1)_+ \right\}&\qquad \text {if } m\ge \frac{2}{n} \end{aligned}$$
(1.12)
$$\begin{aligned} \text { or }\quad 0 \le \kappa&<\frac{\alpha }{n}+\min \left\{ \frac{1}{2},\frac{n-1}{n}-\frac{m}{2}\right\}&\qquad \text {if } m\in \left( 0,\frac{2}{n}\right) , \end{aligned}$$
(1.13)

then there exists \(0 \le u_0 \in {C^{0}(\overline{\Omega })}\) with \(\int _\Omega u_0 = M_0\) leading to finite-time blow-up; that is, there is a classical solution (uv) to (JL) fulfilling (1.11) for some finite \(T_{\max }\).

Moreover, for \(m = 1\), the condition (1.12)–(1.13) is equivalent to

$$\begin{aligned} 0 \le \kappa < {\left\{ \begin{array}{ll} \frac{1}{3} + \frac{\alpha }{3}, &{} n = 3, \\ \frac{1}{2} + \frac{\alpha }{n}, &{} n \ge 4. \end{array}\right. } \end{aligned}$$

(ii) Provided that \(m \ge 1\) and

$$\begin{aligned} 0 \le \kappa < \frac{\alpha [(m-1)n + 1]}{n(n-1)} + \min \left\{ \frac{(m-1) n + 1}{2(n-1)}, \frac{n - 2 - (m-1) n}{n(n-1)} \right\} , \end{aligned}$$
(1.14)

an initial datum \(0 \le u_0 \in {C^{0}(\overline{\Omega })}\) with \(\int _\Omega u_0 = M_0\) and a solution (uv) of (PE) satisfying (1.11) for some \(T_{\max }< \infty \) can be constructed.

Furthermore, for \(m = 1\), the condition (1.14) reads

$$\begin{aligned} 0 \le \kappa < {\left\{ \begin{array}{ll} \frac{1}{6} + \frac{\alpha }{6}, &{} n = 3, \\ \frac{1}{2(n-1)} + \frac{\alpha }{n(n-1)}, &{} n \ge 4. \end{array}\right. } \end{aligned}$$

Before we provide a more detailed comparison to the conditions on the existence of solutions blowing up in finite-time established in previous works, let us note the following.

Remark 1.3

  1. (i)

    The condition \(m < \frac{2n-2}{n}\) in Theorem 1.2 is optimal: Even without any degradation terms (i.e., \(\lambda = \mu \equiv 0\)), classical solutions to (JL) and (PE) are always global in time for \(m > \frac{2n-2}{n}\) (cf. [52] and [25], respectively). In particular, for (JL) the upper bound on the admissible diffusion exponents in Theorem 1.2 (i) coincides with the conjectured critical exponent in [28, Remark 1.2].

  2. (ii)

    The assumption \(p \ge n\) in Theorem 1.1 is a natural limitation. In fact, since (1.10) implies \(\int _\Omega u^q \le K^q \omega _{n-1} \int \limits _0^R r^{n-1 + pq} \,\mathrm {d}r< C(q)\) for all \(q < \frac{n}{p}\) and certain \(C(q) > 0\), assuming that (1.10) holds for some \(p < n\) and a large class of initial data, these initial data would automatically be uniformly bounded in \({L^{\frac{n+p}{2p}}(\Omega )}\), say, by \(C'\). However, as can be seen by applying Hölder’s inequality, their mass on \(B_{r_1}(0)\) would then be bounded by \(C' |B_{r_1}(0)|^\frac{n-p}{n+p}\), which converges to 0 for \(r_1 \searrow 0\). Thus, it would not be clear if one of these initial data could still fulfill (1.9) for the value of \(r_1\) given by Theorem 1.1.

  3. (iii)

    To the best of our knowledge, Theorem 1.2 provides the first detection of finite-time blow-up for Keller–Segel systems with nonlinear diffusion and superlinear damping terms in space dimensions 3 and 4. For (PE) and \(m\ne 1\), it is furthermore the first such result in higher dimensions.

  4. (iv)

    The finite-time blow-up result for (PE) also constitutes a partial answer to the second part of Open Problem (i) in [58].

Now, let us take a more in-depth look at the new ranges for the parameter \(\kappa \) in some different spatial dimensions under the assumption of \(\alpha =0\) for some special values of m. In this setting, earlier works have established a certain \(\kappa ^*\) (provided in Table 1) for which blow-up has been proven for \(\kappa <\kappa ^*\).

Table 1 Supremum of the range of \(\kappa \) for which finite-time blow-up has been detected

Evidently, the findings of [45] and [28] (and also of the related [59]) only cover higher dimensions. For \(n\ge 5\) and \(m\in [1,\frac{2n-4}{n})\), however, these results still provide better ranges than the one we could attain with our method. To the best of our knowledge, for larger values of m or for small space dimensions, however, our results provide the first proof of finite-time blow-up in (JL).

Regarding (PE), for the linear diffusion case we are able to match the range previously established in [49], while also providing first results for the nonlinear diffusion setting in higher dimensions.

When comparing the parameter ranges across the two different systems for \(m = 1\) and \(n \in \{2, 3\}\), we see that our results for (JL) yield a wider regime for \(\kappa \) than the corresponding results obtained in [49] for (PE). Indeed, \(\frac{1}{3} > \frac{1}{6}\) and \(\frac{1}{2} > \frac{1}{6}\). In general, known results for (JL) are stronger than for (PE). However, lacking global existence results for \(\kappa < 2\), it is yet unclear whether blow-up is actually more prominent in (JL) or just easier to detect.

Main ideas As is meanwhile well established in the context of finite-time blow-up proofs for chemotaxis systems and has first been proposed by Jäger and Luckhaus in [19], we consider the mass accumulation function

$$\begin{aligned} w(s, t) {:}{=}\int \limits _0^{s^\frac{1}{n}} \rho ^{n-1} u(\rho , t) \,\mathrm {d}\rho , \quad s \in [0, R^n], t \in [0, T_{\max }), \end{aligned}$$

which transforms (JL) into the scalar equation

$$\begin{aligned} w_t&= n^2 s^{2-\frac{2}{n}} w_{ss} + n w w_s - n \overline{m}(t) s w_s + n \int \limits _0^s \lambda (\sigma ^\frac{1}{n}) w_s(\sigma , t) \,\mathrm {d}\sigma - n \int \limits _0^s \mu (\sigma ^\frac{1}{n}) w_s^{\kappa +1}(\sigma , t) \,\mathrm {d}\sigma \end{aligned}$$

(and (PE) at least into a system that is easier to handle than (PE) itself). The main difficulty for detecting finite-time blow-up lies in the fact that the term \(+n w w_s\), stemming from the cross-diffusion in (JL), has to counter the quite different terms \(n^2 s^{2-\frac{2}{n}} w_{ss}\) and \(- n \int \limits _0^s \mu (\sigma ^\frac{1}{n}) w_s^2(\sigma , t) \,\mathrm {d}\sigma \) originating from the diffusion and logistic terms, respectively.

Following [49], our approach consists of showing that for certain initial data, \(\gamma \in (0, 1)\) and \(s_0 \in (0, R^n)\), the function

$$\begin{aligned} \phi (s_0,\cdot ): [0, T_{\max }) \rightarrow \mathbb {R}, \quad t \mapsto \int \limits _0^{s_0}s^{-\gamma } (s_0-s) w(s, t) \,\mathrm {d}s\end{aligned}$$

cannot exist globally in time, which due to the blow-up criterion asserted in Lemma 2.1 implies the desired finite-time blow-up result (1.11). That is, in Sect. 3 we show that \(\phi \) is a supersolution to the ODI \(\phi ' = a \phi ^2 - b\) for certain \(a,b > 0\) and in Sect. 4, we conclude the existence of initial data leading to finite-time blow-up of \(\phi \) and hence u.

Let us briefly discuss how we deal with the two most problematic terms stemming from the degradation and diffusion terms, respectively. As we will see in Lemma 3.5, in order to handle the former, we essentially need to control

$$\begin{aligned} -\int \limits _0^{s_0} s^\frac{\alpha }{n} (s_0-s) w_s^{1+\kappa }(s, t) \,\mathrm {d}s. \end{aligned}$$
(1.15)

At this point, the assumption (1.10) comes into play, which due to \(w_s(s, t) = \frac{u(s^\frac{1}{n}, t)}{n}\) for \((s, t) \in [0, R^n] \times [0, T_{\max })\) implies \(w(s, t) \le \frac{C}{n} s^\frac{p}{n}\) for \((s, t) \in [0, R^n] \times [0, T_{\max })\). Thus, as a starting point, we can apply this estimate to \(w_s^\kappa \) in (1.15) and then integrate by parts. Moreover, by (3.11), the term arising from the diffusion can be estimated against (some positive multiple of)

$$\begin{aligned} -\int \limits _0^{s_0}s^{1-\frac{2}{n}-\gamma }(s_0-s)(nw_s+1)^m\,\mathrm {d}s. \end{aligned}$$

For \(m \ge 1\), we can proceed as above; that is, we apply the pointwise upper bound to \(w_s^{m-1}\) and then integrate by parts, while for \(m < 1\) we can follow at least two different paths: For \(m \in (0, \frac{2}{p})\), we apply this bound to \(w^m\) and do not integrate by parts and for \(m \ge \frac{2}{p}\), we estimate \((n w_s + 1)^m \le n w_s + 1\) and integrate by parts without using the pointwise upper estimate for \(w_s\) at all. The fact that depending on the value of m we employ two different methods here is the reason for the different conditions in (1.6) and (1.7).

