Abstract

We investigate a nonlocal reaction diffusion equation with absorption under Neumann boundary. We obtain optimal conditions on the exponents of the reaction and absorption terms for the existence of solutions blowing up in finite time, or for the global existence and boundedness of all solutions. For the blowup solutions, we also study the blowup rate estimates and the localization of blowup set. Moreover, we show some numerical experiments which illustrate our results.

1. Introduction

In this paper, we devote our attention to the singularity analysis of the following nonlocal diffusion equation: 𝑢𝑡(𝑥,𝑡)=Ω𝐽(𝑥𝑦)(𝑢(𝑦,𝑡)𝑢(𝑥,𝑡))𝑑𝑦+𝑢𝑝(𝑥,𝑡)𝑘𝑢𝑞(𝑢𝑥,𝑡),𝑥Ω,𝑡>0,(𝑥,0)=𝑢0(𝑥),𝑥Ω.(1.1) Here Ω is a bounded connected and smooth domain, which contains the origin, and 𝐽𝑁 is a nonnegative, bounded, symmetric radially and strictly decreasing function with 𝑁𝐽(𝑧)𝑑𝑧=1, and 𝑝, 𝑞, 𝑘 are all positive constants. We take the initial datum, 𝑢0(𝑥), nonnegative and nontrivial.

Equation in (1.1) is called a nonlocal diffusion equation in the sense that the diffusion of the density 𝑢 at a point 𝑥 and time 𝑡 does not only depend on 𝑢(𝑥,𝑡), but on all the values of 𝑢 in a neighborhood of 𝑥 through the convolution term. Maybe the simplest linear version of nonlocal model (1.1) is 𝑢𝑡(𝑥,𝑡)=𝑅𝑁𝐽(𝑥𝑦)(𝑢(𝑦,𝑡)𝑢(𝑥,𝑡))𝑑𝑦=𝐽𝑢(𝑥,𝑡)𝑢(𝑥,𝑡).(1.2) In recent years, the linear equation (1.2) and its variations have been widely used to model diffusion process, for example, in biology, dislocations dynamics, phase transition model, material science, and network model and so forth (see [17] and the references therein). The idea hidden inside those model is simple to understand. As stated in [6], if 𝑢(𝑥,𝑡) is thought of as a density at the point 𝑥 and time 𝑡 and 𝐽(𝑥𝑦) is thought of as the probability distribution of jumping from location 𝑦 to location 𝑥, then the convolution (𝐽𝑢)(𝑥,𝑡)=𝑁𝐽(𝑥𝑦)𝑢(𝑦,𝑡)𝑑𝑦 is the rate at which individuals are arriving at 𝑥 from all other places and 𝑢(𝑥,𝑡)=𝑁𝐽(𝑥𝑦)𝑢(𝑥,𝑡)𝑑𝑦 is the rate at which they are leaving location 𝑥 to travel to all other sites.

In the past decades, some works have shown that (1.2) shares many properties with the classical heat equation 𝑢𝑡Δ𝑢=0(1.3) such as the existence of constant bounded stationary solutions and the maximum principle, but there is no regularizing effect in general for the former [8]. Therefore, as mentioned in [911], it is an interesting topic to compare the properties of solutions to the nonlocal diffusion equation with that of the corresponding local diffusion cases.

To motivate our works, we would like to remark that, in recent years, (1.3) and its variations have been extensively studied. In particular, a considerable effort has been devoted to studying the blowup properties of the following classical diffusion equation with reaction (𝛼>0) and/or absorption (𝑘>0) under Dirichlet or Neumann boundary 𝑢𝑡Δ𝑢=𝛼𝑢𝑝𝑘𝑢𝑞,𝑥Ω,𝑡>0,(1.4) which provides a simple biological or physical model. For instance, by constructing the self-similar weak subsolutions, Bedjaoui and Souplet [12] obtained the following conclusion for (1.4) with 𝛼=1 under Dirichlet or Neumann boundary: if 𝑝<max{𝑞,1}, then all solutions are global. If 𝑝>max{𝑞,1}, there exist solutions of (1.4) which blow up in finite time. In the critical case 𝑝=max{𝑞,1}, the results may depend on the size of the coefficient 𝑘. In the Dirichlet boundary case, Xiang et al. [13] also studied the blowup rate estimates and obtained the following results: if 𝑝>max{𝑞,1} and the solution 𝑢(𝑥,𝑡) of (1.4) blows up at 𝑇, then there exists constants 𝐶>𝑐>0 such that max[]Ω×0,𝑡𝑢(𝑥,𝜏)𝑐(𝑇𝑡)1/(𝑝1),max[]Ω×0,𝑡𝑢(𝑥,𝜏)𝐶(𝑇𝑡)1/(𝑝1)2if1<𝑝<1+.𝑁+1(1.5)

In this paper, we will deal with the blowup properties of the nonlocal diffusion problem (1.1) and compare them with that of problem (1.4). In our model (1.1), the absorption term 𝑘𝑢𝑞(𝑥,𝑡) represents a rate of consumption due to an internal reaction, and we are integrating in Ω and thus imposing the condition that the diffusion takes place only in Ω, which means that the individual may enter or leave the domain. This is so called Neumann boundary conditions, see [14, 15]. We remark that García-Melián and Quirós [16] recently proved the existence of a critical exponent of Fujita type for the nonlocal diffusion problem 𝑢𝑡(𝑥,𝑡)=𝑅𝑁𝐽(𝑥𝑦)(𝑢(𝑦,𝑡)𝑢(𝑥,𝑡))𝑑𝑦+𝑢𝑝(𝑥,𝑡)=𝐽𝑢(𝑥,𝑡)𝑢(𝑥,𝑡)+𝑢𝑝(𝑥,𝑡).(1.6) As mentioned in [17], nonlocal diffusion systems are more accurate than the classical diffusion systems in modeling the spatial diffusion of the individuals in some biology areas, such as embryological development process. For more study about the nonlocal diffusion operator, we refer to [8, 1822] and references therein.

