• Aucun résultat trouvé

Concentration in Lotka-Volterra parabolic or integral equations: a general convergence result

N/A
N/A
Protected

Academic year: 2022

Partager "Concentration in Lotka-Volterra parabolic or integral equations: a general convergence result"

Copied!
23
0
0

Texte intégral

(1)

Concentration in Lotka-Volterra parabolic or integral equations:

a general convergence result

Guy Barles, Sepideh Mirrahimi, Benoˆıt Perthame †‡

Abstract

We study two equations of Lotka-Volterra type that describe the Darwinian evolution of a population density. In the first model a Laplace term represents the mutations. In the second one we model the mutations by an integral kernel. In both cases, we use a nonlinear birth-death term that corresponds to the competition between the traits leading to selection.

In the limit of rare or small mutations, we prove that the solution converges to a sum of moving Dirac masses. This limit is described by a constrained Hamilton-Jacobi equation. This was already proved in [8] for the case with a Laplace term. Here we generalize the assumptions on the initial data and prove the same result for the integro-differential equation.

Key-Words: Adaptive evolution, Lotka-Volterra equation, Hamilton-Jacobi equation, viscosity solu- tions, Dirac concentrations.

AMS Class. No: 35B25, 35K57, 47G20, 49L25, 92D15

1 Introduction

We continue the study, initiated in [8], of the asymptotic behavior of Lotka-Volterra parabolic equa- tions. The model we use describes the dynamics of a population density. Individuals respond differ- ently to the environment, i.e. they have different abilities to use the available resources. To take this fact into account, population models can be structured by a parameter, representing a physiological (phenotypical) trait inherited from the parent, and that we denote by x ∈Rd. We denote by n(t, x) the density of trait x. The mathematical modeling in accordance with Darwin’s theory consists of two effects: natural selection and mutations between the traits (see [18, 24, 27, 25] for literature in adaptive evolution). We represent the birth and death rates of the phenotypical traits bya net growth rate R(x, I). The term I(t) is an ecological parameter that corresponds to a measure of the total population, whatever the trait, and that represents in the simpler possible way the resources (more precisely the inverse of it). We use two different models for mutations. A first possibility is to represent them by a Laplacian and, in an extreme and irrealistic simplification, we take them independent of birth, so as to write

Laboratoire de Math´ematiques et Physique Th´eorique, CNRS UMR 6083, F´ed´eration Denis Poisson, Universit´e Fran¸cois Rabelais, Parc de Grandmont, 37200 Tours, France. Email: barles@lmpt.univ-tours.fr

Universit´e Pierre et Marie Curie-Paris 6, UMR 7598 LJLL, BC187, 4, place Jussieu, F-75252 Paris cedex 5. Email:

mirrahimi@ann.jussieu.fr

Institut Universitaire de France. Email: benoit.perthame@upmc.fr

(2)

(∂tn−4n= nR(x, I(t)), x∈Rd, t≥0,

n(t= 0) =n0 ∈L1(Rd), n0 ≥0, (1)

I(t) = Z

Rd

ψ(x)n(t, x)dx. (2)

Here is a small term that we introduce to consider only rare mutations. It is also used to re-scale time to consider a much larger time than a generation scale.

A more natural way to model mutations is to use, instead of a Laplacian, an integral term that describes directly the mutation probability to generate a new-born of traitxfrom a mother with trait y. This yields

(∂tn= nR(x, I(t)) + 1 R 1

dK(y−x )b(y, I)n(t, y)dy, x∈Rd, t≥0,

n(t= 0) =n0 ∈L1(Rd), n0 ≥0, (3)

I(t) = Z

Rd

n(t, x)dx. (4)

Both types of models can be derived from individual based stochastic processes in the limit of large populations depending on the scales in mutations birth and death (see [13, 14]).

In this paper, we study the asymptotic behavior of equations (1)-(2) and (3)-(4) when vanishes.

Our purpose is to show that under some assumptions on R(x, I), n(t, x) concentrates as a sum of Dirac masses that are traveling. In biological terms, at every moment one or several dominant traits coexist while other traits disappear. The dominant traits change in time due to the presence of mu- tations.

We use the same assumptions as [8]. We assume that there exist two constantsψmM such that 0< ψm< ψ < ψM <∞, ψ∈W2,∞(Rd). (5) We also assume that there are two constants 0< Im< IM <∞ such that

min

x∈Rd

R(x, Im) = 0, max

x∈Rd

R(x, IM) = 0, (6)

and there exists constantsKi >0 such that, for any x∈Rd,I ∈R,

−K1 ≤ ∂R

∂I(x, I)≤ −K1−1 <0, (7)

sup

Im

2 ≤I≤2IM

kR(·, I)kW2,∞(Rd)< K2. (8) We also make the following assumptions on the initial data

Im ≤ Z

Rd

ψ(x)n0(x)≤IM, and ∃A, B >0, n0 ≤e−A|x|+B , (9)

(3)

and that there exist a pointx0∈Rd and positive constants L0 and M0 such that

eM0 ≤n0(x), for all |x−x0| ≤L0, (10) Note that assumption (10) means that initially we have some kind of biodiversity since it can be interpreted as different traits being sufficiently represented in the population.

