Skip to content
BY 4.0 license Open Access Published by De Gruyter May 12, 2020

On a Class of Reaction-Diffusion Equations with Aggregation

  • Li Chen , Laurent Desvillettes and Evangelos Latos EMAIL logo

Abstract

In this paper, global-in-time existence and blow-up results are shown for a reaction-diffusion equation appearing in the theory of aggregation phenomena (including chemotaxis). Properties of the corresponding steady-state problem are also presented. Moreover, the stability around constant equilibria and the non-existence of nonconstant solutions are studied in certain cases.

1 Introduction

We consider the following initial boundary value problem:

(1.1) { t u = Δ [ ( a - b u ) u ] + ( c - d u ) u in  Ω × ( 0 , T ) , [ u ] = 0 on  Ω × ( 0 , T ) , u ( x , 0 ) = u 0 ( x ) 0 ,

where a,b>0 and c,d. Here, Ω is an open bounded domain in n, and [u] denotes a boundary operator of Neumann or Dirichlet type, i.e.

u | Ω = 0 , or ( a - 2 b u ) u γ | Ω = 0 ,

where γ is the outer unit normal vector of Ω. For the sake of simplicity, we take |Ω|=1.

One of the motivations to study such an equation comes from the structure similarities that exist with the parabolic-elliptic Keller–Segel models for chemotaxis, i.e.

t u - ( u - u V * u ) = 0 ,

where V is the fundamental solution of Poisson equation (or some other given potential in the case of general diffusion aggregation equations). If formally the interaction potential V is replaced by a the Dirac mass δ0, then the above equation is reduced to

t u - ( u - u u ) = 0 .

In [2], the authors propose a microscopic particle model which converges at the formal level towards such a PDE. This microscopic particle model, corresponding to (1.1) with c=d=0 and Ω=n, is the following:

d X t i = 2 a d B t i + 1 N j i V ε ( | X t i - X t j | ) d t ,
X 0 i = ξ i i.i.d. random variables with distribution  u 0 ,

where Vε(x)=ε-nV(xε) for ε>0, nV(x)𝑑x=2b, and Bti are i.i.d. Brownian motions.

It is well known (for example in [11, 9] that under suitable assumptions on V, the particle model converges (when N goes to infinity) to the following intermediate (with fixed ε>0) nonlocal problem:

d X ¯ t i = 2 a d B t i + n V ε ( | X ¯ t i - y | ) u ε ( y , t ) 𝑑 y 𝑑 t ,
X ¯ 0 i = ξ i  i.i.d. random variables with distribution  u 0 ,

where uε is the distribution of the i.i.d. random processes X¯ti at time t. By Itô’s formula one can obtain the following nonlocal partial differential equation for uε:

(1.2) t u ε - ( a u ε - u ε V ε * u ε ) = 0 .

Furthermore, in the parabolic regime, i.e. when 0u0<a2b, it is proved in [2] that the limit of uε satisfies equation (1.1) with c=d=0:

(1.3) t u - Δ ( a u - b u 2 ) = 0 .

The physical meaning of the unknown u is that of a concentration, therefore one considers only nonnegative solutions corresponding to nonnegative initial data.

Furthermore (like in the case of Keller–Segel system), problem (1.1) with c=d=0 (and the homogeneous Neumann boundary condition) possesses the following entropy structure:

d d t ( t ) := d d t Ω ( a u ( log u - 1 ) - b u 2 ) 𝑑 x = - Ω 1 u ( a - 2 b u ) 2 | u | 2 𝑑 x 0 .

This entropy is a combination of a positive part from the diffusion and a negative one from the aggregation. It needs to be pointed out that here the aggregation phenomenon is much stronger than the one appearing in Keller–Segel systems because of the singular potential that appeared in (1.2)–(1.3).

As for the reaction term, it is considered to be of logistic (mono-stable) type so that (when d>0) a significant dampening effect is exercised on the density u at those points where u becomes large.

The arrangement of the paper is the following. In Section 2, global existence and uniqueness of classical solutions are obtained for initial data such that parabolicity is expected to hold. The rest of the paper concerns cases in which parabolicity is expected to be lost at some point, so that blowup may happen. Considerations on the possible steady states and their stability as well as direct estimates of blowup are presented. Section 3 is devoted to the study of the steady states. The non-existence of nontrivial steady states is proved via Pohozaev’s type arguments. Furthermore, the linear stability of constant steady states is investigated. Finally, in Section 4, blow-up (in finite time) results are presented. Two different procedures are carried out: Kaplan’s method is used for the problem with Dirichlet boundary condition on one hand, and the concavity method is used for the problem with Neumann boundary condition on the other hand. In the end, we present an annex where blowup is directly observed in a class of explicit solutions linked to Barenblatt profiles, and we draw conclusions in a final section.

2 Global Existence

In this section, the global existence and uniqueness of a solution is obtained thanks to Leray–Schauder fixed point theorem, under the condition that the initial datum belongs to the parabolic region.

Observing that

Δ [ ( a - b u ) u ] = - b Δ [ ( u - a 2 b ) 2 ] ,

and using the notation v:=u-a2b, (1.1) can be rewritten as

(2.1) { t v = - b Δ ( v 2 ) + ( c - d a 2 b - d v ) ( v + a 2 b ) in  Ω × ( 0 , T ) , [ v + a 2 b ] = 0 on  Ω × ( 0 , T ) , v ( x , 0 ) = v 0 ( x ) .

It can be expected that global existence holds in the case when the parabolicity can be kept in the evolution (that is, when for all time v<0, or equivalently u<a2b). At the same time, the logistic term u(c-du) and the expected nonnegativity of u imply that the estimate 0ucd should also hold. Therefore, a natural sufficient condition for getting global existence for equation (1.1) is 0u0<a2b, together with cd<a2b. The theorem below states a precise result in this direction:

Theorem 1 (Global Existence and Uniqueness).