It is noteworthy that the method used in [12, 13] for problem (1.4) is invalid in our current setting due to the appearance of the nonlocal diffusion term. For example, instead of constructing self-similar weak subsolutions, we will use technique of integration to prove the finite time blowup. As far as the blowup rate is concerned, the scaling argument in [13] is not applicable.

We now state our main results. Our first result determines the complete classification of the parameters for which the solution blows up in finite time or exists globally.

Theorem 1.1. (i) If 𝑝max{𝑞,1}, then all solutions of (1.1) are global. Moreover, if 𝑝<𝑞 or 𝑝=𝑞 and 𝑘1, all solutions are uniformly bounded, while if 𝑝=𝑞 and 𝑘<1, there exist unbounded global solutions.
(ii) If 𝑝>max{𝑞,1}, then (1.1) with large initial data have solutions blowing up in finite time, while the solutions of (1.1) with small initial data exist globally.

Once we have obtained the values of the parameters for which blowup occurs, the next step is to concern the blowup rate, that is the speed at which solutions are blowing up. Different from the result of problem (1.4), we could have a unified upper and low estimate here.

Theorem 1.2. Let 𝑝>max{𝑞,1} and 𝑢(𝑥,𝑡) be a solution of (1.1) blowing up at time 𝑇. Then lim𝑡𝑇(𝑇𝑡)1/(𝑝1)max𝑥Ω𝑢(𝑥,𝑡)=(𝑝1)1/(𝑝1).(1.7)

Remark 1.3. From this result we find that the nonlocal diffusion term plays no role when determining the blowup rate and the blowup rate is just same as that of the ODE 𝑢𝑡=𝑢𝑝. And this phenomena is the same as that of local diffusion case.

Next we consider the spacial location of the blowup set. As usual, the blowup set of solution 𝑢(𝑥,𝑡) is defined as follows: 𝐵(𝑢)=𝑥𝑥Ω;thereexist𝑛,𝑡𝑛𝑥(𝑥,𝑇)suchthat𝑢𝑛,𝑡𝑛,(1.8) where 𝑇 is the maximal existence time of 𝑢. For a general domain Ω we can localize the blowup set near any pint in Ω just by taking an initial condition being very large near that point and not so large in the rest of the domain. This is the following result.

Theorem 1.4. Let 𝑝>max{𝑞,2}. For any 𝑥0Ω and 𝜀>0, there exists an initial data 𝑢0 such that the corresponding solution 𝑢(𝑥,𝑡) of (1.1) blows up at finite time 𝑇 and its blowup set 𝐵(𝑢) is contained in 𝐵𝜖(𝑥0)={𝑥Ω;𝑥𝑥0<𝜖}.

Next we consider the radially symmetric case. In this case, single point blowup occurs.

Theorem 1.5. Let 𝑝>max{𝑞,2} and Ω=𝐵𝑅={|𝑥|<𝑅}. If the initial data 𝑢0𝐶1(𝐵𝑅) is a radial nonnegative function with a unique maximum at the origin, that is, 𝑢0=𝑢0(𝑟)0, 𝑢0(𝑟)<0 for 0<𝑟𝑅, 𝑢(0)=0 and 𝑢0(0)<0, then the blowup set 𝐵(𝑢) of the solution 𝑢 of (1.1) consists only of the original point 𝑥=0.

The remainder of this paper is organized as follows. In Section 2, we give the existence and uniqueness of the solutions as well as the comparison principle. In Section 3, we prove the blowup and global existence condition. And then we prove the blowup rate and blowup set results in Sections 4 and 5, respectively. And in the last section we will give some numerical experiments to demonstrate our results.

2. Existence, Uniqueness, and Comparison Principle

We begin our study of problem (1.1) with a result of existence and uniqueness of continuous solutions and comparison principle.

Firstly, existence and uniqueness of solutions is a consequence of Banach's fixed point theorem. We look for 𝑢𝐶1((0,𝑇);𝐶(Ω))𝐶([0,𝑇);𝐶(Ω)) satisfying (1.1). Fix 𝑡0>0, and consider the Banach space 𝑋𝑡0=𝑢𝐶1((0,𝑇);𝐶(Ω))𝐶([0,𝑇);𝐶(Ω)) with the norm 𝜔𝑋𝑡0=max0𝑡𝑡0𝜔(,𝑡)𝐶(Ω)+max0<𝑡𝑡0𝜔𝑡(,𝑡)𝐶(Ω).(2.1)

We define the following operator 𝑇𝑋𝑡0𝑋𝑡0𝑇𝜔0(𝜔)(𝑥,𝑡)=𝜔0(𝑥)+𝑡0Ω+𝐽(𝑥𝑦)(𝜔(𝑦,𝑠)𝜔(𝑥,𝑠))𝑑𝑦𝑑𝑠𝑡0|𝜔|𝑝1𝜔(𝑥,𝑠)𝑘|𝜔|𝑞1𝜔(𝑥,𝑠)𝑑𝑠.(2.2)

Similar to [10, 15] we could prove the solution to (1.1) is a fixed point of operator 𝑇 in a convenient ball of 𝑋𝑡0. Thus, we could obtain the following result.