At last, we show that pointwise upper estimates of the form (1.10) are indeed available both for (JL) and (PE). While for the former system we make use of the comparison principle applied to \(u_r\) in Lemma 5.1, for the latter we resort to the recent study on blow-up profiles [9] to obtain the desired bounds in Lemma 5.2.

2 Preliminaries

We henceforth always assume \(n \ge 3\) and \(\Omega {:}{=}B_R(0) \subset \mathbb {R}^n\) for some \(R > 0\). Furthermore, in Sects. 24, we also fix \(m > 0\), \(\kappa \ge 0\), \(\alpha \ge 0\), \(M_0>0\), \(M_1\in (0,M_0)\), \(\lambda _1 > 0\) as well as functions \(\lambda , \mu \) complying with \(\lambda \le \lambda _1\), (1.4) and (1.5). To simplify the notation, we also fix an initial datum \(u_0\) satisfying (1.8) with \(\int _\Omega u_0 = M_0\), but emphasize that all constants below, unless otherwise stated, are independent of \(u_0\).

By (uv), we will refer to a solution to either of the systems (JL) or (PE), and we also set \(\overline{M}(t) {:}{=}|\Omega |^{-1}\int _\Omega u(\cdot ,t)\) for \(t \in [0, T_{\max })\).

Lemma 2.1

Suppose that \(u_0 :{\overline{\Omega }}\rightarrow [0, \infty )\) is Hölder continuous. Then, for each of the systems (JL) and (PE) there exist \(T_{\max }\in (0, \infty ]\) and a classical solution (uv), uniquely determined by

$$\begin{aligned} u \in C^0(\overline{\Omega }\times [0, T_{\max })) \cap C^{2, 1}(\overline{\Omega }\times (0, T_{\max })), \end{aligned}$$
(2.1)
$$\begin{aligned} v \in \bigcap _{q > n} C^0([0, T_{\max }); W^{1, q}(\Omega )) \cap C^{2, 1}(\overline{\Omega }\times (0, T_{\max })) \end{aligned}$$

and, in case of (JL),

$$\begin{aligned} \int \limits _{\Omega } v(\cdot , t) = 0 \quad \hbox { for all}\ t \in (0, T_{\max }). \end{aligned}$$

Moreover, \(u \ge 0\) in \(\Omega \times (0, T_{\max })\) and if \(T_{\max }< \infty \), then

$$\begin{aligned} \limsup _{t \nearrow T_{\max }} \Vert u(\cdot , t)\Vert _{{L^{\infty }(\Omega )}} = \infty . \end{aligned}$$

If, finally, \(u_0\) is radially symmetric, then so are \(u(\cdot ,t)\) and \(v(\cdot ,t)\) for any \(t\in (0,T_{\max })\).

Proof

Local existence can be proved by a standard fixed point argument, which is explained in more detail in [7] or [38], for instance, while nonnegativity of u follows by the maximum principle and preservation of radial symmetry is a consequence of uniqueness. \(\square \)

As a first basic observation, we note that, at least locally in time, the mass of u can be controlled by the parameters we fixed above—and thus, independently of the precise choice of \(u_0\).

Lemma 2.2

For all \(t \in (0, T_{\max })\), we have

$$\begin{aligned} \int \limits _{\Omega } u(\cdot , t) \le M_0 \mathrm {e}^{\lambda _1 t}. \end{aligned}$$

Proof

Due to \(\lambda \le \lambda _1\) and nonnegativity of \(\mu \), integrating the first equation over \(\Omega \) gives

$$\begin{aligned} \frac{\mathrm {d}}{\mathrm {d}t}\int \limits _{\Omega } u = \int \limits _{\Omega } \lambda u - \int \limits _{\Omega } \mu u^{1+\kappa } \le \lambda _1 \int \limits _{\Omega } u \qquad \text {in }(0, T_{\max }), \end{aligned}$$

so that the statement follows by an ODE comparison argument. \(\square \)

3 Proving finite-time blow-up

Following [2, 19, 49], we define

$$\begin{aligned} w(s, t) {:}{=}\int \limits _0^{s^\frac{1}{n}} \rho ^{n-1} u(\rho , t) \,\mathrm {d}\rho , \quad s \in [0, R^n], t \in [0, T_{\max }) \end{aligned}$$

and, given \(s_0\in (0,R^n)\) and \(\gamma \in (0,1)\), introduce the functions

$$\begin{aligned} \phi (s_0,\cdot ): [0, T_{\max }) \rightarrow \mathbb {R}, \quad t \mapsto \int \limits _0^{s_0}s^{-\gamma } (s_0-s) w(s, t) \,\mathrm {d}s\end{aligned}$$
(3.1)

(cf. [49, equation (4.1)]) and

$$\begin{aligned} \psi (s_0,\cdot ): [0, T_{\max }) \rightarrow \mathbb {R}, \quad t \mapsto \int \limits _0^{s_0}s^{-\gamma } (s_0-s) w(s, t)w_s(s,t) \,\mathrm {d}s. \end{aligned}$$

If, by the usual slight abuse of notation, we identify the radially symmetric function \(u\in C^0({\overline{\Omega }}\times [0,T_{\max }))\) with \(u\in C^0([0,R]\times [0,T_{\max }))\) and write \(u_r\) for its radial derivative, we can compute the spatial derivatives of w:

Lemma 3.1

We have

$$\begin{aligned} w\in C^{1,0}([0,R^n]\times [0,T_{\max }))\cap C^{2,1}([0,R^n]\times (0,T_{\max }))\cap C^{3,0}((0,R^n]\times (0,T_{\max })), \end{aligned}$$
(3.2)

and

$$\begin{aligned} w_s(s,t)=\frac{1}{n} u(s^{\frac{1}{n}},t), \quad w_{ss}(s,t)=\frac{1}{n^2}s^{\frac{1}{n}-1}u_r(s^{\frac{1}{n}},t) \qquad \text {for } s\in (0,R^n], t\in (0,T_{\max }) \end{aligned}$$
(3.3)

and, with K and T from (1.10),

$$\begin{aligned} w_s(s,t)\le \frac{K}{n} s^{-\frac{p}{n}}\qquad \text {for } (s,t)\in (0,R^n]\times (0,T). \end{aligned}$$
(3.4)

Proof

For the regularity, we rely on (2.1); the final estimate (3.4) results from (3.3) and (1.10). \(\square \)

For \(\phi \), which we later want to show to blow up, the following differential inequality holds:

Lemma 3.2

For any choice of \(\gamma \in (0, 1)\) and \(s_0 \in (0, R^n)\), the function \(\phi \) of (3.1) belongs to \(C^0([0, T_{\max })) \cap C^1((0, T_{\max }))\) and fulfills

(3.5a)

for all \(t \in (0, T_{\max })\) in the case of (JL). For (PE), the same estimate holds with

$$\begin{aligned} I_3(s_0,t)=-n\int \limits _0^{s_0} s^{-\gamma }(s_0-s)w_s(s,t) z(s,t)\,\mathrm {d}s, \qquad t \in (0, T_{\max }), \end{aligned}$$
(3.5b)

where

$$\begin{aligned} z(s,t) {:}{=}\int \limits _0^{s^{\frac{1}{n}}}\rho ^{n-1}v(\rho ,t)\,\mathrm {d}\rho \qquad \text {for }s \in [0, R^n]\hbox { and }t \in [0, T_{\max }). \end{aligned}$$

Proof

The regularity of \(\phi \) follows from (3.2). Written in radial coordinates, the differential equations in (JL) read

$$\begin{aligned} {\left\{ \begin{array}{ll} u_t(r,t) &{}= r^{1-n}\left( r^{n-1}\left( (u(r,t)+1)^{m-1}u_r(r,t)-u(r,t)v_r(r,t)\right) \right) _r+\lambda (r)u(r,t)-\mu (r)u^{1+\kappa }(r,t),\\ 0 &{}= r^{1-n}\left( r^{n-1}v_r(r,t)\right) _r -\overline{M}(t) + u(r,t) \end{array}\right. } \end{aligned}$$
(3.6)

for \((r,t)\in [0,R)\times (0,T_{\max })\), where the second equation can be transformed into

$$\begin{aligned} v_r(r,t) = \overline{M}(t) r^{1-n}\int \limits _0^r \rho ^{n-1}d\rho - r^{1-n}\int \limits _0^r \rho ^{n-1}u(\rho ,t)d\rho = \frac{r}{n} \overline{M}(t) -r^{1-n} w(r^n,t), \end{aligned}$$
(3.7)