Let Ω be a smooth bounded domain in Rn. Assume a,b,d>0 and a2b>cd. Let also 0u0Cα(Ω¯), α(0,1) and

max x Ω ¯ u 0 ( x ) < a 2 b ,

with compatibility condition B[u0]=0. Then problem (1.1), together with homogeneous Neumann or homogeneous Dirichlet boundary condition B[u]=0 has a unique global-in-time classical solution. In addition, it holds that

0 u ( x , t ) < a 2 b for all  x Ω ¯ , t 0 .

Proof.

We first observe that we can take ε0>0 small enough in such a way that maxxΩ¯u0(x)a2b-ε0 and cda2b-ε0. Then we will prove the existence and uniqueness of a solution u of the problem, which satisfies the estimate

0 u ( x , t ) a 2 b - ε 0 for all  x Ω ¯ , t 0 .

For any fixed T>0, we will use the Leray–Schauder fixed point theorem to prove the existence. Let

X = { w C α , α 2 ( Ω ¯ × [ 0 , T ] ) : 0 w ( x , t ) a 2 b - ε 0  for all  x Ω ¯ , t 0 } .

We define an operator as follows: for given wX and σ[0,1], let u:=𝒯(w,σ) be the C2+α,1+α2(Ω¯×[0,T]) solution (see [8, Chapter V, Theorem 7.4] for the existence and uniqueness of the solution) of the following problem:

{ t u - ( 1 - σ ) Δ u - σ ( a - 2 b w ) Δ u + 2 σ b | u | 2 = σ u ( c - d u ) in  Ω × ( 0 , T ) , [ u ] = 0 on  Ω × ( 0 , T ) , u ( x , 0 ) = σ u 0 ( x ) .

In order to build up the map 𝒯, we have to show that 0u(x,t)a2b-ε0 for all (x,t)Ω¯×[0,T].

For σ=0, it is obvious that 0ua2b-ε0, since in that case u satisfies the heat equation.

For σ(0,1], we first prove that u0. To this end, let ε>0 be small and let uε be the solution of

{ t u ε - ( 1 - σ ) Δ u ε - σ ( a - 2 b w ) Δ u ε + 2 σ b | u ε | 2 = σ u ε ( c - d u ε ) + ε in  Ω × ( 0 , T ) , [ u ε ] = 0 on  Ω × ( 0 , T ) , u ε ( x , 0 ) = σ u 0 ( x ) .

The solution uεC2+α,1+α2(Ω¯×[0,T]) possesses a uniform in ε estimate uε2+α,1+α2C, see [8, Chapter V, Theorem 7.4]. With u00, if minΩ¯×[0,T]uε(x,t)<0, then there exists (x1,t1)Ω¯×(0,T] such that

(2.2) 0 = u ε ( x 1 , t 1 ) = min Ω ¯ u ε ( x , t 1 ) with  t u ε ( x 1 , t 1 ) 0 .

More precisely, we take here t10 as the last time before the solution takes some negative value. If x1Ω, we have uε(x1,t1)=0. If x1Ω, in the case of the homogeneous Neumann boundary condition, we also have uε(x1,t1)=0. Then we get (still in the case of Neumann boundary condition)

0 ( t u ε - ( 1 - σ ) Δ u ε - σ ( a - 2 b w ) Δ u ε + 2 σ b | u ε | 2 ) | ( x 1 , t 1 ) = σ u ε ( c - d u ε ) | ( x 1 , t 1 ) + ε > 0 ,

which is a contradiction. Therefore, uε0.

If x1Ω (and uε(x,t1)>0 for all xΩ), in the case of the homogeneous Dirichlet boundary condition, we can prove (see the sequel of the proof) that there exists a sequence xnΩ satisfying xnx1 and such that

(2.3) lim n u ε ( x n , t 1 ) - u ε ( x 1 , t 1 ) | x n - x 1 | = lim n u ε ( x n , t 1 ) | x n - x 1 | = 0 .

This limit, together with the fact that uε is smooth and that the tangential derivative of uε vanishes because of the homogeneous Dirichlet boundary condition, shows that uε|(x1,t1)=0. Thus we can follow the same argument as above.

In order to show the limit (2.3), we consider tn=t1+1n a sequence that converges to t1 and xn (n2) one of the minimal points of uε(x,tn) such that xnx1 (note that if several points x1Ω satisfy (2.2), one at least can be selected in such a way that the construction above makes sense). Then uε(xn,tn)uε(x1,t1) because of the continuity of uε. We get therefore

u ε ( x n , t 1 ) > 0 and u ε ( x n , t n ) < 0 .

Then the mean value theorem implies that there exists a sequence tn*(t1,tn) such that uε(xn,tn*)=0. Therefore,

lim n u ε ( x n , t 1 ) - u ε ( x 1 , t 1 ) | x n - x 1 | = lim n u ε ( x n , t 1 ) - u ε ( x n , t n * ) | x n - x 1 |
= lim n 1 | x n - x 1 | u ε t ( x n , t n * * ) ( t 1 - t n * )
= lim n 1 | x n - x 1 | ( u ε t ( x n , t n * * ) - u ε t ( x 1 , t n * * ) ) ( t 1 - t n * ) ,

where tn**(t1,tn*) and the last line comes again from the homogeneous Dirichlet boundary condition. As a consequence, since we are working with bounded classical solutions, the limit vanishes.