Theorem 2.1. For every 𝑢0𝐶(Ω) there exists a unique solution 𝑢 of (1.1) such that 𝑢𝐶1((0,𝑇);𝐶(Ω))𝐶([0,𝑇);𝐶(Ω)) and 𝑇 (finite or infinite) is the maximal existence time of solution.

Next, we will study the comparison principle. As usual we first give the definition of supersolution and subsolution.

Definition 2.2. A function 𝑢𝐶1((0,𝑇);𝐶(Ω))𝐶([0,𝑇);𝐶(Ω)) is a supersolution of (1.1) if it satisfies 𝑢𝑡(𝑥,𝑡)Ω𝐽(𝑥𝑦)𝑢(𝑦,𝑡)𝑢(𝑥,𝑡)𝑑𝑦+𝑢𝑝(𝑥,𝑡)𝑘𝑢𝑞(𝑥,𝑡),𝑢(𝑥,0)𝑢0(𝑥).(2.3) Subsolutions are defined similarly by reversing the inequalities.

To obtain the comparison principle for problem (1.1), we first give a maximum principle.

Lemma 2.3. Suppose that 𝑤(𝑥,𝑡)𝐶1((0,𝑇);𝐶(Ω))𝐶([0,𝑇);𝐶(Ω)) is nontrivial and satisfies𝑤𝑡(𝑥,𝑡)Ω𝐽(𝑥𝑦)(𝑤(𝑦,𝑡)𝑤(𝑥,𝑡))𝑑𝑦+𝑐1𝑤(𝑥,𝑡),𝑥Ω,𝑡>0(2.4) with 𝑤(𝑥,0)0 for 𝑥Ω, where 𝑐1 is a bounded function, then 𝑤(𝑥,𝑡)>0 for 𝑥Ω and 𝑡>0.

Proof. We first show 𝑤(𝑥,𝑡)0 for 𝑥Ω,𝑡>0. Assume that 𝑤(𝑥,𝑡) is negative somewhere. Let 𝜃(𝑥,𝑡)=𝑒𝜆𝑡𝑤(𝑥,𝑡)(𝜆>0,𝜆2sup|𝑐1|). If we take (𝑥0,𝑡0) a point where 𝜃 attains its negative minimum, there holds 𝑡0>0 and 𝜃𝑡𝑥0,𝑡0=𝜆𝑒𝜆𝑡0𝑤𝑥0,𝑡0+𝑒𝜆𝑡0𝑤𝑡𝑥0,𝑡0𝑒𝜆𝑡0Ω𝐽𝑥0𝑤𝑦𝑦,𝑡0𝑥𝑤0,𝑡0𝑑𝑦+𝜆+𝑐1𝑤𝑥0,𝑡0>0,(2.5) which is a contradiction. Thus 𝜃(𝑥,𝑡)0 for 𝑥Ω,𝑡>0. And so does 𝑤(𝑥,𝑡).
Now, we suppose 𝜃(𝑥1,𝑡1)=0 for some (𝑥1,𝑡1); that is, 𝜃 attains its minimum at (𝑥1,𝑡1) from the first step. Notice that the hypotheses on 𝐽 imply 𝐽(0)>0, so that 𝜃(𝑥1,𝑡1)=0 implies that 𝜃(𝑥,𝑡1)=0 for 𝑥 in a neighborhood of 𝑥1. Thus a standard connectedness argument yields 𝜃0. This is a contradiction. So we obtain our conclusion.

Lemma 2.4. If 𝑝1, 𝑞1 and 𝑢,  𝑢_are super and subsolutions to (1.1), respectively then 𝑢(𝑥,𝑡)𝑢_(𝑥,𝑡) for every (𝑥,𝑡)Ω×[0,𝑇).

Proof. Letting  𝑤(𝑥,𝑡)=𝑢𝑢_, it is easy to verify that 𝑤(𝑥,𝑡) satisfies (2.4) when 𝑝1,𝑞1. We could obtain our conclusion from Lemma 2.3.

Remark 2.5. When 𝑝<1 or 𝑞<1, the conclusion is also validity if 𝑢 and  𝑢_  are bounded away from 0.

3. Blowup and Global Existence

In this section, we will analyze the blowup condition and give the proof of Theorem 1.1.