for \(r\in [0,R)\), \(t\in (0,T_{\max })\), and the first equation results in

$$\begin{aligned} w_t(s,t)&= \int \limits _0^{s^{\frac{1}{n}}}\rho ^{n-1}u_t(\rho ,t)\,\mathrm {d}\rho \\&= \int \limits _0^{s^{\frac{1}{n}}}\rho ^{n-1}\rho ^{1-n} \left( r^{n-1}\left( (u(r,t)+1)^{m-1}u_r(r,t)-u(r,t)v_r(r,t)\right) \right) _r\big \vert _{r=\rho } \,\mathrm {d}\rho \\&\quad +\int \limits _0^{s^{\frac{1}{n}}}\rho ^{n-1}\lambda (\rho )u(\rho ,t)\,\mathrm {d}\rho -\int \limits _0^{s^{\frac{1}{n}}}\rho ^{n-1}\mu (\rho )u^{1+\kappa }(\rho ,t)\,\mathrm {d}\rho \\&= s^{\frac{n-1}{n}}u_r(s^{\frac{1}{n}},t) \left( u(s^{\frac{1}{n}},t)+1\right) ^{m-1} - s^{\frac{n-1}{n}}u(s^{\frac{1}{n}},t)v_r(s^{\frac{1}{n}},t)\\&\quad +\int \limits _0^{s^{\frac{1}{n}}}\rho ^{n-1}\lambda (\rho )u(\rho ,t)\,\mathrm {d}\rho -\int \limits _0^{s^{\frac{1}{n}}}\rho ^{n-1}\mu (\rho )u^{1+\kappa }(\rho ,t)\,\mathrm {d}\rho \end{aligned}$$

for \((s, t) \in (0, R^n) \times (0, T_{\max })\). If we insert (3.3) into (3.7) and use nonnegativity of \(\lambda \), we obtain that

$$\begin{aligned} w_t \ge n^2 s^{2-\frac{2}{n}} (n w_s + 1)^{m-1} w_{ss} + n w w_s - \overline{M}(t) s w_s - n^{\kappa } \mu \int \limits _0^s \sigma ^\frac{\alpha }{n} w_s^{1+\kappa }(\sigma , t) \,\mathrm {d}\sigma \end{aligned}$$
(3.8)

in \((0, R^n) \times (0, T_{\max })\), which by multiplication with \(s^{-\gamma }(s_0-s)\) and integration implies the statement for (JL).

For (PE), in (3.6), \(-\overline{M}(t)\) has to be replaced by \(-v\), so that after essentially the same computation, in (3.8) the term \(-\overline{M}(t)sw_s\) is substituted by \(-nw_s\int \limits _0^{s^{\frac{1}{n}}} \rho ^{n-1} v(\rho ,t)\,\mathrm {d}\rho =-nw_sz\). \(\square \)

In the remaining part of this section, we further estimate the terms of the right-hand side of (3.5), aiming to show that \(\phi (s_0, \cdot )\) fulfills a certain superlinear ODE. These results will then be combined in Sect. 4; ultimately, the consolidation of the lemmata will show that at least for certain values of \(\gamma \) and \(s_0\), \(\phi (s_0, \cdot )\) cannot exist globally.

In order to streamline the arguments below, let us first state two elementary lemmata.

Lemma 3.3

For all \(\alpha >-1\) and \(\beta >-1\) and any \(s_0\ge 0\), we have

$$\begin{aligned} \int \limits _0^{s_0}s^\alpha (s_0-s)^\beta \,\mathrm {d}s=B(\alpha +1,\beta +1) s_0^{\alpha +\beta +1}. \end{aligned}$$

Proof

This is an evident consequence of the properties of the beta function. \(\square \)

Lemma 3.4

Let \(\gamma \in (0, 1)\) and \(s_0 \in (0, R^n)\). Then,

$$\begin{aligned} w(s, t) \le \sqrt{2} s^\frac{\gamma }{2} (s_0-s)^{-\frac{1}{2}} \sqrt{\psi (s_0,t)} \end{aligned}$$

holds true for all \((s, t) \in (0, s_0) \times (0, T_{\max })\).

Proof

This inequality, in its essence based on the fundamental theorem of calculus, is a direct consequence of [49, Lemma 4.2], which for every \(t\in (0,T_{\max })\) can be applied to \(\varphi =w(\cdot ,t)\), because \(\varphi (0)=0\), \(\varphi '\ge 0\) in \((0,s_0)\) and \(\varphi \in C^1([0,s_0])\). \(\square \)

The estimate (3.4), which originates in the crucial assumption (1.10), will come into play at two different places. The first of these is the following lemma, where said upper estimate is the most important ingredient for controlling the term arising from the logistic source, namely \(I_4\) in (3.5).

Lemma 3.5

Let \(\gamma \in (0, 1)\) and \(p \ge n\) satisfy \(\frac{p \kappa }{n} -\frac{\alpha }{n} < \frac{\gamma }{2}\). Whenever (1.10) is fulfilled for some \(K > 0\), \(T>0\) and \(s_0 \in (0, R^n)\), then \(I_4\) from Lemma 3.2 satisfies

$$\begin{aligned} I_4(s_0,t) \ge -K^{\kappa } C s_0^{\frac{3-\gamma }{2} - \frac{p\kappa }{n}+\frac{\alpha }{n} }\sqrt{\psi (s_0,t)} \text {for all }t \in (0, \min \{T,T_{\max }\}), \end{aligned}$$

with \(C=\frac{\mu _1\sqrt{2}((\frac{p\kappa }{n}-\frac{\alpha }{n})_++1)}{1-\gamma }B(\frac{\alpha }{n}-\frac{p\kappa }{n}+\frac{\gamma }{2},\frac{1}{2})\).

Proof

This can be proved analogously to [49, Lemma 4.5]: Firstly, Fubini’s theorem asserts

$$\begin{aligned} I_4(s_0,t)&= - c_1 \int \limits _0^{s_0}\sigma ^\frac{\alpha }{n} w_s^{1+\kappa }(\sigma , t) \left( \int \limits _\sigma ^{s_0} s^{-\gamma } (s_0-s) \,\mathrm {d}s\right) \,\mathrm {d}\sigma \\&\ge - c_1 \int \limits _0^{s_0} \sigma ^\frac{\alpha }{n} (s_0-\sigma ) w_s^{1+\kappa }(\sigma , t) \left( \int \limits _0^{s_0} s^{-\gamma } \,\mathrm {d}s\right) \,\mathrm {d}\sigma \\&= -c_2 s_0^{1-\gamma } \int \limits _0^{s_0} s^\frac{\alpha }{n} (s_0-s) w_s^{1+\kappa }(s, t) \,\mathrm {d}s\end{aligned}$$

for all \(t \in (0, T_{\max })\), where \(I_4\) is as in (3.5), \(c_1 {:}{=}n^{\kappa } \mu _1 > 0\) and \(c_2 {:}{=}\frac{c_1}{1-\gamma } > 0\). Next, we use (3.4) and integrate by parts to see that

holds for all \(t \in (0, \min \{T,T_{\max }\})\) with \(c_3 {:}{=}\frac{(\frac{p \kappa }{n} - \frac{\alpha }{n})_+ + 1}{n^{\kappa }} > 0\). Here, we apply Lemma 3.4 and obtain

$$\begin{aligned} \int \limits _0^{s_0}s^{\frac{\alpha }{n} - \frac{p \kappa }{n} - 1} w(s, t) \,\mathrm {d}s&\le \sqrt{2} \int \limits _0^{s_0}s^{\frac{\alpha }{n} -\frac{p \kappa }{n} - 1 + \frac{\gamma }{2}} (s_0-s)^{-\frac{1}{2}} \,\mathrm {d}s\sqrt{\psi (s_0,t)} \end{aligned}$$

for all \(t \in (0, \min \{T,T_{\max }\})\). Finally, we note that according to Lemma 3.3

$$\begin{aligned} \int \limits _0^{s_0}s^{\frac{\alpha }{n} - \frac{p \kappa }{n} - 1 + \frac{\gamma }{2}} (s_0-s)^{-\frac{1}{2}} \,\mathrm {d}s= B\left( \frac{\alpha }{n}-\frac{p\kappa }{n} +\frac{\gamma }{2},\frac{1}{2}\right) s_0^{\frac{\alpha }{n} - \frac{p \kappa }{n} + \frac{\gamma }{2}-\frac{1}{2}}, \end{aligned}$$

since \(\frac{\alpha }{n} - \frac{p \kappa }{n} + \frac{\gamma }{2} > 0\) by assumption. The statement follows by combining the estimates above. \(\square \)

We now turn our attention to the integral involving the effects of nonlinear diffusion. This is the second place where (at least for certain m) we make use of the assumption (1.10).

Lemma 3.6

Suppose that (1.10) holds for some \(p \ge n\), \(K > 0\) and \(T>0\) and let \(I_1\) be as in (3.5).