On the other hand, ρε=uε-u satisfies the following linear problem:

{ t ρ ε - ( 1 - σ ) Δ ρ ε - σ ( a - 2 b w ) Δ ρ ε + 2 σ b ( u ε + u ) ρ ε - σ c ρ ε + σ d ( u ε + u ) ρ ε = ε in  Ω × ( 0 , T ) , [ ρ ε ] = 0 on  Ω × ( 0 , T ) , ρ ε ( x , 0 ) = 0 ,

where all the coefficients are uniformly bounded in ε. Therefore by the maximum principle, we have

u ε - u L ( Ω × [ 0 , T ] ) = ρ ε L ( Ω × [ 0 , T ] ) C T ε .

By taking the limit ε0 in uε0, we obtain that u0 in Ω×[0,T].

Next we prove that ua2b-ε0 in Ω×[0,T]. Suppose that there exists (x0,t0)Ω¯×(0,T] such that

a 2 b - ε 0 < u ( x 0 , t 0 ) = max Ω ¯ × [ 0 , T ] u ( x , t ) .

Then we have tu(x0,t0)0, and moreover x0Ω if we consider the Dirichlet boundary condition, so that

0 ( t u - ( 1 - σ ) Δ u - σ ( a - 2 b w ) Δ u + 2 σ b | u | 2 ) | ( x 0 , t 0 ) = σ u ( c - d u ) | ( x 0 , t 0 ) ,

which means

c - d u ( x 0 , t 0 ) 0 u ( x 0 , t 0 ) c d .

In the case of the Neumann boundary conditions, x0 might appear on the boundary, but in this case, we still have u(x0,t0)=0, therefore the above argument also works. This implies a2b-ε0<cd, which is a contradiction with the assumption a2b-ε0cd. Therefore maxΩ¯×[0,T]u(x,t)a2b-ε0.

Thus the map 𝒯:X×[0,1]X is well defined. Due to the compact embedding from C2+α,1+α2 to Cα,α2, we know that 𝒯(,σ):XX is a compact operator.

Next we show that the map 𝒯 is continuous in w and σ. For all wX and σ[0,1], let wjX be a sequence such that wj-wCα,α20 as j, and let σj[0,1] be a sequence such that |σj-σ|0. Let uj=𝒯(wj,σj), the Schauder estimates show that uj2+α,1+α2C uniformly in j. Notice that ρj=uj-u satisfies the following linear problem:

{ t ρ j - ( 1 - σ ) Δ ρ j - σ ( a - 2 b w ) Δ ρ j + 2 σ b ( u j + u ) ρ j + σ d u ρ j - σ ( c - d u j ) ρ j = F j in  Ω × ( 0 , T ) , [ ρ j ] = 0 on  Ω × ( 0 , T ) , ρ j ( x , 0 ) = 0 ,

where

F j = ( σ - σ j ) Δ u j - 2 σ b ( w j - w ) Δ u j + ( σ j - σ ) ( a - 2 b w j ) Δ u j - 2 b ( σ - σ j ) | u j | 2 + ( σ j - σ ) u j ( c - d u j ) .

Using Schauder’s theory for linear parabolic equations, we get the estimate

u j - u 2 + α , 1 + α 2 = ρ j 2 + α , 1 + α 2 C F j α , α 2 C ( w j - w α , α 2 + | σ j - σ | ) .

Hence, 𝒯 is continuous in w and σ.

Furthermore, it is obvious that 𝒯(w,0)=0. Additionally, for any fixed point of 𝒯(u,σ)=u, the uniform estimates for quasilinear parabolic equation show ([8, Chapter V, Theorem 7.2]) that there exists a constant M depending only on a2b, c, u0 such that

u L , t u L M .

Therefore, by Leray–Schauder’s fixed point theorem, there exists a fixed point to the map 𝒯(,1), i.e. u is a solution of the following problem:

{ t u - ( a - 2 b u ) Δ u + 2 b | u | 2 = u ( c - d u ) in  Ω × ( 0 , T ) , [ u ] = 0 on  Ω × ( 0 , T ) , u ( x , 0 ) = u 0 ( x ) ,

which is equivalent to equation (1.1).

The uniqueness of classical solutions follows directly from comparison principles. ∎

3 Steady States

In this section, two results concerning stationary states are given. One of them shows that nontrivial nonnegative solutions do not exist for equation (1.1) with Dirichlet boundary condition. The other one has to do with the linear instability of constant steady states to equation (1.1) with homogeneous Neumann boundary condition.

3.1 Non-existence of Nontrivial Steady States for Dirichlet Boundary Conditions

The steady states corresponding to (1.1) satisfy the equation

(3.1) { - Δ [ ( a - b u ) u ] = ( c - d u ) u in  Ω , u = 0 on  Ω .

We write down a non-existence result (based on Pohozaev method, see [6]) which works for general elliptic problems, and explain how to use it specifically in the case of equation (3.1).

Theorem 2.

Let n>2, let Ω be a star shaped domain of Rn with respect to the origin and suppose that g,h are C1 functions defined on R+ such that

(3.2) n F ( s ) := n 0 s g ( v ) h ( v ) 𝑑 v < n - 2 2 g ( s ) h ( s ) for all  s > 0    𝑎𝑛𝑑    F ( u ( x ) ) 0 for all  x Ω .

Then the problem

(3.3) { - Δ h ( u ) = g ( u ) in  Ω , h ( u ) = 0 on  Ω ,

does not have any nontrivial (that is, different from u0) classical solution.

As a consequence, considering h(s):=(a-bs)s and g(s):=(c-ds)s, the sufficient condition (3.2) implies non-existence of (nontrivial, nonnegative, classical) solutions to equation (3.1) as soon as

b d s 2 + ( n - 6 6 a d - n + 6 6 b c ) s + a c < 0 for all  s > 0 .