Proof of Theorem 1.1. (i) We only need to look for a global supersolution of (1.1). Indeed, it is easy to construct spacial homogeneous global supersolution of (1.1). To see this, we set 𝑢=𝐶𝑒𝛼𝑡, where 𝐶 and 𝛼 are positive constants to be determined.
For any given initial data 𝑢0, we note that 𝑢(𝑡0)𝑢0 for 𝑡0 sufficiently large and 𝑢 is bounded away from 0. Thus by the comparison principle and Remark 2.5, to make 𝑢 a supersolution of (1.1), we only need to show the existence of 𝐶 and 𝛼 satisfying𝐶𝑝𝑒𝑝𝛼𝑡𝑘𝐶𝑞𝑒𝑞𝛼𝑡+𝛼𝐶𝑒𝛼𝑡.(3.1)If 𝑝1<𝑞, for any given 𝛼, we can take 𝐶=𝑘1/(𝑝𝑞) such that (3.1) holds. If 𝑞1 and thus 𝑝1, we can choose 𝐶 and 𝛼 satisfying 𝐶𝑝1=𝛼, which make (3.1) validity.
Next, we show all global solutions are uniformly bounded when 𝑝<𝑞 or 𝑝=𝑞 and 𝑘1. In fact, (1.1) has constant supersolution 𝑢=𝐴 whenever 𝑝<𝑞 or 𝑝=𝑞 and 𝑘1. To see this, we choose 𝐴 large enough such that 𝑘𝐴𝑞𝐴𝑝𝑢,𝐴0,(3.2) which imply that 𝑢 is a supersolution of (1.1).
At last we show there exist global unbounded solutions when 𝑝=𝑞 and 𝑘<1. Indeed, let (𝑓(𝑡)=1𝑘)(1𝑝)𝑡+𝑓1𝑝(0)1/(1𝑝)𝑒,𝑝=𝑞<1,((1𝑘)/2)𝑡,𝑝=𝑞=1.(3.3) It is easy to see that if 𝑓(0)maxΩ𝑢0(𝑥), 𝑓(𝑡) is a subsolution of (1.1). It is obvious that when 𝑝<1, 𝑓(𝑡) is unbounded.
(ii) We first show that if the initial data 𝑢0(𝑥) is large enough, solutions of (1.1) blow up in finite time.
In the case of 𝑝>𝑞>1. Integrating (1.1)1 in Ω and applying Fubini's theorem, we get 𝑑𝑑𝑡Ω𝑢(𝑥,𝑡)𝑑𝑥=Ω𝑢𝑝(𝑥,𝑡)𝑑𝑥𝑘Ω𝑢𝑞(𝑥,𝑡)𝑑𝑥.(3.4) Using Hölder's inequality, we could get 𝑑𝑑𝑡Ω𝑢(𝑥,𝑡)𝑑𝑥Ω𝑢𝑝||Ω||(𝑥,𝑡)𝑑𝑥𝑘(𝑝𝑞)/𝑝Ω𝑢𝑝(𝑥,𝑡)𝑑𝑥𝑞/𝑝=Ω𝑢𝑝(𝑥,𝑡)𝑑𝑥𝑞/𝑝Ω𝑢𝑝(𝑥,𝑡)𝑑𝑥(𝑝𝑞)/𝑝||Ω||𝑘(𝑝𝑞)/𝑝,(3.5) where |Ω| is assumed to be the measure of Ω. Given positive constant 𝑚>𝑘1/(𝑝𝑞) and 𝑢0𝑚, we have by the comparison principle that the solution 𝑢(𝑥,𝑡) of problem (1.1) satisfies 𝑢(𝑥,𝑡)𝑚. Thus 𝑑𝑑𝑡Ω𝑢(𝑥,𝑡)𝑑𝑥Ω𝑢𝑝(𝑥,𝑡)𝑑𝑥𝑞/𝑝𝑚𝑝𝑞||Ω||(𝑝𝑞)/𝑝||Ω||𝑘(𝑝𝑞)/𝑝.(3.6) Then we use Jensen's inequality to obtain 𝑑𝑑𝑡Ω𝑢(𝑥,𝑡)𝑑𝑥>𝐶Ω𝑢(𝑥,𝑡)𝑑𝑥𝑞,(3.7) where 𝐶 is a positive constant independent of the solution 𝑢. From this inequality, we could easily obtain that 𝑢(𝑥,𝑡) blow up in finite time.
In the case of 𝑝>1𝑞, it follows from 𝑢𝑞𝑢+1 and Jensen's inequality that Ω𝑢𝑝(𝑥,𝑡)𝑑𝑥𝑘Ω𝑢𝑞(𝑥,𝑡)𝑑𝑥Ω𝑢𝑝(𝑥,𝑡)𝑑𝑥𝑘Ω||Ω||||Ω||𝑢(𝑥,𝑡)𝑑𝑥𝑘1𝑝Ω𝑢(𝑥,𝑡)𝑑𝑥𝑝𝑘Ω𝑢||Ω||.(𝑥,𝑡)𝑑𝑥𝑘(3.8) Substituting this inequality into (3.4), we obtain 𝑑𝑑𝑡Ω||Ω||𝑢(𝑥,𝑡)𝑑𝑥1𝑝Ω𝑢(𝑥,𝑡)𝑑𝑥𝑝𝑘Ω||Ω||.𝑢(𝑥,𝑡)𝑑𝑥𝑘(3.9)
Therefore, if we take the initial data 𝑢0 large enough such that |Ω|1𝑝(Ω𝑢0(𝑥)𝑑𝑥)𝑝𝑘Ω𝑢0(𝑥)𝑑𝑥𝑘|Ω|>0, then Ω𝑢(𝑥,𝑡)𝑑𝑥 blows up in finite time. So does 𝑢(𝑥,𝑡).
Next we show when the initial data 𝑢0(𝑥) is small, solutions of (1.1) exist globally. Consider constant 𝐵. Let 0<𝐵𝑘1/(𝑝𝑞). Then 𝐵𝑡𝐵𝑝𝑘𝐵𝑞. Henceforth, if 𝑢0(𝑥)𝐵, 𝐵 is a supersolution of (1.1). From the comparison principle, we know solutions of (1.1) are global in this case.

4. Blowup Rate Estimate

In this section, we study the blowup rate and prove Theorem 1.2.