(i) Assume that

$$\begin{aligned} 0<m<1+\frac{n-2}{p}\quad&\text {and}\quad 1 - \frac{2}{n} - \frac{p}{n} (m-1)_+<\gamma < 2 - \frac{4}{n} - \frac{2p}{n} (m-1)_+. \end{aligned}$$
(3.9)

Then, there is \(C > 0\) such that for any \(s_0 \in (0, R^n)\) and all \(t\in (0, \min \{T,T_{\max }\})\), we have

$$\begin{aligned} I_1(s_0,t)\ge - C s_0^{\frac{3-\gamma }{2} - \frac{2}{n} - \frac{p}{n}(m-1)_+} \sqrt{\psi (s_0,t)} -C s_0^{3-\gamma -\frac{2}{n}}. \end{aligned}$$

(ii) Assume that

$$\begin{aligned} 0<m<\min \left\{ 1,\frac{2(n-1)}{p}\right\} \quad&\text {and}\quad 0<\gamma <2-\frac{2}{n}-\frac{pm}{n}. \end{aligned}$$
(3.10)

Then, there is \(C > 0\) such that for any \(s_0 \in (0, R^n)\) and all \(t\in (0, \min \{T,T_{\max }\})\), we have

$$\begin{aligned} I_1(s_0,t)\ge -C s_0^{3-\gamma -\frac{2}{n}-\frac{p}{n}m}-C s_0^{3-\gamma -\frac{2}{n}}. \end{aligned}$$

Proof

Direct calculation gives for every \(s_0 \in (0, R^n)\)

in \((0, T_{\max })\). The last two terms therein are positive, since (3.9) and (3.10) both entail \(2 - \frac{2}{n} - \gamma > 0\), leading to

$$\begin{aligned} I_1(s_0, t)\ge -\frac{n}{m}\left( 2-\frac{2}{n}-\gamma \right) \int \limits _0^{s_0}s^{1-\frac{2}{n}-\gamma }(s_0-s)(nw_s+1)^m\,\mathrm {d}s\qquad \text {for }(s_0,t)\in (0,R^n)\times (0,T_{\max }). \end{aligned}$$
(3.11)

Now, let us start by considering the case that (3.9) holds. First, we find that \(w_s\ge 0\) implies for \(m\ge 1\) that \((nw_s+1)^m\le 2^{m-1}(n^m w_s^m+1)\) in \((0,T_{\max })\). By (3.4), this entails that \((nw_s+1)^m\le 2^{m-1} K^{m-1} n s^{-\frac{p}{n}(m-1)}w_s+2^{m-1}\) on \((0,R^n)\times (0,\min \{T,T_{\max }\})\). On the other hand, for \(m\in (0,1)\) we have \((nw_s+1)^m\le n w_s+1\) in \((0,R^n)\times (0,\min \{T,T_{\max }\})\). Thus, letting \(c_1{:}{=}\max \{n,2^{m-1},2^{m-1} n K^{(m-1)_+}\}\) we find that from combining these two estimates, we have

$$\begin{aligned} (n w_s+1)^m\le c_1 s^{-\frac{p}{n}(m-1)_+ } w_s+c_1\qquad \text {in }(0,R^n)\times (0,\min \{T,T_{\max }\}) \end{aligned}$$

and hence, from (3.11),

The third term on the right-hand side is nonnegative, and in the other terms, we use \(s_0 - s \le s_0\) and \(s \le s_0\) for all \(s \in (0, s_0)\) as well as the conditions \(\gamma < 2 - \frac{2}{n}\) and \(\gamma > 1 - \frac{2}{n} - \frac{p}{n} (m-1)_+\) contained in (3.9) to see that

$$\begin{aligned} I_1(s_0, t)&\ge - c_1 \left( 2 - \frac{2}{n} - \gamma \right) \left( \gamma + \frac{2}{n} + \frac{p}{n} (m-1)_+ \right) s_0 \int \limits _0^{s_0}s^{- \frac{2}{n} - \frac{p}{n}(m-1)_+ - \gamma } w \,\mathrm {d}s-c_1 s_0^{3-\gamma -\frac{2}{n}} \end{aligned}$$
(3.12)

in \((0, \min \{T,T_{\max }\})\) and for \(s_0\in (0,R^n)\). To estimate further, we make use of Lemma 3.4, the fact that (3.9) entails \(0<1-\frac{2}{n}-\frac{p}{n}(m-1)_+-\frac{\gamma }{2}\) and Lemma 3.3 to obtain

(3.13)

for \(c_3=\sqrt{2}B(1-\frac{2}{n}-\frac{p}{n}(m-1)_+-\frac{\gamma }{2},\frac{1}{2}) > 0\). Collecting (3.12) and (3.13) proves the estimate of \(I_1\) in the case that (3.9) holds.

To verify the asserted inequality in the case of (3.10), we return to (3.11) and note that due to \(w_s\ge 0\) and \(m\in (0,1)\), we have \((nw_s+1)^m\le n^m w_s^m+1\) on \((0,\min \{T,T_{\max }\})\). Here, we rely on (3.4) to conclude that \((nw_s+1)^m\le K^ms^{-\frac{p m}{n}}+1\) in \((0,\min \{T,T_{\max }\})\) and hence

$$\begin{aligned} I_1(s_0, t)&\ge -\frac{K^m n}{m}\left( 2-\frac{2}{n}-\gamma \right) \int \limits _0^{s_0}s^{1-\frac{2}{n}-\frac{p m}{n}-\gamma }(s_0-s)\,\mathrm {d}s-\frac{n}{m}\left( 2-\frac{2}{n}-\gamma \right) \int \limits _0^{s_0}s^{1-\frac{2}{n}-\gamma }(s_0-s)\,\mathrm {d}s\end{aligned}$$

in \((0,\min \{T,T_{\max }\})\) and for \(s_0\in (0,R^n)\), where the conditions \(\gamma <2-\frac{2}{n}-\frac{p m}{n}\) and \(\gamma <2-\frac{2}{n}\) contained in (3.10) together with \(s_0-s\le s_0\) entail

$$\begin{aligned} I_1(s_0, t)\ge -c_4 s_0^{3-\gamma -\frac{2}{n}-\frac{p m}{n}}-\frac{n}{m} s_0^{3-\gamma -\frac{2}{n}}\quad \text {in }(0,\min \{T,T_{\max }\}) \end{aligned}$$

for \(c_4=\frac{2-\frac{2}{n}-\gamma }{2-\frac{2}{n}-\frac{pm}{n}-\gamma }\cdot \frac{K^mn}{m} >0\), completing the proof. \(\square \)

The arguments for estimating the remaining integrals in (3.5) rely on the following relation between \(\phi \) and \(\psi \), which was also obtained in [49, Lemma 3.4].

Lemma 3.7

Let \(\gamma \in (0,1)\). For every \(s_0\in (0,R^n)\) and \(t\in (0,T_{\max })\),

$$\begin{aligned} \phi (s_0,t)\le Cs_0^{\frac{3-\gamma }{2}} \sqrt{\psi (s_0,t)}, \end{aligned}$$

where \(C=\sqrt{2} B(1-\frac{\gamma }{2},\frac{1}{2})\).

Proof

By Lemma 3.4,

$$\begin{aligned} \phi (s_0,t)\le s_0\int \limits _0^{s_0} s^{-\gamma }w(s,t)\,\mathrm {d}s\le s_0\sqrt{2}\sqrt{\psi }\int \limits _0^{s_0}s^{-\frac{\gamma }{2}}(s_0-s)^{-\frac{1}{2}}\,\mathrm {d}s, \end{aligned}$$

and the claim follows from Lemma 3.3. \(\square \)

Lemma 3.8

Let \(\gamma \in (0, 1)\). Then, for all \(s_0 \in (0, R^n)\) and all \(t \in (0, T_{\max })\)

$$\begin{aligned} I_2(s_0,t) \ge C s_0^{\gamma -3} \phi ^2(s_0,t) \end{aligned}$$

with \(\phi \) and \(I_2\) as in (3.5) and \(C=\frac{n}{2B^2(1-\frac{\gamma }{2},\frac{1}{2})}\).

Proof

As \(I_2=n\psi \), this is a corollary of Lemma 3.7. \(\square \)

Lemma 3.9

Let \(\gamma \in (0, 1)\) and \(T > 0\). Then, there is \(C > 0\) such that for all \(s_0 \in (0, R^n)\) and all \(t \in (0, \min \{T, T_{\max }\})\)

$$\begin{aligned} I_3(s_0,t)\ge -Cs_0^{\frac{2}{n}+1-\gamma } - C s_0^{\frac{2}{n}}\psi (s_0,t) \end{aligned}$$
(3.14)

with \(I_3\) as in (3.5).

Remark 3.10

For (JL), we can even obtain the stronger estimate

$$\begin{aligned} I_3(s_0,t) \ge - C \mathrm {e}^{\lambda _1 t} s_0^{\frac{3-\gamma }{2}} \sqrt{\psi (s_0,t)}\ge -Cs_0^{\frac{2}{n}+3-\gamma }\mathrm {e}^{2\lambda _1t} - C s_0^{\frac{2}{n}}\psi (s_0,t) \end{aligned}$$
(3.15)

in place of (3.14).