Here we takeψ(x)≡1 for equations (3)-(4) because replacingnbyψnleaves the model unchanged.

For equation (3) we assume additionally that the probability kernelK(z) and the mutation birth rate b(z) verify

0≤K(z), Z

K(z)dz = 1, Z

K(z)e|z|2dz <∞, (11) bm ≤b(z, I)≤bM, |∇xb(z, I)|< L1b(z, I), |b(x, I1)−b(x, I2)|< L2|I1−I2|, (12) wherebm,bM,L1 and L2 are positive constants. Finally for equation (3) we replace (6) and (7) by

x∈minRd

R(x, Im) +b(x, Im)

= 0, max

x∈Rd

R(x, IM) +b(x, IM)

= 0, (13)

|R(x, I1)−R(x, I2)|< K3|I1−I2| and −K4 ≤ ∂(R+b)

∂I (x, I)≤ −K4−1<0, (14) whereK3 and K4 are positive constants.

In both cases, in the limit we expect n(t, x) = 0 or R(x, I) = 0, where n(t, x) is the weak limit of n(t, x) as vanishes. If we suppose that the latter is possible at only isolated points, we expect n to concentrate as Dirac masses. Following earlier works on the similar issue [19, 7, 8, 28], in order to study n, we make a change of variable n(t, x) = eu(t,x) . It is easier to study the asymptotic behavior ofuinstead ofn. In section 5 we study the asymptotic behavior ofu whilevanishes. We show that u, after extraction of a subsequence, converge to a functionu that satisfies a constrained Hamilton-Jacobi equation in the viscosity sense (see [3, 20, 16, 22] for general introduction to the theory of viscosity solutions). Our main results are as follows.

Theorem 1.1. Assume (5)-(10). Let n be the solution of (1)-(2), and u = ln(n). Then, after extraction of a subsequence,u converges locally uniformly to a functionu∈C((0,∞)×Rd), a viscosity solution to the following equation:

tu=|∇u|2+R(x, I(t)), max

x∈Rd

u(t, x) = 0, ∀t >0, (15)

I(t)−→

→0 I(t) a.e., Z

ψ(x)n(t, x)dx=I(t) a.e.. (16) In particular, a.e. in t, supp n(t,·) ⊂ {u(t,·) = 0}. Here the measure n is the weak limit of n

as vanishes. If additionally (u0) :=ln(n0) is a sequence of uniformly continuous functions which converges locally uniformly to u0 then u∈C([0,∞)×Rd) and u(0, x) =u0(x) in Rd.

(4)

Theorem 1.2. Assume (8)-(14), and(u0) is a sequence of uniformly Lipschitz-continuous functions which converges locally uniformly to u0. Let n be the solution of (3)-(4) with n0 = eu

0

, and u = ln(n). Then, after extraction of a subsequence, u converges locally uniformly to a function u ∈ C([0,∞)×Rd), a viscosity solution to the following equation:





tu=R(x, I(t)) +b(x, I(t))R

K(z)e∇u·zdz, max

x∈Rd

u(t, x) = 0, ∀t >0, u(0, x) =u0(x),

(17)

I(t)−→

→0 I(t) a.e., Z

n(t, x)dx=I(t) a.e.. (18)

In particular, a.e. in t, supp n(t,·)⊂ {u(t,·) = 0}. As above, the measure nis the weak limit of n

as vanishes.

These theorems improve previous results proved in [19, 8, 7, 29] in various directions. For the case where mutations are described by a Laplace equation, i.e. (1)-(2), Theorem 1.1 generalizes the assumptions on the initial data. This generalization derives from regularizing effects of Eikonal Hamil- tonian (see [26, 1, 2]). But our motivation is more in the case of equations (3)-(4) where mutations are described by an integral operator. Then we can treat cases where the mutation rateb(x, I) really depends onx, which was not available until now. The difficulty here is that Lipschitz bounds on the initial data are not propagated on u and may blow up in finite time (see [12, 5, 15] for regularity results for integral Hamiltonian). However, we achieve to control the Lipschitz norm by −u, that goes to infinity as|x|goes to +∞.

We do not discuss the uniqueness for equations (15) and (17) in this paper. The latter is studied, for some particular cases, in [8, 7].

A related, but different, situation arises in reaction-diffusion equations as in combustion (see [6, 9, 10, 21, 23, 30]). A typical example is the Fisher-KPP equation, where the solution is a progressive front. The dynamics of the front is described by a level set of a solution of a Hamilton-Jacobi equation.

The paper is organized as follows. In section 2 we state some existence results and bounds on n

and I. In section 3 we prove some regularity results foru corresponding to equations (1)-(2). We show that u are locally uniformly bounded and continuous. In section 4 we prove some analogous regularity results for u corresponding to equations (3)-(4). Finally, in section 5 we describe the asymptotic behavior of u and deduce the constrained Hamilton-Jacobi equation (15)-(16).

2 Preliminary results

We recall the following existence results for n and a priori bounds forI (see also [8, 17]).