Proof of Theorem 1.2. Let 𝑈(𝑡)=𝑢(𝑥(𝑡),𝑡)=max𝑥Ω𝑢(𝑥,𝑡). It is easy to see that 𝑈(𝑡) is Lipschitz continuous and thus it is differential almost everywhere [23]. From the first equality of (1.1) we have 𝑈(𝑡)Ω𝐽(𝑥𝑦)(𝑢(𝑦,𝑡)𝑢(𝑥(𝑡),𝑡))𝑑𝑦+𝑢𝑝(𝑥(𝑡),𝑡)𝑘𝑢𝑞(𝑥(𝑡),𝑡)𝑢𝑝(𝑥(𝑡),𝑡)(4.1) at any point of differentiability of 𝑈(𝑡). Here we used 𝑢(𝑥(𝑡),𝑡)=0. Noticing that 𝑝>1 and integrating (4.1) from 𝑡 to 𝑇, we obtain max𝑥Ω𝑢(𝑥,𝑡)(𝑝1)1/(𝑝1)(𝑇𝑡)1/(𝑝1).(4.2)
Next we will establish the upper estimate. For any (𝑥,𝑡)Ω×[0,𝑇), we have 𝑢𝑡(𝑥,𝑡)𝑢(𝑥,𝑡)+𝑢𝑝(𝑥,𝑡)𝑘𝑢𝑞(𝑥,𝑡)=𝑢𝑝(𝑥,𝑡)1𝑢(𝑝1)(𝑥,𝑡)𝑘𝑢(𝑝𝑞).(𝑥,𝑡)(4.3) In particular, 𝑈(𝑡)𝑈𝑝(𝑡)1𝑈(𝑡)(𝑝1)𝑘𝑈(𝑡)(𝑝𝑞).(4.4)
From the lower estimate (4.2) we get 𝑈(𝑡)𝑈𝑝(𝑡)1(𝑝1)(𝑇𝑡)𝑘(𝑝1)(𝑝𝑞)/(𝑝1)(𝑇𝑡)(𝑝𝑞)/(𝑝1).(4.5)
Integrating in (𝑡,𝑇), we get max𝑥Ω𝑢(𝑥,𝑡)(𝑝1)(𝑇𝑡)(𝑝1)22(𝑇𝑡)2𝑘(𝑝1)(3𝑝𝑞2)/(𝑝1)2𝑝𝑞1(𝑇𝑡)(2𝑝𝑞1)/(𝑝1)1/(𝑝1),(4.6) combining with (4.2), the conclusion of Theorem 1.2 is proved if one takes the limit as 𝑡𝑇.

5. Blowup Set

Next we will concern the blowup set for the solution to problem (1.1). We will first localize the blowup set near any point in Ω just by taking an initial condition being very large near that point and not so large in the rest of the domain.