This last condition is satisfied in particular when c and d are negative and 0<ad(n-6)-bc(n+6)<12abcd or ad(n-6)-bc(n+6)<0. Note also that since h(0)=0, the homogeneous Dirichlet boundary condition u=0 on Ω implies that h(u)=0 on Ω, so that the first part of Theorem 2 can be applied.

Proof.

By testing (3.3) with xh(u), we get

(3.4) - Ω ( x h ( u ) ) Δ h ( u ) 𝑑 x = Ω ( x h ( u ) ) g ( u ) 𝑑 x .

First, notice that

[ ( x h ( u ) ) h ( u ) ] = ( x h ( u ) ) Δ h ( u ) + k ( x k [ h ( u ) ] ) x k ( i x i h ( u ) x i )
= ( x h ( u ) ) Δ h ( u ) + | [ h ( u ) ] | 2 + 1 2 i x i x i | h ( u ) | 2 .

Integrating over Ω and applying the divergence lemma to the left-hand side, we get

Ω ( x h ( u ) ) ( h ( u ) ν ) 𝑑 σ = Ω ( x h ( u ) ) Δ h ( u ) 𝑑 x + Ω | h ( u ) | 2 𝑑 x + 1 2 Ω i x i x i | h ( u ) | 2 d x ,

so that using (3.4), we obtain

Ω ( x h ( u ) ) ( h ( u ) ν ) 𝑑 σ = - Ω ( x h ( u ) ) g ( u ) 𝑑 x + Ω | h ( u ) | 2 𝑑 x + 1 2 Ω i x i x i | h ( u ) | 2 d x
(3.5) = : I + I I + I I I .

We first compute

I = - Ω ( x h ( u ) ) g ( u ) 𝑑 x = - Ω g ( u ) h ( u ) i x i u x i d x = - Ω i x i F ( u ) x i d x ,

where F(u)=0ug(s)h(s)𝑑s. Integrating by parts, we get

- Ω i x i F ( u ) x i d x = n Ω F ( u ) 𝑑 x - Ω ( x ν ) F ( u ) 𝑑 σ ,

thus

(3.6) I = n Ω F ( u ) 𝑑 x - Ω ( x ν ) F ( u ) 𝑑 σ .

For the second term, using problem (3.3), we get

(3.7) I I = Ω | h ( u ) | 2 𝑑 x = Ω g ( u ) h ( u ) 𝑑 x .

For the last term in (3.5), we compute

( x 2 | h ( u ) | 2 ) = n 2 | h ( u ) | 2 + 1 2 i x i x i | h ( u ) | 2 ,

so that

I I I = Ω 1 2 i x i x i | h ( u ) | 2 d x = Ω ( div ( x 2 | h ( u ) | 2 ) - n 2 | h ( u ) | 2 ) 𝑑 x
(3.8) = 1 2 Ω x ν | h ( u ) | 2 𝑑 σ - n 2 Ω g ( u ) h ( u ) 𝑑 x .

Plugging (3.6), (3.7), (3.8) into (3.5), we obtain

Ω ( x h ( u ) ) ( h ( u ) ν ) 𝑑 σ = I + I I + I I I
= n Ω F ( u ) 𝑑 x - Ω ( x ν ) F ( u ) 𝑑 σ + Ω g ( u ) h ( u ) 𝑑 x
+ 1 2 Ω x ν | h ( u ) | 2 𝑑 σ - n 2 Ω g ( u ) h ( u ) 𝑑 x .

Using the Dirichlet boundary condition, we see that |h(u)|Ω=|νh(u)|Ω, so that on Ω, we have

( x [ h ( u ) ] ) ( ν [ h ( u ) ] ) = ( x ν ) | h ( u ) | 2 .

Thus, the above relation becomes

1 2 Ω ( x ν ) | h ( u ) | 2 𝑑 σ = Ω ( x h ( u ) ) ( h ( u ) ν ) 𝑑 σ - 1 2 Ω ( x ν ) | h ( u ) | 2 𝑑 σ
= n Ω F ( u ) 𝑑 x - Ω ( x ν ) F ( u ) 𝑑 σ - n - 2 2 Ω g ( u ) h ( u ) 𝑑 x ,

or

(3.9) 1 2 Ω ( x ν ) | h ( u ) | 2 𝑑 σ + Ω ( x ν ) F ( u ) 𝑑 σ = n Ω F ( u ) 𝑑 x - n - 2 2 Ω g ( u ) h ( u ) 𝑑 x .

Since Ω is star shaped, there exists α0 such that

x ν α Ω 𝑑 σ 0 ,

and relation (3.9) yields

n Ω F ( u ) 𝑑 x - n - 2 2 Ω g ( u ) h ( u ) 𝑑 x 0 .

Therefore, a sufficient condition for the non-existence of (nontrivial, nonnegative, classical) solutions is

(3.10) n 0 s g ( v ) h ( v ) 𝑑 v < n - 2 2 g ( s ) h ( s ) for all  s > 0 .

Next, if we set h(s)=(a-bs)s, g(s)=(c-ds)s, then

F ( u ) = b d 2 u 4 - a d + 2 b c 3 u 3 + a c 2 u 2

and

g ( u ) h ( u ) = b d u 4 - ( a d + b c ) u 3 + a c u 2 .

Using (3.10), we get the sufficient condition of non-existence of nontrivial solutions to the corresponding steady-state problem, which consists in finding a,b,c,d,n such that

b d s 2 + ( n - 6 6 a d - n + 6 6 b c ) s + a c < 0 for all  s > 0 .

As stated in the theorem, this happens when for example, c,d<0 and ad(n-6)-bc(n+6)<12abcd. ∎

Remark 3.