Theorem 2.1. With the assumptions (5)-(8), and Im−C2 ≤I(0)≤ IM +C2, there is a unique solution n∈C(R+;L1(Rd)) to equations (1)-(2) and it satisfies

Im0 =Im−C2 ≤I(t)≤IM +C2=IM0 , (19) where C is a constant. This solution, n(t, x), is nonnegative for all t≥0.

(5)

We recall a proof of this theorem in Appendix A. We have an analogous result for equations (3)-(4):

Theorem 2.2. With the assumptions (8), (11)-(14), andIm ≤I(0)≤IM, there is a unique solution n∈C(R+;L1∩L(Rd)) to equations (3)-(4) and it satisfies

Im ≤I(t)≤IM. (20)

This solution, n(t, x), is nonnegative for allt≥0.

This theorem can be proved with similar arguments as Theorem 2.1. A uniform BV bound onI(t) for equations (1)-(2) is also proved in [8]:

Theorem 2.3. With the assumptions (5)-(9), we have additionally to the uniform bounds (19), the locally uniform BV and sub-Lipschitz bounds

d

dtI(t)≥ − C+e−Lt Z

ψ(x)n0(x)R(x, I0)

dx, (21)

d

dt%(t)≥ −Ct+ Z

(1 +ψ(x))n0(x)R(x, I0)

dx, (22)

where C and L are positive constants and %(t) =R

Rdn(t, x)dx. Consequently, after extraction of a subsequence,I(t) converges a.e. to a functionI(t), asgoes to 0. The limitI(t) is nondecreasing as soon as there exists a constant C independent of such that

Z

ψ(x)n0(x)R(x, I0)

≥ −Ceo(1) . We also have a local BV bound onI(t) for equations (3)-(4):

Theorem 2.4. With the assumptions (8)-(14), we have additionally to the uniform bounds (20), the locally uniform BV bound

d

dtI(t)≥ −C0+e−L

0t

Z

n0(x)R(x, I0) +b(x, I0)

dx, (23)

Z T 0

|d

dtI(t)|dt≤2C0T+C00, (24)

where C0, C00 and L0 are positive constants. Consequently, after extraction of a subsequence, I(t) converges a.e. to a function I(t), as goes to 0.

This theorem is proved in Appendix B.

3 Regularity results for equations (1)-(2)

In this section we study the regularity properties of u = lnn, where n is the unique solution of equations (1)-(2). We have

tn = 1

tueu , ∇n = 1

∇ueu , 4n = 1

4u+ 1

2|∇u|2 eu .

(6)

Consequentlyu is a smooth solution to the following equation

(∂tu−4u=|∇u|2+R(x, I(t)), x∈R, t≥0,

u(t= 0) =lnn0. (25)

We have the following regularity results for u.

Theorem 3.1. Assume (5)-(10) and let T >0 be given. Set D= B+ (A2 +K2)T. Then we have u ≤D2. For all t0>0, v =√

2D2−u are locally uniformly bounded and Lipschitz in [t0, T]×Rd,

|∇v| ≤C(T)(1 + 1

√t0), (26)

where C(T) is a constant depending on T, K1, K2, A and B. Moreover, if we assume that (u0) :=

ln(n0) is a sequence of uniformly continuous functions, then u are locally uniformly bounded and continuous in[0,∞[×Rd.

We prove Theorem 3.1 in several steps. We first prove an upper bound, then a regularizing effect inx, then local L bounds, and finally a regularizing effect in t.

3.1 An upper bound for u

From assumption (9) we have u0(x)≤ −A|x|+B. We claim that, with C=A2+K2,

u(t, x)≤ −A|x|+B+Ct, ∀t≥0. (27)

Define φ(t, x) =−A|x|+B+Ct. We have

tφ−4φ− |∇φ|2−R(x, I(t))≥C+A(d−1)

|x| −A2−K2 ≥0.

HereK2 is an upper bound forR(x, I) according to (8). We have also φ(0, x) =−A|x|+B≥u0(x).

So φ is a super-solution to (25) and (27) is proved.

3.2 Regularizing effect in space Letu=f(v), where f is chosen later. We have

tu=f0(v)∂tv, ∂xu=f0(v)∂xv, 4u=f0(v)4v+f00(v)|∇v|2. So equation (25) becomes

tv−4v−

f00(v)

f0(v) +f0(v)

|∇v|2= R(x, I)

f0(v) . (28)

Define p=∇v. By differentiating (28) we have

tpi−4pi−2

f00(v)

f0(v) +f0(v)

∇v· ∇pi

f000(v)

f0(v) −f00(v)2

f0(v)2 +f00(v)

|∇v|2pi

=−f00(v)

f0(v)2R(x, I)pi+ 1 f0(v)

∂R

∂xi

.

(7)

We multiply the equation bypi and sum over i:

t|p|2

2 −X

(4pi)pi−2

f00(v)

f0(v) +f0(v)

∇v· ∇|p|2 2 −

f000(v)

f0(v) −f00(v)2

f0(v)2 +f00(v)

|p|4

=−f00(v)

f0(v)2R(x, I)|p|2+ 1

f0(v)∇xR·p.

First, we compute P

i(4pi)pi. X

i

(4pi)pi =X

i

4p2i 2 −X

|∇pi|2

=4|p|2

2 −X

|∇pi|2

=|p|4|p|+|∇|p||2−X

i

|∇pi|2.