Proof of Theorem 1.4. Given 𝑥0Ω and 𝜀>0, we could construct an initial condition 𝑢0 such that 𝐵(𝑢)𝐵𝜀𝑥0=𝑥Ω𝑥𝑥0<𝜀.(5.1)
In fact, we will consider 𝑢0 concentrated near 𝑥0 and small away from 𝑥0.
Let 𝜑 be a nonnegative smooth function such that supp(𝜑)𝐵𝜀/2(𝑥0) and 𝜑(𝑥)>0 for 𝑥𝐵𝜀/2(𝑥0).
Next, let 𝑢0(𝑥)=𝑀𝜑(𝑥)+𝛿.(5.2) We want to choose 𝑀 large and 𝛿 small such that (5.1) holds.
First we can assume that 𝑇 is as small as we need by taking 𝑀 large enough. Indeed, we have 𝑇𝐶(Ω,𝑝,𝜑)𝑀𝑞1or𝑇𝐶(Ω,𝑝,𝜑)𝑀𝑝1(5.3) from the proof of Theorem 1.1.
Now, from the proof of blowup rate, we havemax𝑥Ω𝑢(𝑥,𝑡)(𝑝1)(𝑇𝑡)(𝑝1)22(𝑇𝑡)2𝑘(𝑝1)(3𝑝𝑞2)/(𝑝1)2𝑝𝑞1(𝑇𝑡)(2𝑝𝑞1)/(𝑝1)1/(𝑝1)𝐶(𝑇𝑡)1/(𝑝1).(5.4) Henceforth, for any 𝑥Ω, 𝑢𝑡=𝑥,𝑡Ω𝐽(𝑥,𝑦)𝑢(𝑦,𝑡)𝑢𝑥,𝑡𝑑𝑦+𝑢𝑝𝑥,𝑡𝑘𝑢𝑞𝑥,𝑡Ω𝐽𝑥,𝑦𝑢(𝑦,𝑡)𝑑𝑦+𝑢𝑝𝑥,𝑡𝐶(𝑇𝑡)1/(𝑝1)+𝑢𝑝.𝑥,𝑡(5.5) Therefore, if 𝑥Ω𝐵𝜖(𝑥0), then 𝑢(𝑥,𝑡) is a subsolution to 𝑤𝑡=𝐶(𝑇𝑡)1/(𝑝1)+𝑤𝑝𝑤(𝑡),(0)=𝛿,(5.6) which shows 𝑢𝑥,𝑡𝑤(𝑡).(5.7)
Next, we only need to prove that a solution 𝑤 to (5.6) remains bounded up to 𝑡=𝑇, provided that 𝛿 and 𝑇 are small enough.
Let 𝑧(𝑠)=(𝑇𝑡)1/(𝑝1)𝑤(𝑡),𝑠=ln(𝑇𝑡).(5.8) Then 𝑧(𝑠) satisfies 𝑧(𝑠)=𝐶𝑒𝑠1𝑝1𝑧(𝑠)+𝑧𝑝(𝑠),𝑧(ln𝑇)=𝛿𝑇1/(𝑝1),(5.9) which show that for 𝑇 and 𝛿 small (𝑇 is small if 𝑀 is large), we have 1𝐶𝑇𝑝1𝛿𝑇1/(𝑝1)+𝛿𝑝𝑇𝑝/(𝑝1)<0.(5.10) So 𝑧(𝑠)<0 for all 𝑠>ln𝑇. From this and Lemma 4.2 of [24], we know 𝑧(𝑠)0,𝑠.(5.11) Combining the equation verified by 𝑧 we obtain that, for given positive constant 𝛾(<1/𝑝(𝑝1)), there exists 𝑠0>0 such that 𝑧(𝑠)𝐶𝑒𝑠1𝑝1𝛾𝑧(𝑠)(5.12) for 𝑠>𝑠0.
Let 𝑣(𝑠) be a solution of 𝑣(𝑠)=𝐶𝑒𝑠1𝑝1𝛾𝑣(𝑠)(5.13) with 𝑣(𝑠0)𝑧(𝑠0). Integrating this equation we get 𝑣(𝑠)=𝐶1𝑒𝑠+𝐶2𝑒(1/(𝑝1)𝛾)𝑠.(5.14) By a comparison argument we could get that for every 𝑠>𝑠0, 𝑧(𝑠)𝑣(𝑠)=𝐶1𝑒𝑠+𝐶2𝑒(1/(𝑝1)𝛾)𝑠.(5.15) Now we go back to 𝑧(𝑠)=𝐶𝑒𝑠(1/(𝑝1))𝑧(𝑠)+𝑧𝑝(𝑠). We have 𝑧(𝑠)+(1/(𝑝1))𝑧(𝑠)=𝐶𝑒𝑠+𝑧𝑝(𝑠),(5.16) then 𝑒(1/(𝑝1))𝑠𝑧(𝑠)=𝑒(1/(𝑝1))𝑠(𝐶𝑒𝑠+𝑧𝑝).(5.17) Integrating form 𝑠0 to 𝑠, one could get 𝑧(𝑠)=𝑒(1/(𝑝1))𝑠𝐶1+𝑠𝑠0𝑒(1/(𝑝1))𝜎(𝐶𝑒𝜎+𝑧𝑝(𝑠))𝑑𝜎=𝑒(1/(𝑝1))𝑠𝐶1+𝑠𝑠0𝑒((𝑝2)/(𝑝1))𝜎(𝐶+𝑒𝜎𝑧𝑝.(𝑠))𝑑𝜎(5.18) Using (5.15) and 𝛾<1/𝑝(𝑝1), we have 𝑒𝑠𝑧𝑝𝐶𝑝1𝑒(𝑝1)𝑠+𝐶𝑝2𝑒(𝑝/(𝑝1)𝑝𝛾1)𝑠0(5.19) as 𝑠+.
And thus from (5.18), we get 𝑧(𝑠)𝑒(1/(𝑝1))𝑠𝐶1+𝐶3𝑠𝑠0𝑒((𝑝2)/(𝑝1))𝜎𝑑𝜎𝐶1𝑒(1/(𝑝1))𝑠+𝐶4𝑒𝑠.(5.20)
As 𝑝>2, we have 𝑧(𝑠)𝐶𝑒(1/(𝑝1))𝑠.(5.21) This implies that 𝑤(𝑡)𝐶, for 0𝑡<𝑇. From the boundedness of 𝑤 and (5.7) we get 𝑢(𝑥,𝑡)𝑤(𝑡)𝐶 for every 𝑥Ω𝐵𝜖(𝑥0), as we wished.

Next, we will consider the radial symmetric case, that is, the proof of Theorem 1.5. For the convenience of writing, we only deal with the one dimensional case, Ω=(𝑙,𝑙). The radial case is analogous; we leave the details to the reader.