Proof of Lemma 3.9 and Remark 3.10

As their expressions for \(I_3\) differ, we treat the cases of (JL) and (PE) separately, beginning with (JL): An integration by parts and Lemma 3.4 show

$$\begin{aligned} I_3(s_0,t)&= \overline{M}(t) (1-\gamma ) \int \limits _0^{s_0}s^{-\gamma } (s_0-s) w(s, t) \,\mathrm {d}s- \overline{M}(t) \int \limits _0^{s_0}s^{1-\gamma } w(s, t) \,\mathrm {d}s- 0 \\&\ge - \overline{M}(t)\sqrt{2} \int \limits _0^{s_0}s^{1-\frac{\gamma }{2}} (s_0-s)^{-\frac{1}{2}} \,\mathrm {d}s\cdot \sqrt{\psi (s_0,t)} \end{aligned}$$

for \(t \in (0, T_{\max })\) and \(s_0\in (0,R^n)\), so that with some \(c_1>0\), \(c_2>0\),

$$\begin{aligned} I_3\ge -\frac{\sqrt{2}M_0}{|\Omega |} \mathrm {e}^{\lambda _1t}c_1 s_0^{\frac{3-\gamma }{2}}\sqrt{\psi (s_0,t)}\ge -c_2 \mathrm {e}^{2\lambda _1t}s_0^{3-\gamma +\frac{2}{n}}-c_2 s_0^{\frac{2}{n}}\psi (s_0,t) \end{aligned}$$

for \(t\in (0,T_{\max })\), \(s_0\in (0,R^n)\) by Lemmata 2.23.3 and Young’s inequality. This proves (3.15) and, since \(s_0^{3-\gamma +\frac{2}{n}}\le R^{2n} s_0^{1-\gamma +\frac{2}{n}}\) for \(s_0\in (0,R^n)\), also (3.14).

As to (PE), we follow [49, Lemma 4.8] but need to make some modifications to remove the additional condition \(\gamma <2-\frac{4}{n}\) required there. To that end, we first fix \(\tilde{\gamma }\in (\max \{\gamma - \frac{4}{n}, 0\}, \min \{\gamma , 2-\frac{4}{n}\})\) and apply a variant of [49, Lemma 4.7] (for \({\tilde{\gamma }} \in (0, 2 - \frac{4}{n})\) instead of \(\gamma \)) to obtain \(\Gamma =\Gamma (R,\lambda ,\gamma ,M_0)>0\) such that

$$\begin{aligned} z(s, t)&\le \Gamma s_0^{\frac{2}{n} - 1} s + \Gamma s_0^{-\frac{1}{2}} s^{\frac{2}{n} + \frac{{\tilde{\gamma }}}{2}} \left( \int \limits _0^{s_0} \sigma ^{-{\tilde{\gamma }}} (s_0 - \sigma ) w(\sigma , t) w_s(\sigma , t) \,\mathrm {d}\sigma \right) ^\frac{1}{2} \\&\le \Gamma s_0^{\frac{2}{n} - 1} s + \Gamma s_0^{-\frac{1}{2}} s^{\frac{2}{n} + \frac{{\tilde{\gamma }}}{2}} \left( \int \limits _0^{s_0} s_0^{\gamma -{\tilde{\gamma }}}\sigma ^{-\gamma } (s_0 - \sigma ) w(\sigma , t) w_s(\sigma , t) \,\mathrm {d}\sigma \right) ^\frac{1}{2} \\&= \Gamma s_0^{\frac{2}{n} - 1} s + \Gamma s_0^{-\frac{1}{2} + \frac{\gamma -{\tilde{\gamma }}}{2}} s^{\frac{2}{n} + \frac{\tilde{\gamma }}{2}} \sqrt{\psi (s_0, t)} \qquad \text {for all }s_0 \in (0, R^n), t \in (0, \min \{T, T_{\max }\}). \end{aligned}$$

Combined with (4.5) of [49], Lemmata 2.23.4 and 3.3, this shows that for any \(s_0\in (0,R^n)\), \(t\in (0, \min \{T, T_{\max }\})\),

$$\begin{aligned} \frac{I_3(s_0,t)}{n (\gamma +1)}&\ge - s_0\int \limits _0^{s_0}s^{-\gamma -1} z(s,t)w(s,t)\,\mathrm {d}s\\&\ge - M_0 \Gamma \mathrm {e}^{\lambda _1 t} s_0^\frac{2}{n} \int \limits _0^{s_0}s^{-\gamma } \,\mathrm {d}s- \Gamma s_0^{\frac{1}{2} + \frac{\gamma -{\tilde{\gamma }}}{2}} \int \limits _0^{s_0} s^{\frac{2}{n} - 1 -\gamma + \frac{{\tilde{\gamma }}}{2}} w(s, t) \,\mathrm {d}s\cdot \sqrt{\psi (s_0, t)} \\&\ge - \frac{M_0 \Gamma \mathrm {e}^{\lambda _1 T}}{1-\gamma } s_0^{1+\frac{2}{n}-\gamma } - \Gamma s_0^{\frac{1}{2} + \frac{\gamma -{\tilde{\gamma }}}{2}} \int \limits _0^{s_0} s^{\frac{2}{n} - 1 - \frac{\gamma - {\tilde{\gamma }}}{2}} (s_0 - s)^{-\frac{1}{2}} \,\mathrm {d}s\cdot \psi (s_0, t) \\&\ge - \frac{M_0 \Gamma \mathrm {e}^{\lambda _1 T}}{1-\gamma } s_0^{1+\frac{2}{n}-\gamma } - \Gamma B\left( \frac{2}{n} - \frac{\gamma - {\tilde{\gamma }}}{2}, \frac{1}{2}\right) s_0^{\frac{2}{n}} \psi (s_0, t), \end{aligned}$$

which again implies (3.14). \(\square \)

Combining these lemmata shows that \(\phi (s_0, \cdot )\) is indeed a supersolution to a superlinear ODE as long as \(s_0\) is sufficiently small and \(\gamma \) can be chosen in a suitable way.

Lemma 3.11

Let \(p \ge n\), \(K > 0\) and \(T > 0\). (i) Suppose \(m\in [\frac{2}{p},1+\frac{n-2}{p})\), \(0\le \kappa <\frac{\alpha }{p}+\min \{\frac{n}{2p},\frac{n-2}{p}-(m-1)_+\}\) and

$$\begin{aligned} \max \left\{ 0,\frac{2p\kappa }{n}-\frac{2\alpha }{n},1-\frac{2}{n}-\frac{p}{n}(m-1)_+\right\}<\gamma <\min \left\{ 2-\frac{4}{n}-\frac{2p}{n}(m-1)_+,1\right\} . \end{aligned}$$
(3.16)

Then, there are \(C_1,C_2>0\), \(\theta \in (0,2)\) and \(s_1\in (0,R^n)\) such that for every solution complying with (1.10), we have

$$\begin{aligned} \frac{\partial }{\partial t}\phi (s_0,t) \ge C_1 s_0^{\gamma -3} \phi ^2(s_0,t) - C_2 s_0^{3-\gamma - \theta } \qquad \text {for }t \in (0, \min \{T,T_{\max }\})\hbox { and } s_0\in (0,s_1). \end{aligned}$$
(3.17)

(ii) Suppose \(m\in (0,\frac{2}{p})\), \(0\le \kappa <\frac{\alpha }{p}+\min \{\frac{n}{2p},\frac{n-1}{p}-\frac{m}{2}\}\) and

$$\begin{aligned} \max \left\{ 0,\frac{2p\kappa }{n}-\frac{2\alpha }{n}\right\}<\gamma <\min \left\{ 2-\frac{2}{n}-\frac{pm}{n},1\right\} . \end{aligned}$$
(3.18)

Then, there are \(C_1,C_2>0\) , \(\theta \in (0,2)\) and \(s_1\in (0,R^n)\) such that (3.17) holds for every solution complying with (1.10) also in this case.

Proof

To verify part (i), we note that the prescribed conditions on \(\kappa \) and \(\gamma \) render Lemma 3.5 and the first part of Lemma 3.6 applicable, in addition to Lemmata 3.8 and 3.9. Inserting their respective results into (3.5) of Lemma 3.2, we find that there are \(c_1, c_2>0\) such that

is valid for all \(t \in (0,\min \{T,T_{\max }\})\) and for every \(s_0\in (0,R^n)\). Young’s inequality shows that thus for every \(\eta >0\), there is \(c_3(\eta )>0\) satisfying

$$\begin{aligned} \frac{\partial }{\partial t}\phi (s_0,t)\ge&c_2 s_0^{\gamma -3}\phi ^2(s_0,t)-c_1s_0^{\frac{2}{n}}\psi (s_0,t)-\eta \psi (s_0,t)\\&-c_3(\eta )\left( s_0^{3-\gamma -\frac{4}{n}-\frac{2p}{n}(m-1)_+}+s_0^{3-\gamma -\frac{2}{n}}+s_0^{1+\frac{2}{n}-\gamma }+s_0^{3-\gamma -\frac{2p\kappa }{n}+\frac{2\alpha }{n}}\right) \end{aligned}$$

for all \((0,\min \{T,T_{\max }\})\). If we employ Lemma 3.7, fix \(s_1\) and \(\eta \) sufficiently small and use that \(s_0\le s_1\le R^n\), we obtain (3.17) with

$$\begin{aligned} 3-\gamma -\theta =\min \left\{ 3-\gamma -\frac{4}{n}-\frac{2p}{n}(m-1)_+,3-\gamma -\frac{2}{n},1 +\frac{2}{n}-\gamma ,3-\gamma -\frac{2p\kappa }{n}+\frac{2\alpha }{n}\right\} . \end{aligned}$$