We also have

|∇|p||2 =X

i

|p·∂xip|2

|p|2 ≤X

i

|∂xip|2=X

i,j

|∂xipj|2 =X

j

|∇pj|2.

It follows that

X

i

(4pi)pi ≤ |p|4|p|.

We deduce

t|p| −4|p| −2

f00(v)

f0(v) +f0(v)

p· ∇|p| −

f000(v)

f0(v) −f00(v)2

f0(v)2 +f00(v)

|p|3 (29)

≤ −f00(v)

f0(v)2R(x, I)|p|+ 1

f0(v)∇xR· p

|p|. From (27) we know that, for 0 ≤t ≤ T, u ≤D(T)2, where D(T) = √

B+CT. Then we define f(v) =−v2+ 2D2, for vpositive, and thus

D(T)≤v, f0(v) =−2v, and | 1

f0(v)|= 1 2v ≤ 1

2D,

f00(v) =−2, and |f00(v) f0(v)2|= 1

2v2 ≤ 1 2D2,

(8)

f000(v) = 0, −

f000(v)

f0(v) −f00(v)2

f0(v)2 +f00(v)

= 2 +1 v2 >2.

From (29), Theorem 2.1, assumption (8) and these calculations we deduce

∂|p|

∂t −4|p| −2

f00(v)

f0(v) +f0(v)

p· ∇|p|+ 2|p|3− K2

2D2|p| − K2 2D ≤0.

Thus forθ(T) large enough we can write

∂|p|

∂t −4|p| −2

f00(v)

f0(v) +f0(v)

p· ∇|p|+ 2(|p| −θ)3 ≤0. (30) Define the function

y(t, x) =y(t) = 1 2√

t +θ.

Since y is a solution to (30), and y(0) =∞ and |p|being a sub-solution we have

|p|(t, x)≤y(t, x) = 1 2√

t+θ.

Thus forv =√

2D2−u, we have

|∇v|(t, x)≤ 1 2√

t+θ(T), 0< t≤T. (31)

See Appendix C for more details on the comparison principle used above.

3.3 Regularity in space of u near t= 0

Assume that u0 are uniformly continuous. We show that u are uniformly continuous in space on [0, T]×Rd.

For δ > 0 we prove that for h small |u(t, x+h) −u(t, x)| < δ. To do so define w(t, x) = u(t, x+h)−u(t, x). Sinceu0 are uniformly continuous, for h small enough |w(0, x)|< δ2. Besides w satisfies the following equation:

tw(t, x)−4w(t, x)−(∇u(t, x+h) +∇u(t, x))· ∇w(t, x) =R(x+h, I(t))−R(x, I(t)).

From Theorem 2.1 and using assumption (8) we have

tw(t, x)−4w(t, x)−(∇u(t, x+h) +∇u(t, x))· ∇w(t, x)≤K2|h|.

Therefore by the maximum principle we arrive at max

Rd

|w(t, x)|<max

Rd

|w(0, x)|+K2|h|t.

So for h small enough|u(t, x+h)−u(t, x)|< δon [0, T]×Rd.

(9)

3.4 Local bounds for u

We show thatuare bounded on compact subsets of ]0,∞[×Rd. We already know from section 3.1 that u is locally bounded from above. We show that it is also bounded from below onC= [t0, T]×B(0, R), for all R >0 and 0< t0< T.

From section 3.1 we have u(t, x) ≤ −A|x|+B+CT. So for R large enough there exists 0 such that for < 0

Z

|x|>R

eu dx <

Z

|x|>R

e−A|x|+B+CT dx < Im0M. We have also from (19) that

Z

Rd

eu dx > Im0 ψM. We deduce that forR large enough and for all 0< < 0

Z

|x|<R

eu dx > Im0M. Therefore there exists1>0 such that, for all < 1

∃x0∈Rd; |x0|< R, u(t, x0)>−1, thus v(t, x0)<p

2D2+ 1.

From Section 3.2 we know that v are locally uniformly Lipschitz

|v(t, x+h)−v(t, x)|< C(T) + 1 2√ t0

|h|,

Thus for all (t, x)∈ C and < 1

v(t, x)< E(t0, T, R) :=p

2D2(T) + 1 + 2 C(T) + 1 2√ t0

R.

It follows that

u(t, x)>2D2(T)−E2(t0, T, R).

We conclude that u are uniformly bounded from below onC.

If we assume additionally that u0 are uniformly continuous, with similar arguments we can show thatu are bounded on compact subsets of [0,∞[×Rd. To prove the latter we use uniform continuity of u instead of the Lipschitz bounds of v.

3.5 Regularizing effect in time

From the above uniform bounds and continuity results we can also deduce uniform continuity in time i.e. for all η > 0, there exists θ > 0 such that for all (t, s, x) ∈ [0, T]×[0, T]×B(0,R2), such that 0< t−s < θ, and for all < 0 we have:

|u(t, x)−u(s, x)| ≤2η.