We obtain (3.10) from (3.9) by neglecting the first boundary term (since Ω is starshaped). Now we keep this first boundary integral (the second boundary integral in (3.9) is 0 because of the boundary conditions) and compute

1 2 Ω ( x ν ) | h ( u ) | 2 𝑑 σ α 2 ( Ω - h ( u ) ν d σ ) 2 = α 2 ( Ω - Δ h ( u ) d x ) 2 = α 2 ( Ω g ( u ) 𝑑 x ) 2 ,

where we have used the geometry of the domain, Cauchy–Schwarz inequality, |Ω|=1, the divergence lemma and the problem itself. Then relation (3.9) yields

n Ω F ( u ) 𝑑 x - n - 2 2 Ω g ( u ) h ( u ) 𝑑 x α 2 ( Ω g ( u ) 𝑑 x ) 2 .

Therefore, we can get a more precise description for the non-existence of solutions, since now we need to check the less stringent inequality

n Ω F ( u ) 𝑑 x - n - 2 2 Ω g ( u ) h ( u ) 𝑑 x < α 2 ( Ω g ( u ) 𝑑 x ) 2 .

Remark 4.

We present here a computation related to the linear stability of steady states for the Neumann boundary condition. We denote by {λk,ek}k=1 the solution of the eigenvalue problem for the Laplacian with homogeneous Neumann boundary condition, with λk0 for k=1,2, and 0=λ1<λ2λ3.

We assume that c,d>0. Then the equilibrium cd for equation (1.1) with homogeneous Neumann boundary condition is asymptotically linearly stable if and only if cda2b.

Indeed, we set u=cd+εw, so that the problem is transformed into

ε t w - Δ [ ( a - b c d - b ε w ) ( c d + ε w ) ] = ( c - d ( c d + ε w ) ) ( c d + ε w ) = - ε d w ( c d + ε w ) .

Thus

t w - Δ [ ( a - 2 b c d ) w ] = - c w + O ( ε ) .

By projecting the equation onto the k-th eigenspace (and by using the notation wk(t):=w(t,),ek), we obtain

d d t w k = ( ( 2 b c d - a ) λ k - c ) w k + O ( ε ) .

The condition for linear asymptotic stability of the steady state cd is therefore, for all k,

( 2 b c d - a ) λ k - c < 0 ,

whence the result.

4 Blow-Up Results

In this section, we present blow-up results (for different boundary conditions). Namely, we show that the solution to equation (1.1) blows up, under appropriate conditions, for both Dirichlet and Neumann boundary conditions, by using two different classical methods, i.e. Kaplan’s and concavity method.

4.1 A Sufficient Blow-Up Condition via Kaplan’s Method for Dirichlet Boundary Conditions

The problem under consideration in this subsection is

(4.1) { t u = Δ [ ( a - b u ) u ] + ( c - d u ) u in  Ω × ( 0 , T ) , u = 0 , on  Ω × ( 0 , T ) , u ( x , 0 ) = u 0 ( x ) 0 .

Let (μ,ϕ) be the solution to the eigenvalue problem

(4.2)

(4.2a) - Δ ϕ = μ ϕ , x Ω ,
(4.2b) ϕ = 0 , x Ω ,

where μ is the first eigenvalue and Ω is a connected bounded domain. Then μ>0 and ϕ is strictly positive and bounded in Ω. For convenience, we also impose the normalization condition Ωϕ(x)𝑑x=1. The main result in this subsection is:

Theorem 5.

Assume that Ω is a bounded smooth domain of Rn and let u0L1(Ω;ϕdx) satisfy

A 0 := Ω u 0 ϕ > max { μ a - c , 0 } μ b - d 𝑎𝑛𝑑 μ b > d ,

where (μ,ϕ) is the solution to (4.2). Then any nonnegative solution to problem (4.1) blows up in finite time in L1(Ω;ϕdx).

Remark 6.

When c=d=0, the above blow-up condition on the initial data can be roughly translated as u0>Cstab, which is coherent with our global existence result, and with the assumption of Theorem 7. Note also that the homogeneous Dirichlet boundary condition u=0 on Ω could be replaced in the theorem above by the less stringent condition (a-bu)u=0 on Ω.

Proof.

We begin, motivated by [5], with testing (4.1) with the eigenfunction ϕ, and set

A ( t ) = Ω ϕ ( x ) u ( x , t ) 𝑑 x ,

so that

A ( t ) = Ω ϕ Δ [ ( a - b u ) u ] 𝑑 x + Ω ( c - d u ) u ϕ 𝑑 x
= - μ Ω ( a - b u ) u ϕ 𝑑 x + Ω ( c - d u ) u ϕ 𝑑 x
= ( c - μ a ) Ω u ϕ + ( μ b - d ) Ω u 2 ϕ ,

where we have used problem (4.2). Next we recall that

μ b > d .

After applying Jensen’s inequality, we remind that Ωϕ(x)𝑑x=1, we get

A ( t ) ( c - μ a ) A ( t ) + ( μ b - d ) A 2 ( t ) ,

from which the blowup of the solution can be obtained. Namely, by using the change of variables

Ξ ( t ) = e - ( c - μ a ) t A ( t ) ,

we can obtain

Ξ ( t ) = - ( c - μ a ) e - ( c - μ a ) t A ( t ) + e - ( c - μ a ) t A ( t )
( μ b - d ) e - ( c - μ a ) t A 2 ( t ) = ( μ b - d ) e ( c - μ a ) t Ξ 2 ( t )

or (as long as Ξ(t)>0, remembering that Ξ(0)=A(0)>0)

( - 1 Ξ ) ( t ) ( μ b - d ) e ( c - μ a ) t 1 Ξ ( t ) 1 Ξ ( 0 ) - μ b - d c - μ a ( e ( c - μ a ) t - 1 ) .