Proof of Theorem 1.5. Under the hypothesis on the initial condition imposed in Theorem 1.5 we have that the solution is symmetric and 𝑢𝑥<0 in (0,𝑙]×(0,𝑇) from the stand parabolic theorem and Lemma 4.1 of [10]. Therefore the solution has a unique maximum at the origin for every 𝑡(0,𝑇).
Let us perform the following change of variables 𝑧(𝑥,𝑠)=(𝑇𝑡)1/(𝑝1)𝑢(𝑥,𝑡),𝑠=ln(𝑇𝑡).(5.22) Our remainder proof consist of two steps.
Step 1. We first prove the only blowup point that verifies the blowup estimate (1.7) is 𝑥=0. And this shows that for 𝑥0, 𝑧(𝑥,𝑠) does not converge to 𝐶𝑝=(𝑝1)1/(𝑝1) as 𝑠+.
We conclude by contradiction. Assume that (𝑇𝑡)1/(𝑝1)𝑢(𝑥0,𝑡)𝐶𝑝 for a 𝑥0>0.
Let 𝑣(𝑡)=𝑢(0,𝑡)𝑢(𝑥0,𝑡). Then 𝑣(𝑡)=𝑙𝑙𝐽(𝑦)(𝑢(𝑦,𝑡)𝑢(0,𝑡))𝑑𝑦𝑙𝑙𝐽𝑥0𝑥𝑦𝑢(𝑦,𝑡)𝑢0,𝑡𝑑𝑦+𝑝𝜉𝑝1(𝑡)𝑣(𝑡)𝑘𝑞𝜂𝑞1(𝑡)𝑣(𝑡),(5.23) where 𝜉(𝑡) and 𝜂(𝑡) are between 𝑢(0,𝑡) and 𝑢(𝑥0,𝑡). Hence 𝑣(𝑡)𝑙𝑙𝑥𝐽(𝑦)𝐽0+𝑦𝑢(𝑦,𝑡)𝑑𝑦𝑙𝑙𝐽𝑦𝑥0𝐽(𝑦)𝑢(0,𝑡)𝑑𝑦𝑣(𝑡)+𝑝𝜉𝑝1(𝑡)𝑣(𝑡)𝑘𝑞𝜂𝑞1(=𝑡)𝑣(𝑡)𝑙𝑙𝐽𝑦𝑥0𝐽(𝑦)(𝑢(0,𝑡)𝑢(𝑦,𝑡))𝑑𝑦𝑣(𝑡)+𝑝𝜉𝑝1(𝑡)𝑣(𝑡)𝑘𝑞𝜂𝑞1(𝑡)𝑣(𝑡)𝐶1+𝑝𝜉𝑝1(𝑡)𝑘𝑞𝜂𝑞1(𝑡)𝑣(𝑡),(5.24) for some positive constant.
Integrating the above inequality, we obtain 𝑡ln(𝑣)(𝑡)ln(𝑣)0𝑡𝑡0𝐶1+𝑝𝜉𝑝1(𝑠)𝑘𝑞𝜂𝑞1(𝑠)𝑑𝑠.(5.25) Remember that (𝑇𝑡)1/(𝑝1)𝑢(𝑥0,𝑡)𝐶𝑝, (𝑇𝑡)1/(𝑝1)𝑢(0,𝑡)𝐶𝑝, we have lim𝑡𝑇𝜉(𝑡)(𝑇𝑡)1/(𝑝1)=lim𝑡𝑇𝜂(𝑡)(𝑇𝑡)1/(𝑝1)=𝐶𝑝.(5.26) And this implies that 𝑡𝑡0𝐶1+𝑝𝜉𝑝1(𝑠)𝑘𝑞𝜂𝑞1(𝑠)𝑑𝑠𝑝𝑡𝑡0𝐶𝑝𝑝1𝛿1𝑇𝑠𝑑𝑠𝑘𝑞𝑡𝑡0𝐶𝑝𝑞1+𝛿2(𝑇𝑠)(𝑞1)/(𝑝1)𝑑𝑠𝐶2.(5.27)𝑝>𝑞 implies that (𝑇𝑠)(𝑞1)/(𝑝1)𝛿3(𝑇𝑠)1 as 𝑠𝑇 for given 𝛿3>0. Hence 𝑡𝑡0𝐶1+𝑝𝜉𝑝1(𝑠)𝑘𝑞𝜂𝑞1(𝑠)𝑑𝑠𝑝𝑡𝑡0𝐶𝑝𝑝1𝛿𝑇𝑠𝑑𝑠𝐶2𝐶=𝑝𝑝𝑝1𝛿ln(𝑇𝑡)𝐶2(5.28) for some 𝛿>0.
Hence 𝑣(𝑡)𝐶(𝑇𝑡)𝑝(𝐶𝑝𝑝1𝛿)=𝐶(𝑇𝑡)𝑝𝛿𝑝/(𝑝1).(5.29) Using this fact, we have 0=lim𝑡𝑇(𝑇𝑡)1/(𝑝1)𝑣(𝑡)𝐶lim𝑡𝑇(𝑇𝑡)1/(𝑝1)𝑝/(𝑝1)+𝑝𝛿=+.(5.30) This contradiction proves our claim.
Step 2. We will show the only possible blowup point is 𝑥=0.
Remembering the transform (5.22), 𝑧(𝑥,𝑠) satisfies 𝑧𝑠=𝑒𝑠𝑙𝑙1𝐽(𝑥𝑦)(𝑧(𝑦,𝑠)𝑧(𝑥,𝑠))𝑑𝑦𝑝1𝑧+𝑧𝑝𝑘𝑒((𝑞𝑝)/(𝑝1))𝑠𝑧𝑞.(5.31) Note that the blowup rate of 𝑢 implies that 𝑧(𝑥,𝑠)𝐶 for every (𝑥,𝑠)[𝑙,𝑙]×(ln𝑇,). Therefore, 𝑧𝑠(𝑥,𝑠)𝐶𝑒𝑠1𝑝1𝑧(𝑥,𝑠)+𝑧𝑝(𝑥,𝑠).(5.32) From this we know that if there exists 𝑠0 such that 𝑧𝑝(𝑥,𝑠0)(1/(𝑝1))𝑧(𝑥,𝑠0)<𝐶𝑒𝑠0, then 𝑧(𝑥,𝑠)0 as 𝑠 (see Lemma 4.2 in [24]).
Moreover, if there exists 𝑠0 such that 𝑧𝑝(𝑥,𝑠0)(1/(𝑝1))𝑧(𝑥,𝑠0)>𝐶𝑒𝑠0 then 𝑧(𝑥,𝑠) blows up in finite time 𝑠. This follows from Lemma 4.3 of [24] using that 𝑧𝑠(𝑥,𝑠)𝐶𝑒𝑠1𝑝1𝑧(𝑥,𝑠)+𝑧𝑝(𝑥,𝑠).(5.33)
Thus if 𝑧(𝑥,𝑠) does not converge to zero and does not blow up in finite time, then 𝑧(𝑥,𝑠) satisfies 𝐶𝑒𝑠𝑧𝑝1(𝑥,𝑠)𝑝1𝑧(𝑥,𝑠)𝐶𝑒𝑠.(5.34) Henceforth, 𝑧𝑝1(𝑥,𝑠)𝑝1𝑧(𝑥,𝑠)0(𝑠+).(5.35) As 𝑧(𝑥,𝑠) is continuous, bounded and does not go to zero, we conclude that 𝑧(𝑥,𝑠)𝐶𝑝.
Now we could conclude that 𝑧(𝑥,𝑠) verifies 𝑧(𝑥,𝑠)0(𝑠+), or 𝑧(𝑥,𝑠)𝐶𝑝(𝑠+), or 𝑧(𝑥,𝑠) blows up in finite time.
From Step 1 we know for 𝑥0, 𝑧(𝑥,𝑠) is bounded and does not converge to 𝐶𝑝, so 𝑧(𝑥,𝑠)0 as 𝑠+. Combined with inequality (5.32), we could get 𝑧𝑠(𝑥,𝑠)𝐶𝑒𝑠1𝑝1𝜃𝑧(𝑥,𝑠)(5.36) for any 𝜃>0.
By a comparison argument as in the proof of Theorem 1.4, it follows that 𝑧(𝑥,𝑠)𝐶1𝑒𝑠+𝐶2𝑒(1/(𝑝1)𝜃)𝑠.(5.37) Going back to the equation verified by 𝑧(𝑥,𝑡) we obtain 𝑒(1/(𝑝1))𝑠𝑧(𝑥,𝑠)𝑠=𝑒(1/(𝑝1))𝑠𝑒𝑠𝑙𝑙𝐽(𝑥𝑦)(𝑧(𝑦,𝑠)𝑧(𝑥,𝑠))𝑑𝑦+𝑧𝑝(𝑥,𝑠)𝑘𝑒((𝑞𝑝)/(𝑝1))𝑠.𝑧(𝑥,𝑠)(5.38) Integrating we get 𝑧(𝑥,𝑠)=𝑒(1/(𝑝1))𝑠𝐶1+𝑠𝑠0𝑒((𝑝2)/(𝑝1))𝜎𝑙𝑙𝐽(𝑥𝑦)(𝑧(𝑦,𝑠)𝑧(𝑥,𝑠))𝑑𝑦+𝑒𝜎𝑧𝑝(𝑥,𝑠)𝑘𝑒((𝑞1)/(𝑝1))𝜎.𝑧(𝑥,𝑠)𝑑𝜎(5.39) On the other hand, (5.37) implies that 𝑒𝑠𝑧𝑝(𝑥,𝑠)0 as 𝑠. Henceforth, 𝑧(𝑥,𝑠)𝑒(1/(𝑝1))𝑠𝐶1+𝐶2𝑠𝑠0𝑒((𝑝2)/(𝑝1))𝜎.𝑑𝜎(5.40) Using that 𝑝>2, one could have 𝑧(𝑥,𝑠)𝐶3𝑒(1/(𝑝1))𝑠.(5.41) Remembering the transform (5.22), we have 𝑢(𝑥,𝑡)=𝑒(1/(𝑝1))𝑠𝑧(𝑥,𝑠)𝑐3.(5.42) And so our proof is complete.