(10)

We prove this with the same method as that of Lemma 9.1 in [4] (see also [11] for another proof of this claim). We prove that for any η >0, we can find positive constantsA,B large enough such that, for any x∈B(0,R2),s∈[0, T] and for every < 0,

u(t, y)−u(s, x)≤η+A|x−y|2+B(t−s), for every (t, y)∈[s, T]×B(0, R), (32) and

u(t, y)−u(s, x)≥ −η−A|x−y|2−B(t−s), for every (t, y)∈[s, T]×B(0, R). (33) We prove inequality (32), the proof of (33) is analogous. We fix (s, x) in [0, T[×B(0,R2). Define

ξ(t, y) =u(s, x) +η+A|y−x|2+B(t−s), (t, y)∈[s, T[×B(0, R),

where A and B are constants to be determined. We prove that, for A and B large enough, ξ is a super-solution to (25) on [s, T]×B(0, R) andξ(t, y)> u(t, y) for (t, y)∈ {s}×B(0, R)∪[s, T]×∂B(0, R).

According to section 3.4,u are locally uniformly bounded, so we can take A a constant such that for all < 0,

A≥ 8kukL([0,T]×B(0,R))

R2 .

With this choice, ξ(t, y)> u(t, y) on [0, T]×∂B(0, R), for all η,B andx∈B(0,R2). Next we prove that, for A large enough, ξ(s, y) > u(s, y) for all y ∈ B(0, R). We argue by contradiction. Assume that there exists η >0 such that for all constantsA there existsyA, ∈B(0, R) such that

u(s, yA,)−u(s, x)> η+A|yA,−x|2. (34) It follows that

|yA,−x| ≤ r2M

A ,

where M is a uniform upper bound for k u kL([0,T]×B(0,R)). Now let A → ∞. Then for all ,

|yA,−x| → 0. According to Section 3.3, u are uniformly continuous on space. Thus there exists h >0 such that if|yA,−x| ≤hthen|u(s, yA,)−u(s, x)|< η2, for all. This is in contradiction with (34). Therefore ξ(s, y) > u(s, y) for all y ∈ B(0, R). Finally, noting that R is bounded we deduce that forB large enough,ξ is a super-solution to (25) in [s, T]×B(0, R). Sinceu is a solution of (25) we have

u(t, y)≤ξ(t, y) =u(s, x) +η+A|y−x|2+B(t−s) for all (t, y)∈[s, T]×B(0, R).

Thus (32) is satisfied for t≥s. We can prove (33) for t≥s analogously. Then we putx =y and we conclude taking θ < Bη.

4 Regularity results for equations (3)-(4)

In this section we study the regularity properties of u = lnn, where n is the unique solution of equations (3)-(4) as given in Theorem 2.2. From equation (3) we deduce that u is a solution to the following equation

(11)

(

tu =R(x, I(t)) +R

K(z)b(x+z, I)eu(t,x+z)− u(t,x)dz, x∈R, t≥0,

u(t= 0) =lnn0. (35)

We have the following regularity results for u.

Theorem 4.1. Let n be the solution of (3)-(4) with n0 =eu

0

, and u=ln(n). With the assump- tions (8)-(14), and if we assume that (u0) is a sequence of uniformly bounded functions in W1,∞, thenu are locally uniformly bounded and Lipschitz in [0,∞[×Rd.

As in section 3 we prove Theorem 4.1 in several steps. We first prove an upper and a lower bound on u, then local Lipschitz bounds in space and finally a regularity result in time.

4.1 Upper and lower bounds on u

From assumption (9) we have u0(x)≤ −A|x|+B. As in section 3.1 we claim that

u(t, x)≤ −A|x|+B+Ct, ∀t≥0. (36)

Define v(t, x) =−A|x|+B+Ct, whereC=bMR

K(z)eA|z|dz+K2. Using (8) and (12) we have

tv−R(x, I(t))− Z

K(z)b(x+z, I)e

v(t,x+z)−v(t,x)

dz ≥C−K2−bM

Z

K(z)eA|z|dz ≥0.

We also have v(0, x) = −A|x|+B ≥u0(x). So v is a supersolution to (35). Since (3) verifies the comparison property, equation (35) verifies also the comparison property, i.e. if v and u are respec- tively super and subsolutions of (35) then u≤v. Thus (36) is proved.

To prove a lower bound onu we assume thatu0 are locally uniformly bounded. Then from equation (35) and assumption (8) we deduce

tu(t, x)≥ −K2, and thus

u(t, x)≥ −ku0kL(B(0,R))−K2t, ∀x∈B(0, R).

Moreover,|∇u0|being bounded, we can give a lower bound inRd

u(t, x)≥inf

u0(0)− k∇u0kL|x| −K2t, ∀x∈Rd. (37) 4.2 Lipschitz bounds

Here we assume thatu is differentiable inx (See [15]). See also Appendix D for a proof without any regularity assumptions onu.

Letp =∇u·χ, whereχis a fixed unit vector. By differentiating (35) with respect toχ we obtain

(12)

tp(t, x) =∇R(x, I(t))·χ+ Z

K(z)∇b(x+z, I)·χ eu(t,x+z)− u(t,x)dz +

Z

K(z)b(x+z, I)p(t, x+z)−p(t, x)

eu(t,x+z)−u(t,x)

dz.