Observing that \(\frac{4}{n}+\frac{2p}{n}(m-1)_+<\frac{4}{n}+\frac{2p}{n}\cdot \frac{n-2}{p}=2\) due to \(m < 1 + \frac{n-2}{p}\) and \(\frac{2p\kappa }{n}-\frac{2\alpha }{n}<\frac{2\alpha +n}{n}-\frac{2\alpha }{n}=1<2-\frac{2}{n}<2\) because of \(\kappa < \frac{\alpha }{p} + \frac{n}{2p}\), we conclude

$$\begin{aligned} \theta =\max \left\{ \frac{4}{n}+\frac{2p}{n}(m-1)_+,\frac{2}{n},2-\frac{2}{n},\frac{2p\kappa }{n}-\frac{2\alpha }{n}\right\} \in (0,2). \end{aligned}$$

As to part (ii), similarly as before with Lemma 3.6, part (ii), in place of Lemma 3.6, part (i), we obtain

for all \((0,\min \{T,T_{\max }\})\) and \(s_0\in (0,R^n)\) and conclude as before, with

$$\begin{aligned} \theta =\max \left\{ \frac{2}{n}+\frac{pm}{n},\frac{2}{n},2-\frac{2}{n},\frac{2p\kappa }{n}-\frac{2\alpha }{n}\right\} , \end{aligned}$$

where again \(\frac{2p\kappa }{n}-\frac{2\alpha }{n}<\frac{2\alpha +n}{n}-\frac{2\alpha }{n}<2\) and \(\frac{2}{n}+\frac{pm}{n}<\frac{2}{n}+\frac{2}{n}<2\), so that \(\theta \in (0,2)\). \(\square \)

Let us close this section by verifying that the parameter choices of Lemma 3.11 are indeed feasible.

Lemma 3.12

Under the conditions on m and \(\kappa \) in Lemma 3.11, part (i) or (ii), there is \(\gamma >0\) satisfying (3.16) or (3.18), respectively. Moreover, for every choice of m as indicated there, the respective ranges of \(\kappa \) are nonempty.

Proof

(i) First, we note that due to \(p \ge n > 2\), the interval \(A{:}{=}[\frac{2}{p},1+\frac{n-2}{p})\) is not empty. For \(m\in A\), due to \(m<1+\frac{n-2}{p}\) and \(\alpha \ge 0\), we then see that \(B{:}{=}[0,\frac{\alpha }{p}+\min \{\frac{n}{2p},\frac{n-2}{p}-(m-1)_+\})\) is also not empty. Now, to check that for \(m\in A\) and \(\kappa \in B\), the condition (3.16) is not empty, we first note that \(\kappa <\frac{\alpha }{p}+\frac{n-2}{p}-(m-1)_+\) and \(m<1+\frac{n-2}{p}\) imply that

$$\begin{aligned}&\frac{2p\kappa }{n}-\frac{2\alpha }{n}<2-\frac{4}{n}-\frac{2p}{n}(m-1)_+\quad \text {and}\quad 1-\frac{2}{n}-\frac{p}{n}(m-1)_+<2-\frac{4}{n}-\frac{2p}{n}(m-1)_+. \end{aligned}$$

Moreover, we find that \(\kappa <\frac{n}{2p}+\frac{\alpha }{p}\) and \(m<1+\frac{n-2}{p}\) entail that

$$\begin{aligned}&\frac{2p\kappa }{n}-\frac{2\alpha }{n}<1\quad \text {and}\quad 0<1-\frac{2}{n}-\frac{p}{n}(m-1)_+<1, \end{aligned}$$

so that indeed \(\max \{0,\frac{2p\kappa }{n}-\frac{2\alpha }{n},1-\frac{2}{n}-\frac{p}{n}(m-1)_+\}<\min \{2-\frac{4}{n}-\frac{2p}{n}(m-1)_+,1\}\).

(ii) We first note that \(m<\frac{2}{p}\) ensures that \(\frac{n-1}{p}-\frac{m}{2}> \frac{n-2}{p}>0\), so that \(\kappa \) with \(0\le \kappa <\frac{\alpha }{p}+\min \{\frac{n}{2p},\frac{n-1}{p}-\frac{m}{2}\}\) exists. From \(\kappa <\frac{\alpha }{p}+\frac{n}{2p}\), we conclude that \(\frac{2p\kappa }{n}-\frac{2\alpha }{n}< \frac{2p}{n}(\frac{\alpha }{p}+\frac{n}{2p})-\frac{2\alpha }{n}=1\), and the condition that \(\kappa <\frac{\alpha }{p}+\frac{n-1}{p}-\frac{m}{2}\) shows that \(\frac{2p\kappa }{n}-\frac{2\alpha }{n}<\frac{2p}{n}(\frac{\alpha }{p}+\frac{n-1}{p}-\frac{m}{2})-\frac{2\alpha }{n}=2-\frac{2}{n}-\frac{pm}{n}\). Together with the fact that \(2-\frac{2}{n}-\frac{pm}{n}>2-\frac{2}{n}-\frac{2}{n}\ge 0\), this shows that \( \max \left\{ 0,\frac{2p\kappa }{n}-\frac{2\alpha }{n}\right\} <\min \left\{ 2-\frac{2}{n}-\frac{pm}{n},1\right\} . \) \(\square \)

4 First conclusion: Proof of Theorem 1.1

While Lemmata 3.11 and 3.12 already show that \(\phi (s_0, \cdot )\) is (for certain \(s_0\) and \(\gamma \), at least) a supersolution to a superlinear ODE, we still need to show that \(\phi (s_0, 0)\) can be arranged to be suitably large. We take care of this last step in the following lemma; Theorem 1.1 will then be proven directly thereafter.

Lemma 4.1

Let \(\gamma \in (0, 1)\), \(s_0 \in (0, R), M_1 \ge 0\) as well as \(\eta \in (0, 1)\) and set \(s_{\eta } {:}{=}(1-\eta ) s_0\) as well as \(r_1 {:}{=}s_{\eta }^{\frac{1}{n}}\). If

$$\begin{aligned} \int \limits _{B_{r_1}(0)} u_0 \ge M_1, \end{aligned}$$

then

$$\begin{aligned} \phi (s_0,0) \ge \frac{\eta ^2M_1}{\omega _{n-1}}\cdot s_0^{2-\gamma }. \end{aligned}$$

Proof

We use positivity and monotonicity of \(w_0{:}{=}w(\cdot ,0)\), \(w_0(s_{\eta })=\frac{1}{\omega _{n-1}}\int \limits _{B_{r_1}}u_0\) and \(s_0-s_1=\eta s_0\) as well as the fact that \(1-(1-\eta )^{1-\gamma } \ge \inf _{\xi \in (0, \eta )} (1-\gamma )(1-\xi )^{-\gamma } \eta = (1-\gamma )\eta \) holds by the mean value theorem, to see that

$$\begin{aligned} \phi (s_0,0)&= \int \limits _0^{s_0}s^{-\gamma } (s_0 - s) w_0(s) \,\mathrm {d}s\ge w_0(s_{\eta }) \int \limits _{s_{\eta }}^{s_0} s^{-\gamma } (s_0 - s_{\eta }) \,\mathrm {d}s\\&\ge \frac{\eta M_1}{(1-\gamma ) \omega _{n-1}} s_0 \left( s_0^{1-\gamma } - s_{\eta }^{1-\gamma } \right) \ge \frac{\eta ^2 M_1}{\omega _{n-1}} \cdot s_0^{2-\gamma }. \end{aligned}$$

\(\square \) \(\square \)

These preparations now allow us to indeed prove Theorem 1.1.

Proof of Theorem 1.1

We observe that for \(a,b,y_0>0\) with \(\sqrt{\frac{a}{b}}y_0>1\), the solution

$$\begin{aligned} y(t)=\sqrt{\frac{b}{a}} \cdot \frac{1+\frac{\sqrt{\frac{a}{b}} y_0 -1}{\sqrt{\frac{a}{b}} y_0 +1} \mathrm {e}^{2\sqrt{ab}t}}{1-\frac{\sqrt{\frac{a}{b}} y_0 -1}{\sqrt{\frac{a}{b}} y_0 +1} \mathrm {e}^{2\sqrt{ab}t}} \quad \text {of}\quad {\left\{ \begin{array}{ll} y' = a y^2 - b, \\ y(0) = y_0 \end{array}\right. } \end{aligned}$$

blows up at the finite time t with \(\mathrm {e}^{2\sqrt{ab}t}=\frac{\sqrt{\frac{a}{b}} y_0 +1}{\sqrt{\frac{a}{b}} y_0 -1}\). If \(\sqrt{\frac{a}{b}}y_0>2\), then \(\frac{\sqrt{\frac{a}{b}} y_0 +1}{\sqrt{\frac{a}{b}} y_0 -1}<\frac{\sqrt{\frac{a}{b}} y_0 +\frac{1}{2} \sqrt{\frac{a}{b}} y_0}{\sqrt{\frac{a}{b}} y_0 -\frac{1}{2} \sqrt{\frac{a}{b}} y_0}=3\) and blow-up hence occurs before time \(\frac{1}{2\sqrt{ab}}\ln 3\). Given m, n, p, \(\kappa \), \(M_0\), \(M_1\), K, T from Theorem 1.1, we use Lemma 3.12 to find \(\gamma \in (0,1)\) such that Lemma 3.11 is applicable and let \(C_1\) and \(C_2\), \(\theta \) and \(s_1\) be as defined there. We furthermore set \(C_3{:}{=}\frac{M_1}{4\omega _{n-1}}\) and introduce

$$\begin{aligned} a(s_0)=C_1s_0^{\gamma -3},\quad b(s_0)=C_2 s_0^{3-\gamma -\theta }\quad \text {and} \quad \phi _0(s_0)=C_3s_0^{2-\gamma }. \end{aligned}$$