6. Numerical Experiments

At the end of this paper, we will use several numerical examples to demonstrate our results about the location of blowup points. For this purpose, we discretize the problem in the spacial variable to obtain an ODE system. Taking Ω=[4,4] and 4=𝑥𝑁<<𝑥𝑁=4,𝑁=100, we consider the following system: 𝑢𝑖(𝑡)=𝑁𝑗=𝑁𝐽𝑥𝑖𝑥𝑗𝑢𝑗(𝑡)𝑢𝑖+𝑢(𝑡)𝑖𝑝𝑢(𝑡)𝑘𝑖𝑞𝑢(𝑡),𝑖(0)=𝑢0𝑥𝑖.(6.1) Next we choose 𝑝=3, 𝑞=1, 𝑘=1 and 1𝐽(𝑧)=1,|𝑧|,1100,|𝑧|>.10(6.2)

In Figure 1 we choose a nonsymmetric initial condition very large near the point 𝑥0=1,  𝑢0(𝑥)=1/4+100(1|𝑥1|)+. We observe that the blowup set is localized in a neighborhood of 𝑥0=1.

Next we choose a symmetric initial condition with a unique maximum at the origin, 𝑢0(𝑥)=16𝑥20. We observe that the solution blows up only at the origin, Figure 2.

Acknowledgments

Y. Wang is supported by the Key Scientific Research Foundation of Xihua University (no.  Z0912611), the Scientific Research Found of Sichuan Provincial Education Department (no. 09ZB081) and the Research Fund of Key Disciplinary of Application Mathematics of Xihua University (no. XZD0910-09-1). Z. Xiang is supported by NNSF of China (11101068), the Sichuan Youth Science & Technology Foundation (2011JQ0003), the Fundamental Research Funds for the Central Universities (ZYGX2009X019) and the SRF for ROCS, SEM. J. Hu is supported by the Scientific Research Found of Sichuan Provincial Education Department (no. 11ZB009).