When c-μa>0, we see that Ξ-1(t)>0 cannot remain true for

t t * := ( c - μ a ) - 1 log ( μ b - d ) A ( 0 ) + ( c - μ a ) ( μ b - d ) A ( 0 ) ,

so that blowup occurs before time t*. When c-μa<0, a similar computation shows that a blowup also occurs, under the extra assumption A0>μa-cμb-d. ∎

4.2 A Sufficient Blow-Up Condition by the Concavity Method for Neumann Boundary Condition

As has been stated in the beginning of Section 2, after the transformation v=u-a2b, the equation can be rewritten into

t v = - b Δ v 2 + ( c - d a 2 b - d v ) ( v + a 2 b ) .

In this subsection, we consider the following more general equation with homogeneous boundary condition,

(4.3) { t v = - b Δ v m + h ( v ) , in  Ω × ( 0 , T ) , v m ν = 0 , on  Ω × ( 0 , T ) , v ( x , 0 ) = v 0 ( x ) 0 in  Ω ,

and after giving a result about the blowup for the above general problem, we explain how (and under which conditions) it applies to problem (2.1). We refer the interested reader to [7, 3, 4, 10]. The main result of this subsection is the following:

Theorem 7.

Suppose that m>1, and that h is a continuous real function such that for all s0, one has smh(s)2H(s), where H(s):=0smtm-1h(t)𝑑t. We assume that v:=v(x,t) is a smooth nonnegative solution to problem (4.3) on [0,T] such that

(4.4) b 2 Ω | v 0 ( x ) m | 2 𝑑 x + Ω H ( v 0 ( x ) ) 𝑑 x > 0 .

Then there exists t*>0 (depending only on m, b, h and v(0,)) such that T<t*.

In other words, a blowup occurs before t*. More precisely,

(4.5) lim t t * 0 t Ω v m + 1 ( x , τ ) 𝑑 x 𝑑 τ = + .

Remark 8.

Note that the function h is not assumed to be nonnegative. Actually, Theorem 7 still holds when h is negative, or when it changes sign.

Remark 9.

When we consider problem (2.1) with homogeneous Neumann boundary condition, we are led to use m=2 and h(s):=(c-da2b-ds)(s+a2b) in Theorem 7. The condition smh(s)2H(s) (for all s0) becomes (remember that a,b>0, but the sign of c,d is not fixed)

(4.6) c min { a d b , a d 2 b } .

Coming back to the original unknown u (instead of v), Theorem 7 states that (under assumption (4.6)) the (smooth) solutions to (2.1) which are such that ua2b is pointwise true, cannot exist globally.

Note that a significant limitation of this result is related to the assumption that ua2b pointwise. Indeed, this estimate is propagated at the formal level by the equation only in very special cases in which a,b,c,d are linked by some equality, like when c=ad2b0.

Proof.

The proof is given by contradiction argument. Assume that the solution is global, and define

Ψ ( t ) := 0 t Ω v m + 1 ( x , τ ) 𝑑 x 𝑑 τ 0 .

The idea of the concavity method is to find an α>0 and a t00 such that Ψ-α is a concave function on [t0,+[. Then, from the concavity property of Ψ-α written at time t0 in a differential way, we get

Ψ - α ( t ) Ψ - α ( t 0 ) - α Ψ - α - 1 ( t 0 ) Ψ ( t 0 ) ( t - t 0 ) for all  t t 0 .

Using this inequality together with the fact that Ψ(t)0 for all t0, we obtain an upper bound t* for the blow-up time (that is, the first time t* such that Ψ-α(t*)=0)

t * Ψ - α ( t 0 ) + α t 0 Ψ - α - 1 ( t 0 ) Ψ ( t 0 ) α Ψ - α - 1 ( t 0 ) Ψ ( t 0 ) = Ψ ( t 0 ) + α t 0 Ψ ( t 0 ) α Ψ ( t 0 ) .

To prove this concavity property, we compute

( Ψ - α ) ′′ = α Ψ - α - 2 ( ( α + 1 ) ( Ψ ) 2 - Ψ Ψ ′′ ) ,

from which it can be deduced that a sufficient condition for Ψ-α to be concave (on [t0,+[) is that

Ψ ( t ) Ψ ′′ ( t ) - ( α + 1 ) ( Ψ ( t ) ) 2 0 for all  t t 0 .

In fact, we start by computing the derivative of the functional

Ψ ( t ) = Ω v m + 1 ( x , t ) 𝑑 x = 0 t Ω ( v m + 1 ) τ 𝑑 x 𝑑 τ + Ω v 0 m + 1 𝑑 x ,

and its second derivative

(4.7) Ψ ′′ ( t ) = Ω ( v m + 1 ) t 𝑑 x = ( m + 1 ) Ω v m v t 𝑑 x .

Next, we test (4.3) with (m+1)vm, and get

( m + 1 ) Ω v m v t 𝑑 x = - b ( m + 1 ) Ω v m Δ v m 𝑑 x + ( m + 1 ) Ω v m h ( v ) 𝑑 x
(4.8) = b ( m + 1 ) Ω | v m | 2 𝑑 x + ( m + 1 ) Ω v m h ( v ) 𝑑 x ,

so that substituting (4.8) into (4.7), we obtain

(4.9) Ψ ′′ ( t ) = b ( m + 1 ) Ω | v m | 2 𝑑 x + ( m + 1 ) Ω v m h ( v ) 𝑑 x .

We now test (4.3) with (vm)t, and get

0 Ω m v m - 1 ( v t ) 2 𝑑 x = - b Ω ( v m ) t Δ v m 𝑑 x + Ω ( v m ) t h ( v ) 𝑑 x
(4.10) = d d t { b 2 Ω | v m | 2 d x + Ω H ( v ) d x } = : d d t E ( t ) ,

where we recall that

H ( v ) := 0 v m s m - 1 h ( s ) 𝑑 s .