Thus, using assumptions (8) and (12), we have

tp(t, x)≤K2+L1

Z

K(z)b(x+z, I)eu(t,x+z)−u(t,x)

dz (38)

+ Z

K(z)b(x+z, I)p(t, x+z)−p(t, x)

eu(t,x+z)− u(t,x)dz.

Define w(t, x) = p(t, x) +L1u(t, x) and ∆(t, x, z) = u(t,x+z)−u (t,x). From (38) and (35) we deduce

tw−K2(1 +L1)− Z

K(z)b(x+z, I)w(t, x+z)−w(t, x)

e(t,x,z)dz

≤2L1

Z

K(z)b(x+z, I)e(t,x,z)dz

−L1 Z

K(z)b(x+z, I)∆(t, x, z)e(t,x,z)dz

=L1 Z

K(z)b(x+z, I)e(t,x,z) 2−∆(t, x, z) dz

≤L1bMe,

noticing that e is the maximum of the function g(t) = et(2−t) in R. Therefore by the maximum principle, withC1=K2(1 +L1) +L1bMe, we have

w(t, x)≤C1t+ max

Rd

w(0, x).

It follows that

p(t, x)≤C1t+k ∇u0 kL +L1(B+Ct) +L1 k∇u0kL|x|+K2t−u0(x= 0)

(39)

=C2t+C3|x|+C4,

whereC2,C3 andC4 are constants. Since this bound is true for any|χ|= 1, we obtain a local bound on |∇u|.

4.3 Regularity in time

In section 4.2 we proved thatu is locally uniformly Lipschitz in space. From this we can deduce that

tu is also locally uniformly bounded.

LetC= [0, T]×B(x0, R) andS1 be a constant such thatkukL(C)< S1 for all >0. Assume that R0 is a constant large enough such that we haveu(t, x)<−S1 in [0, T]×Rd\B(x0, R0). According to (36) there exists such constant R0. We choose a constant S2 such that k ∇u kL([0,T]×B(x0,R0))< S2 for all >0. We deduce

(13)

|∂tu| ≤ |R(x, I(t))|+ Z

K(z)b(x+z, I)eu(t,x+z)−u(t,x)

1|x+z|<R0+1|x+z|≥R0

dz

≤K2+bM Z

K(z)eS2|z|1|x+z|<R0dz+bM Z

K(z)1|x+z|≥R0dz

≤K2+bM 1 + Z

K(z)eS2|z|dz . This completes the proof of Theorem 4.1.

5 Asymptotic behavior of u

Using the regularity results in sections 3 and 4, we can now describe the asymptotic behavior of u

and prove Theorems 1.1 and 1.2. Here we prove Theorem 1.1. The proof of Theorem 1.2 is analogous, except the limit of the integral term in equation (17). The latter has been studied in [19, 12, 7, 29].

Proof of theorem 1.1. step 1 (Limit) According to section 3, u are locally uniformly bounded and continuous. So by Arzela-Ascoli Theorem after extraction of a subsequence, u converges locally uni- formly to a continuous functionu.

step 2 (Initial condition)We proved that if u0 are uniformly continuous then u will be locally uniformly bounded and continuous in [0, T]×Rd. Thus we can apply Arzela-Ascoli neart= 0 as well.

Therefore we haveu(0, x) = lim

→0u(0, x) =u0(x).

step 3 max

x∈Rd

u = 0

Assume that for some t, x we have 0 < a ≤ u(t, x). Since u is continuous u(t, y)≥ a2 on B(x, r), for somer >0. Thus we haven(t, y)→ ∞, while→0. ThereforeI(t)→ ∞ while →0. This is a contradiction with (19).

To prove that max

x∈Rd

u(t, x) = 0, it suffices to show that lim

→0n(t, x)6= 0, for somex∈Rd. From (27) we have

u(t, x)≤ −A|x|+B+Ct.

It follows that for M large enough

lim→0

Z

|x|>M

n(t, x)dx≤lim

→0

Z

|x|>M

e−A|x|+B+Ct = 0. (40)

From this and (19) we deduce

→0lim Z

|x|≤M

n(t, x)dx≥ Im0 ψM. If u(t, x) < 0 for all |x| < M then lim

→0 eu(t,x) = 0 and thus lim

→0

R

|x|≤Mn(t, x)dx = 0. This is a contradiction with (40). It follows that max

x∈Rd

u(t, x) = 0, ∀t >0.

(14)

step 4 (supp n(t,·)⊂ {u(t,·) = 0})Assume thatu(t0, x0) =−a <0. Sinceuare uniformly contin- uous in a small neighborhood of (t0, x0), (t, x)∈[t0−δ, t0+δ]×B(x0, δ), we haveu(t, x)≤ −a2 <0 for small. We deduce that R

[t0−δ,t0+δ]×B(x0,δ)n dtdx=R

[t0−δ,t0+δ]×B(x0,δ)lim

→0eu(t,x) dtdx= 0. Therefore we have supp n(t,·)⊂ {u(t,·) = 0} for almost everyt.

step 5 (Limit equation)Finally we recall, following [8], how to pass to the limit in the equation.