By positivity of \(\theta \), \(\sqrt{ab}\rightarrow \infty \) as \(s_0\rightarrow 0\); additionally, \(\phi _0^2\frac{a}{b}\rightarrow \infty \) as \(s_0\rightarrow 0\), because the exponent of \(s_0\) in this expression is negative according to

$$\begin{aligned} 2(2-\gamma )+\gamma -3-(3-\gamma -\theta ) =-2+\theta <0. \end{aligned}$$

We therefore can pick \(s_0\in (0,s_1)\) so small that \(\frac{\ln 3}{2\sqrt{a(s_0)b(s_0)}}<T\) and \(\phi _0\sqrt{\frac{a(s_0)}{b(s_0)}}>2\), and finally let \(r_1=(\frac{1}{2} s_0)^{\frac{1}{n}}\). Then, by Lemma 3.11 in conjunction with Lemma 4.1 for \(\eta =\frac{1}{2}\), for every solution obeying (1.8), (1.9) and (1.10), the function \(\phi {:}{=}\phi (s_0,\cdot )\) from (3.1) satisfies

$$\begin{aligned} \phi ^{\prime }(t)\ge a \phi ^2(t) - b \quad \text {for every } t\in (0,\min \{T,T_{\max }\}),\qquad \phi (0)\ge \phi _0 \end{aligned}$$

and hence \(\phi (t)\ge y(t)\) for all \(t\in (0,\min \{T,T_{\max }\})\), which implies \(T_{\max }<\frac{\ln 3}{2\sqrt{a(s_0)b(s_0)}}<T\) and (1.11). \(\square \)

5 Pointwise upper estimates for u: Proof of Theorem 1.2

The goal of this section is to prove Theorem 1.2. To that end, we first derive estimates of the form (1.10) both for (JL) (Lemma 5.1) and (PE) (Lemma 5.2) and then apply Theorem 1.1.

Lemma 5.1

Assume that \(\lambda , \mu \) not only comply with (1.4) and (1.5) but additionally satisfy \(-\lambda ', \mu ' \ge 0\). Let moreover \(\kappa \ge 0\), \(m > 0\), \(T, M_0 > 0\) and suppose that \(u_0\) satisfies (1.8) and \(\int _\Omega u_0 = M_0\). Then, every solution \((u, v)\in \left( C^0({\overline{\Omega }}\times [0, T))\cap C^{2,1}({\overline{\Omega }}\times (0, T))\right) ^2\) of (JL) fulfills

$$\begin{aligned} u(r, t) \le \frac{M_0 n \mathrm {e}^{\lambda t}}{\omega _{n-1}} \cdot r^{-n} \quad \text {for all }r \in (0, R)\hbox { and }t \in (0, T). \end{aligned}$$

Proof

As in [10, Lemma 3.7] (cf. also [48, Lemma 2.2]), we employ the comparison principle to see that u remains radially decreasing throughout evolution. To that end, we first note that by an approximation argument as in [48, Lemma 2.2], we may without loss of generality assume \(\mu \in C^2([0, R])\) and \(u_0 \in {C^{2}(\overline{\Omega })}\) with \(\partial _\nu u_0 = 0\) on \(\partial \Omega \). For henceforth fixed \(T \in (0, T_{\max })\), these assumptions then assert that \(u_r\) belongs to \(C^0([0, R] \times [0, T)) \cap C^{2, 1}([0, R] \times (0, T))\): Indeed, elliptic regularity (cf. [8, Theorem 1.19.1]) asserts \(\nabla v \in L^\infty (\Omega \times (0, T))\), which in conjunction with [22, Theorem IV.5.3] implies \(u \in C^{1, 0}({\overline{\Omega }}\times [0, T)) \cap C^{3, 1}({\overline{\Omega }}\times (0, T))\).

Thus, setting \(f(s) {:}{=}(s + 1)^{m-1}\) for \(s \ge 0\), and re-interpreting (JL) in the form of \(u_t=\nabla \cdot (f(u)\nabla u)-\nabla u\cdot \nabla v-u(\overline{M} - u) +\lambda u-\mu u^{1+\kappa }\) as an equation for the radial function \(u\in C^{1,0}([0,R]\times [0,T))\cap C^{3,1}([0,R]\times (0,T))\), we may compute the radial derivative

$$\begin{aligned} u_{rt}&= \left( (f(u) u_r)_r + \frac{n-1}{r} f(u) u_r - u_r v_r + u^2 - \overline{M}(t) u + \lambda u - \mu u^{1+\kappa } \right) _r \nonumber \\&= (f(u) u_r)_{rr} + a_1 (f(u) u_r)_r + a_2 u_{rr} + b u_r + c \qquad \hbox { in}\ (0, R) \times (0, T), \end{aligned}$$
(5.1)

where

$$\begin{aligned} a_1(r, t)&{:}{=}\frac{n-1}{r}, \qquad a_2(r, t) {:}{=}- v_r(r, t), \qquad c(r, t) {:}{=}\lambda '(r) u(r, t) - \mu '(r) u^{1+\kappa }(r, t)\quad \text {and}\\ b(r, t)&{:}{=}- \frac{n-1}{r^2}f(u(r,t)) - v_{rr}(r, t) + 2 u(r, t) - \overline{M}(t) + \lambda (r) - (1+\kappa ) \mu (r) u^\kappa (r, t) \end{aligned}$$

for \((r, t) \in (0, R) \times (0, T)\). Note that \(-\lambda ', \mu ' \ge 0\) imply \(c \le 0\) in \((0, R) \times (0, T)\). Moreover, by the second equation in (JL), \((r^{n-1} v_r(r, t))_r \le r^{n-1} \overline{M}(t)\) and hence \(v_r(r, t) \le \frac{r}{n} \overline{M}(t)\) for \((r, t) \in (0, R) \times (0, T)\). This implies

$$\begin{aligned} -v_{rr}(r, t) = u(r, t) - \overline{M}(t) + \frac{n-1}{r} v_r(r, t) \le u(r, t) - \frac{1}{n} \overline{M}(t) \le u(r, t) \qquad \text {for }(r, t) \in (0, R) \times (0, T), \end{aligned}$$

so that by setting \(c_1 {:}{=}\sup _{(r, t) \in (0, R) \times (0, T)} (3u(r, t) + \lambda (r))\), we obtain

$$\begin{aligned} b(r, t) \le c_1 \qquad \hbox { for}\ (r, t) \in (0, R) \times (0, T). \end{aligned}$$
(5.2)

Using that \(u\in C^2([0,R]\times [0,T])\), and, in particular, that \(a_1u_r=(n-1)\frac{u_r}{r}\le (n-1)u_{rr}\) is bounded in \((0,R)\times (0,T)\), we can moreover introduce

$$\begin{aligned} c_2{:}{=}\sup _{(r,t)\in (0,R)\times (0,T)} ((f(u)_{rr}+a_1(f(u))_r)<\infty \end{aligned}$$

and set \(c_3{:}{=}c_1+c_2+1\).

By (5.1), since \(u_r(r, t) = 0\) for \(r \in \{0, R\}\) and \(t \in (0, T)\) (because u is radially symmetric and (uv) solves (JL)) and as \(u_{0r} \le 0\) by assumption, the function \(y :[0, R] \times [0, T] \rightarrow \mathbb {R}, (r, t) \mapsto u_r(r, t) - \varepsilon \mathrm {e}^{c_3 t}\) belongs to \(C^0([0, R] \times [0, T)) \cap C^{2, 1}([0, R] \times (0, T))\) and fulfills

$$\begin{aligned} {\left\{ \begin{array}{ll} y_t \le (f(u) (y + \varepsilon \mathrm {e}^{c_3 t}))_{rr} + a_1 (f(u) (y + \varepsilon \mathrm {e}^{c_3 t}))_r + a_2 y_r + b y + \varepsilon \mathrm {e}^{\varepsilon t}b - c_3 \varepsilon \mathrm {e}^{c_3 t}\\ \quad =(f(u) y)_{rr} + a_1 (f(u) y)_r + a_2 y_r + b y+ \varepsilon \mathrm {e}^{c_3 t}[(f(u))_{rr}+ a_1 (f(u))_r + b - c_3]\\ \quad \le {(f(u) y)_{rr} + a_1 (f(u) y)_r + a_2 y_r + b y - \varepsilon \mathrm {e}^{c_3 t}} &{} \text {in }(0, R) \times (0, T), \\ y< 0 &{} \text {on }\{0, R\} \times (0, T), \\ y(\cdot , 0) < 0, &{} \text {in }(0, R). \end{array}\right. } \end{aligned}$$
(5.3)