From (4.10), we can also deduce that

(4.11) E ( t ) - E ( 0 ) = 0 t Ω ( v m ) τ v τ 𝑑 x 𝑑 τ = m 0 t Ω v m - 1 2 v τ v m - 1 2 v τ 𝑑 x 𝑑 τ = 4 m ( m + 1 ) 2 0 t Ω [ ( v m + 1 2 ) τ ] 2 𝑑 x 𝑑 τ .

If the initial energy is strictly positive, namely

(4.12) E ( 0 ) > 0 ,

which is assumption (4.4) in Theorem 7, then we see that for all t0, E(t)>0, thanks to (4.11).

With the help of (4.10), identity (4.9) becomes

Ψ ′′ ( t ) = b ( m + 1 ) Ω | v m | 2 + ( m + 1 ) Ω v m h ( v ) 𝑑 x
= 2 ( m + 1 ) E ( t ) + ( m + 1 ) Ω v m h ( v ) 𝑑 x - 2 ( m + 1 ) Ω H ( v ) 𝑑 x .

At this point, we use the assumption on h,H to obtain

s m h ( s ) 2 H ( s ) = 2 0 s m r m - 1 h ( r ) 𝑑 r for all  s 0 ,

and conclude that

(4.13) Ψ ′′ ( t ) 2 ( m + 1 ) E ( t ) 2 ( m + 1 ) E ( 0 ) > 0 for all  t 0 .

From the above inequality, we also get that tΨ(t) is strictly increasing. Furthermore,

Ψ ′′ ( t ) Ψ ( t ) ( 4.13 ) 2 ( m + 1 ) E ( t ) Ψ ( t ) = 2 ( m + 1 ) E ( t ) 0 t Ω ( v m + 1 2 ) 2 𝑑 x 𝑑 τ
= ( 4.11 ) 2 ( m + 1 ) ( 4 m ( m + 1 ) 2 0 t Ω [ ( v m + 1 2 ) τ ] 2 𝑑 x 𝑑 τ + E ( 0 ) ) 0 t Ω ( v m + 1 2 ) 2 𝑑 x 𝑑 τ
> ( 4.12 ) 8 m m + 1 0 t Ω [ ( v m + 1 2 ) τ ] 2 𝑑 x 𝑑 τ 0 t Ω ( v m + 1 2 ) 2 𝑑 x 𝑑 τ
Cauchy–Schwarz     2 m m + 1 ( 0 t Ω ( v m + 1 ) τ 𝑑 x 𝑑 τ ) 2 = 2 m m + 1 ( Ψ ( t ) - Ω v 0 m + 1 𝑑 x ) 2
= 2 m m + 1 ( Ψ ( t ) - Ψ ( 0 ) ) 2 .

We now prove that there exists α>0 and t0>0 such that

2 m m + 1 ( Ψ ( t ) - Ψ ( 0 ) ) 2 ( α + 1 ) ( Ψ ( t ) ) 2 for all  t t 0 ,

or equivalently

(4.14) [ 1 - ( ( m + 1 ) ( α + 1 ) 2 m ) 1 2 ] Ψ ( t ) Ψ ( 0 ) for all  t t 0 .

In order to do so, we choose 0<α<m-1m+1 (remember that m>1). Due to the fact that Ψ(t)+ as t+ (because of (4.13)), we can indeed choose t0>0 large enough for (4.14) to hold.

Therefore, we finally obtain

Ψ ′′ ( t ) Ψ ( t ) ( α + 1 ) ( Ψ ( t ) ) 2 for all  t t 0 .

As observed at the beginning of the proof, the above inequality implies that we cannot extend the solution for all times, since (4.5) holds at a some point t*>0. ∎

5 Annex: Self-Similar Solutions Blowing Up in the Whole Space

In this annex, we provide a few explicit computations concerning problem (1.1) in the case when c=d=0:

t u = Δ [ ( a - b u ) u ] ,

and we still consider only the nonnegative solutions.

We recall that it is equivalent to studying the following problem:

(5.1) t v + b Δ v 2 = 0 ,

with u0 if and only if v-a2b.

This equation is a (reverse in time) porous medium equation, for which explicit solutions of Barenblatt type [1] can be computed (on a given time interval [0,T*) for the first type given below):

v ( x , t ) = 1 b ( T * - t ) ( ( T * - t ) 2 n + 2 - x 2 4 ( n + 2 ) ) + in  n × [ 0 , T * ) ,

and

v ( x , t ) = - 1 b ( T * + t ) ( ( T * + t ) 2 n + 2 - x 2 4 ( n + 2 ) ) + in  n × [ 0 , + ) .

It is possible to take linear combinations of those solutions and still get solutions, though the equation is nonlinear, as long as the support of those solutions remain separate (more precisely, when each two solutions have support with empty intersections during the time of existence of the solutions).

For an initial datum

v 0 ( x ) = i = 1 A 1 b T i ( T i 2 n + 2 - ( x - x i ) 2 4 ( n + 2 ) ) + - j = A + 1 A + B 1 b T j ( T j 2 n + 2 - ( x - x j ) 2 4 ( n + 2 ) ) + ,

with A,B, Tk>0, xkn (k=1,,A+B), the function defined by

v ( t , x ) = i = 1 A 1 b ( T i - t ) ( ( T i - t ) 2 n + 2 - ( x - x i ) 2 4 ( n + 2 ) ) + - j = A + 1 A + B 1 b ( t + T j ) ( ( t + T j ) 2 n + 2 - ( x - x j ) 2 4 ( n + 2 ) ) +

is a solution to equation (5.1) on the time interval [0,τ[ for

τ := min i = 1 , , A T i

(if A=0, τ=+) provided that

( τ + T j ) 1 n + 2 + ( τ + T k ) 1 n + 2 < | x k - x j | 2 n + 2 for all  j , k = A + 1 , , A + B , j k ,

and

(5.2) ( t + T j ) 1 n + 2 + ( T l - t ) 1 n + 2 < | x l - x j | 2 n + 2 for all  j = A + 1 , , A + B , l = 1 , , A , t [ 0 , τ ] .