Since u is a solution to (25), it follows that φ(t, x) = u(t, x)−Rt

0R(x, I(s))ds is a solution to the following equation

tφ(t, x)−4φ(t, x)− |∇φ(t, x)|2−2∇φ(t, x)· Z t

0

∇R(x, I(s))ds

= Z t

0

4R(x, I(s))ds+| Z t

0

∇R(x, I(s))ds|2.

Note that we have I(s) → I(s) for all s≥ 0 as goes to 0, and on the other hand, the function R(x, I) is smooth. It follows that we have the locally uniform limits

→0lim Z t

0

R(x, I(s))ds= Z t

0

R(x, I(s))ds,

→0lim Z t

0

∇R(x, I(s))ds= Z t

0

∇R(x, I(s))ds,

lim→0

Z t 0

4R(x, I(s))ds= Z t

0

4R(x, I(s))ds,

for all t ≥ 0. Moreover the functions Rt

0R(x, I(s))ds, Rt

0∇R(x, I(s))ds and Rt

0 4R(x, I(s))ds are continuous. According to step 1,u(t, x) converge locally uniformly to the continuous functionu(t, x) as vanishes. Therefore φ(t, x) converge locally uniformly to the continuous function φ(t, x) = u(t, x)−Rt

0 R(x, I(s))ds asvanishes. It follows that φ(t, x) is a viscosity solution to the equation

tφ(t, x)− |∇φ(t, x)|2−2∇φ(t, x)· Z t

0

∇R(x, I(s))ds

=| Z t

0

∇R(x, I)ds|2. In other wordsu(t, x) is a viscosity solution to the following equation

tu(t, x) =|∇u(t, x)|2+R(x, I(t)).

A Proof of theorem 2.1

A.1 Existence

LetT >0 be given and A be the following closed subset:

A ={u∈C [0, T], L1(Rd)

, u≥0, ku(t,·)kL1≤a},

(15)

wherea= R n0dx

eK2T. Let Φ be the following application:

Φ : A→A u7→v, wherev is the solution to the following equation

(∂tv−4v= vR(x, I¯ u(t)), x∈R, t≥0,

v(t= 0) =n0. (41)

Iu(t) = Z

Rd

ψ(x)u(t, x)dx, (42)

and ¯R is defined as below

R(x, I) =¯





R(x, I) if I2m < I <2IM, R(x,2IM) if 2IM ≤I, R(x,Im2 ) if I ≤ I2m. We prove that

(a) Φ defines a mapping ofA into itself, (b) Φ is a contraction forT small.

With these properties, we can apply the Banach-Picard fixed point theorem and iterate the con- struction with T fixed.

Assume that u ∈A. In order to prove (a) we show that v, the solution to (41), belongs to A. By the maximum principle we know thatv≥0. To prove the L1 bound we integrate (41)

d dt

Z

vdx= Z v

R(x, I¯ u(t))dx≤ 1 max

x∈Rd

R(x, I¯ u(t)) Z

vdx≤ K2

Z vdx,

and we conclude from the Gronwall Lemma that kvkL1

Z n0dx

eK2T =a.

Thus (a) is proved. It remains to prove (b). Let u1, u2 ∈A, v1= Φ(u1) and v2= Φ(u2). We have

t(v1−v2)−4(v1−v2) = 1

(v1−v2) ¯R(x, Iu1) +v2 R(x, I¯ u1)−R(x, I¯ u2) .

Noting that k v2 kL1≤ a, and |R(x, I¯ u1)−R(x, I¯ u2)| ≤ K1|Iu1 −Iu2| ≤ K1ψM k u1−u2 kL1 we obtain

d

dt kv1−v2kL1≤ K2

kv1−v2 kL1 +aK1ψM

ku1−u2 kL1 . Usingv1(0,·) =v2(0,·) we deduce

(16)

kv1−v2kL

t L1x≤ aK1ψM K2

(e

K2T

−1)ku1−u2 kL

t L1x . Thus, for T small enough such that eK2T(eK2T −1)< 2K K2

1ψMR

n0, Φ is a contraction. Therefore Φ has a fixed point and there existsn∈A a solution to the following equation

(∂tn−4n = nR(x, I(t)),¯ x∈R,0≤t≤T, n(t= 0) =n0.

I(t) = Z

Rd

ψ(x)n(t, x)dx,

With the same arguments as A.2 we prove that I2m < I(t) < 2IM and thus n is a solution to equations (1)-(2) for t∈ [0, T]. We fix T small enough such that eK2T(eK2T −1)< 4KK2ψm

1ψMIM. Then we can iterate in time and find a global solution to equations (1)-(2).

A.2 Uniform bounds on I(t) We have

dI dt = d

dt Z

Rd

ψ(x)n(t, x)dx= Z

Rd

ψ(x)4n(t, x)dx+ 1

Z

Rd

ψ(x)n(t, x)R(x, I(t))dx.

We define ψL = χL·ψ ∈ W2,c(Rd), where χL is a smooth function with a compact support such thatχL|B(0,L)≡1,χL|R\B(0,2L)≡0. Then by integration by parts we find

Z

Rd

ψL(x)4n(t, x)dx= Z

Rd

L(x)n(t, x)dx.