By the estimate for \(y(\cdot ,0)\) in (5.3) and continuity of y, the time \(t_0 {:}{=}\sup \{\, t \in (0, T): y \le 0 \text { in } [0,R]\times (0, t) \,\} \in (0, T]\) is well defined. Suppose \(t_0 < T\), then there exists \(r_0 \in [0, R]\) such that \(y(r_0, t_0) = 0\) and \(y(r, t) \le 0\) for all \(r \in [0, R]\) and \(t \in [0, t_0]\); hence, \(y_t(r_0, t_0) \ge 0\). As \(f \ge 0\) in \([0, \infty )\), not only \(y(\cdot , t_0)\) but also \(h :(0, R) \rightarrow \mathbb {R}, r \mapsto f(u(r, t_0)) y(r, t_0) \) attains its maximum 0 at \(r_0\). Since the second inequality in (5.3) asserts \(r_0 \in (0, R)\), we conclude \(h_{rr}(r_0) \le 0\), \(h_r(r_0) = 0\) and \(y_r(r_0, t_0) = 0\). However, from the first inequality in (5.3) and (5.2), we could then infer the contradiction

$$\begin{aligned} 0&\le y_t(r_0, t_0) \\&\le h_{rr}(r_0) + a_1(r_0, t_0) h_r(r_0) + a_2(r_0, t_0) y_{r}(r_0, t_0) + b(r_0, t_0) y(r_0, t_0) -\varepsilon \mathrm {e}^{c_3 t_0} \\&\le - \varepsilon \mathrm {e}^{c_3 t_0} < 0, \end{aligned}$$

so that \(t_0 = T\), implying \(y \le 0\) in \([0, R] \times [0, T]\) and hence also \(u_r \le \varepsilon \mathrm {e}^{c_3 T}\) in \([0, R] \times [0, T]\). Letting first \(\varepsilon \searrow 0\) and then \(T \nearrow T_{\max }\), this indeed gives \(u_r \le 0\) in \([0, R] \times [0, T_{\max })\).

This, together with Lemma 2.2, implies that

$$\begin{aligned} M_0 \mathrm {e}^{\lambda _1 t} \ge \int _\Omega u(\cdot , t) = \omega _{n-1} \int \limits _0^R \rho ^{n-1} u(\rho , t) \,\mathrm {d}\rho \ge \omega _{n-1} \int \limits _0^r \rho ^{n-1} u(r, t) \,\mathrm {d}\rho = \frac{\omega _{n-1}}{n} r^n u(r, t) \end{aligned}$$

for all \((r, t) \in (0, R) \times (0, T_{\max })\). \(\square \)

Arguments based on the comparison principle as used in Lemma 5.1 are apparently not expedient for the less simplified system (1.3). Therefore, we rely on the pointwise upper bounds gained in [9] instead and argue similar as in [49, Lemma 3.3], whose proof instead of [9] relies on its predecessor [51] dealing with linear diffusion only.

Lemma 5.2

Let \(m \in [1, 2 - \frac{2}{n})\), \(\kappa \ge 0\), \(T, L, M_0 > 0\) and suppose that \(\lambda , \mu \) satisfy (1.4) and (1.5). For any \(\varepsilon > 0\), there is \(C > 0\) such that with \(p {:}{=}\frac{n(n-1)}{(m-1)n + 1} + \varepsilon \), the following holds: Whenever \(u_0 \in {C^{0}(\overline{\Omega })}\) complies with (1.8),

$$\begin{aligned} \int _\Omega u_0 \le M_0 \quad \text {as well as} \quad u_0(x) \le L |x|^{-p} \hbox { for all}\ x \in \Omega \end{aligned}$$

and \((u, v)\in \left( C^0({\overline{\Omega }}\times [0,T))\cap C^{2,1}({\overline{\Omega }}\times (0,T))\right) ^2\) is a classical solution of (PE), then

$$\begin{aligned} u(x, t) \le C \cdot |x|^{-p} \qquad \text {for all }x \in \Omega \hbox { and }t \in (0, T). \end{aligned}$$

Proof

Aiming to apply the pointwise upper bounds for quite general parabolic equations in divergence form obtained in [9], we first derive estimates for the second equation in (PE). Thus, writing the second equation in (PE) in radial coordinates, we see that

$$\begin{aligned} r^{n-1} v_r(r, t)&= \int \limits _0^r (\rho ^{n-1} v_r(\rho , t))_{\rho } \,\mathrm {d}\rho = \int \limits _0^r \rho ^{n-1} (v(\rho , t) - u(\rho , t)) \,\mathrm {d}\rho \le \frac{1}{\omega _{n-1}} \left( \int _\Omega v(\cdot , t) + \int _\Omega u(\cdot , t) \right) \end{aligned}$$
(5.4)

for all \((r, t) \in (0, R) \times (0, T)\). Since integrating the second equation in (PE) reveals that \(\int _\Omega v = \int _\Omega u\) in \((0, T_{\max })\), from (5.4) and Lemma 2.2, we infer that

$$\begin{aligned} r^{n-1} v_r(r, t)&\le \frac{2 \mathrm {e}^{\lambda _1 T} M_0}{\omega _{n-1}} {=}{:}c_1 \qquad \text {for all }(r, t) \in (0, R) \times (0, T). \end{aligned}$$

We now choose \(\theta > n\) so large that \(m-1 > \frac{1}{\theta }- \frac{1}{n}\) as well as \(p = \frac{n(n-1)}{(m-1)n + 1} + \varepsilon > \frac{n(n-1)}{(m-1)n + 1 - \frac{n}{\theta }}\). Then,

$$\begin{aligned} \int _\Omega |x|^{(n-1)\theta } |\nabla v(x, t)|^\theta \,\mathrm {d}x\le c_1^{\theta }|\Omega | {=}{:}c_2 \end{aligned}$$

, and since \({\tilde{u}}(x, t) {:}{=}\mathrm {e}^{-\lambda _1 t} u(x, t)\), \((x, t)\in {\overline{\Omega }}\times [0, T)\), solves

$$\begin{aligned} {\left\{ \begin{array}{ll} {\tilde{u}}_t \le \nabla \cdot ((\mathrm {e}^{\lambda _1 t} {\tilde{u}}+1)^{m-1} \nabla {\tilde{u}} - {\tilde{u}} \nabla v) &{} \text {in }\Omega \times (0, T), \\ ((\mathrm {e}^{\lambda _1 t} {\tilde{u}}+1)^{m-1} \nabla {\tilde{u}} - {\tilde{u}} \nabla v) \cdot \nu = 0 &{} \text {on }\partial \Omega \times (0, T), \\ {\tilde{u}}(\cdot , 0) = u_0 &{} \text {in }\Omega , \end{array}\right. } \end{aligned}$$

classically, an application of [9, Theorem 1.1] (with \(\alpha {:}{=}p\), \(q {:}{=}1\), \(K_{D, 1} {:}{=}1\), \(K_{D, 2} {:}{=}\max \{\mathrm {e}^{\lambda _1T(m-1)},1\}\), \(K_S {:}{=}1\), \(K_f {:}{=}c_2\), \(M {:}{=}M_0\), \(\beta {:}{=}n-1\), \(\mathbb {p} {:}{=}1\), \(D(x, t, {\tilde{u}}) {:}{=}(\mathrm {e}^{-\lambda _1 t} {\tilde{u}}+1)^{m-1}\) and \(S(x, t, {\tilde{u}}) {:}{=}{\tilde{u}}\)) yields \(c_3 > 0\) such that \({\tilde{u}}(x, t) \le c_3 |x|^{-p}\) and hence \(u(x, t) \le c_3 \mathrm {e}^{\lambda _1 T} |x|^{-p}\) for all \(x \in \Omega \) and \(t \in (0, T)\). \(\square \)

With these upper estimates at hand, we can now prove Theorem 1.2.

Proof of Theorem 1.2

Part (i) follows directly from Theorem 1.1 and Lemma 5.1.

Regarding part (ii), we first set \(p_0 {:}{=}\frac{n(n-1)}{(m-1)n + 1} \ge n\) and note that the assumption \(m \ge 1\) implies \(m > \frac{2}{n} \ge \frac{2}{p_0}\). Thus, (1.6)–(1.7) reduces to (1.6) and if \(\kappa \) satisfies (1.14) then also (1.6) for some \(p > p_0\) sufficiently close to \(p_0\).

For arbitrary \(L > 0\) and \(\varepsilon {:}{=}p - p_0 > 0\), we fix \(C > 0\) as given by Lemma 5.2. With \(r_1\) given by Theorem 1.1, we then choose \(u_0\) satisfying (1.8), having compact support in \(B_{r_1}(0)\) and fulfilling \(\int _\Omega u_0 = M_0\) as well as \(u_0(x) \le L |x|^{-p}\) for all \(x \in \Omega \). Note that this is indeed possible since \(p \ge n\) implies \(\int \limits _{B_{r_1}(0)} L|x|^{-p} \,\mathrm {d}x= \infty \). Finally, Theorem 1.1 asserts that the solution (uv) emanating from \(u_0\) blows up in finite time.