Condition (5.2) can be rewritten without any direct reference to the time t in the following way:

  1. If Tl-Tj0, then

    T j 1 n + 2 + T l 1 n + 2 < | x l - x j | 2 n + 2 .

  2. If Tl-Tj2τ, then

    ( τ + T j ) 1 n + 2 + ( T l - τ ) 1 n + 2 < | x l - x j | 2 n + 2 .

  3. If Tl-Tj[0,2τ[, then

    2 n + 1 n + 2 ( T l + T j ) 1 n + 2 < | x l - x j | 2 n + 2 .

Note also that v(t,x)-b2a for all t[0,τ[ and xn as soon as for all j=A+1,,A+B, Tj>(a2)-1-2n.

The explicit solutions defined above feature in an explicit way the properties of blowup discussed previously. The value v=0 (or u=a2b) plays a decisive role in the existence or not of a blowup, as can be guessed from the study of the parabolicity regions of the equation.

Finally, we propose a figure illustrating the computations above. In Figure 1, a solution is drawn, with one positive bump and two negative ones, with the specific feature that when t=0 the branches of the bump coincide and connect. For this solution, we drew three different time instances.

Figure 1 
          Explicit solution shown at different times.
Figure 1

Explicit solution shown at different times.

6 Conclusion

This paper is a first attempt to tackle problems of the form

{ t u = Δ [ ( a - b u ) u ] + ( c - d u ) u in  Ω × ( 0 , T ) , [ u ] = 0 on  Ω × ( 0 , T ) , u ( x , 0 ) = u 0 ( x ) 0 ,

whose main characteristic is the fact that the quantity inside the Laplacian does not a priori have a fixed sign, so that global-in-time existence of solutions does not always hold. We proved the global existence and uniqueness of classical solutions for initial data and parameters such that the problem is of parabolic type. The non-existence of nontrivial steady states is studied, and some blow-up results using Kaplan’s method on the one hand, and the concavity method on the other hand, are also presented.


Communicated by Laurent Veron


Funding statement: This work was supported by DFG Project CH 955/3-1. The authors have been partially supported by the French “ANR blanche” project Kibord, grant ANR-13-BS01-0004, and from the Université Sorbonne Paris Cité, in the framework of the “Investissements d’Avenir”, grant ANR-11-IDEX-0005.

Acknowledgements

We thank Stephan Knapp for providing us with Figure 1.

References

[1] G. I. Barenblatt, On a class of exact solutions of the plane one-dimensional problem of unsteady filtration of a gas in a porous medium, Akad. Nauk SSSR. Prikl. Mat. Meh. 17 (1953), 739–742. Search in Google Scholar

[2] L. Chen, S. Göttlich and S. Knapp, Modeling of a diffusion with aggregation: rigorous derivation and numerical simulation, ESAIM: M2AN 52 (2018), 123–145. 10.1051/m2an/2018028Search in Google Scholar

[3] K. Deng and H. A. Levine, The role of critical exponents in blow-up theorems: The sequel, J. Math. Anal. Appl. 243 (2000), no. 1, 85–126. 10.1006/jmaa.1999.6663Search in Google Scholar

[4] M. Fila, Boundedness of global solutions of nonlinear diffusion equations, J. Differential Equations 98 (1992), no. 2, 226–240. 10.1016/0022-0396(92)90091-ZSearch in Google Scholar

[5] S. Kaplan, On the growth of solutions of quasilinear parabolic equations, Comm. Pure Appl. Math. 16 (1963), 327–343. 10.1002/cpa.3160160307Search in Google Scholar

[6] I. Kuzin and S. Pohozaev, Entire Solutions of Semilinear Elliptic Equations, Progr. Nonlinear Differential Equations Appl. 33, Birkhäuser, Basel, 1997. Search in Google Scholar

[7] H. A. Levine, The role of critical exponents in blowup theorems, SIAM Rev. 32 (1990), no. 2, 262–288. 10.1137/1032046Search in Google Scholar

[8] O. Ladyzenskaya, V. A. Solonnikov and N. N. Ural’ceva, Linear and Quasilinear Equations of Parabolic Type, American Mathematical Society, Providence, 1968. 10.1090/mmono/023Search in Google Scholar

[9] K. Oelschläger, Large systems of interacting particles and the porous medium equation, J. Differential Equations 88 (1990), 294–346. 10.1016/0022-0396(90)90101-TSearch in Google Scholar

[10] P. Quittner and P. Souplet, Superlinear Parabolic Problems. Blow-up, Global Existence and Steady States, Birkhäuser Adv. Texts Basler Lehrbücher, Birkhäuser, Basel, 2007. Search in Google Scholar

[11] A. S. Sznitman, Topics in propagation of chaos, École d’Été de Probabilités de Saint-Flour XIX-1989, Lecture Notes in Math. 1464, Springer, Berlin (1991), 165–251. 10.1007/BFb0085169Search in Google Scholar

Received: 2020-02-03
Revised: 2020-04-22
Accepted: 2020-04-29
Published Online: 2020-05-12
Published in Print: 2021-02-01

© 2020 Walter de Gruyter GmbH, Berlin/Boston

This work is licensed under the Creative Commons Attribution 4.0 International License.

Downloaded on 25.4.2024 from https://www.degruyter.com/document/doi/10.1515/ans-2020-2092/html
Scroll to top button