AsL→ ∞,ψL converges toψ inWloc2,∞(Rd). Therefore we obtain

L→∞lim Z

Rd

L(x)ndx= Z

Rd

4ψ(x)ndx,

L→∞lim Z

Rd

ψL(x)4n(t, x)dx= Z

Rd

ψ(x)4n(t, x)dx.

From these calculations we conclude dI

dt = Z

Rd

4ψ(x)n(t, x)dx+1

Z

Rd

ψ(x)n(t, x)R(x, I(t))dx.

It follows that

−C1

ψmI+1 Imin

x∈Rd

R(x, I)≤ dI

dt ≤C1

ψmI+ 1 Imax

x∈Rd

R(x, I).

LetC = Cψ1K1

m . As soon asI overpasses IM +C2, we haveR(x, I)<−CK2

1 =−2ψC1

m and thus dIdt becomes negative. Similarly, as soon as I becomes less than Im −C2, dIdt becomes positive. Thus (19) is proved.

(17)

B A locally uniform BV bound on I

for equations (3)-(4)

In this appendix we prove Theorem 2.4. We first integrate (3) overRd to obtain d

dtI(t) = 1

Z

n(t, x) R(x, I(t)) +b(x, I(t)) dx.

Define J(t) = dtdI(t). We differentiateJ and we obtain d

dtJ(t) = 1 J(t)

Z

n(t, x)∂(R+b)

∂I (x, I(t))dx + 1

2 Z

R(x, I) +b(x, I)

n(t, x)R(x, I) + Z

K(y−x)b(y, I)n(t, y)dy dx.

We rewrite this equality in the following form d

dtJ(t) = 1 J(t)

Z

n(t, x)∂(R+b)

∂I x, I(t)

dx+ 1 2

Z

n(t, x) R x, I(t)

+b x, I(t)2

dx + 1

2 Z Z

K(y−x) R x, I(t)

−R y, I(t)

b y, I(t)

n(t, y)dydx + 1

2 Z Z

K(y−x) b x, I(t)

−b y, I(t)

b y, I(t)

n(t, y)dydx.

It follows that d

dtJ(t)≥ 1 J(t)

Z

n(t, x)∂(R+b)

∂I x, I(t) dx+ 1

2 Z

n(t, x) R x, I(t)

+b x, I(t)2

dx

−K2+bML1

Z Z

K(z)|z|b x+z, I(t)

n(t, x+z)dzdx

≥ 1 J(t)

Z

n(t, x)∂(R+b)

∂I x, I(t) dx+ 1

2 Z

n(t, x) R x, I(t)

+b x, I(t)2

dx−C1 , whereC1 is a positive constant. Consequently, using (14) we obtain

d

dt(J(t))≤ C1

−C2

(J(t)), with (J(t))= max(0,−J(t)). From this inequality we deduce

(J(t))≤ C1

C2

+ (J(0))e

C2t . With similar arguments we obtain

(J(t))+≥ −C10

C20 + (J(0))+e

C0 2t ,

with (J(t))+ = max(0, J(t)). Thus (23) is proved. Finally, we deduce the locally uniform BV bound (24)

Z T 0

|d

dtI(t)|dt= Z T

0

d

dtI(t)dt+ 2 Z T

0

(d

dtI(t))dt

≤IM −Im+ 2C0T +O(1).

(18)

C Complement to the proof of the regularizing effect (26)

In this section, we provide some details for the comparison principle used in the proof of (26). In Subsection 3.2 we proved thatp=∇v satisfies the following (see the inequality (30))

∂|p|

∂t −4|p| −2h v −2vi

p· ∇|p|+ 2(|p| −θ)3≤0.

To apply the comparison principle we first claim the following lemma that we will prove at the end of this section.

Lemma C.1. Assume (8) and (10). Then, there exist positive constants A1, B1 and D1 such that, for all t1 >0 and ≤1,

−A1|x|2+B1+D1t

t1 ≤u(t, x), for (t, x)∈(t1,+∞)×Rd. (43) The above lemma implies that

D(T)≤v≤ r

2D2+ 1

t1(B1+D1T +A1|x|2), for (t, x)∈(t1,+∞)×Rd. We deduce that, for some positive constants A2 andD2(T),

∂|p|

∂t −4|p| − 1

√t1

[A2|x|+D2(T)]|p| · ∇|p|+ 2(|p| −θ)3 ≤0, for (t, x)∈(t1,+∞)×Rd. (44) Define, for (t, x)∈(t1, T]×BR(0) and for A3 to be chosen later,

z(t, x) = 1 2√

t−t1 + A3R2

√t1(R2− |x|2) +θ.

We prove that, forA3 =A3(T) chosen large enough,zis a strict supersolution of (44) in (t1, T]×BR(0).

To this end, we compute

tz(t, x) =− 1 4(t−t1)√

t−t1

,

∇z(t, x) = 2A3R2x

√t1(R2− |x|2)2,

∆z(t, x) = 2A3R2

√t1(R2− |x|2)2 + 8A3R2|x|2

√t1(R2− |x|2)3. We then replace this in (44) to obtain

∂z

∂t −∆z−1

t1(A2|x|+D2(T))|z∇z|+ 2(z−θ)3 =

4(t−t1

1)

t−t1t 2A3R2

1(R2−|x|2)2 +t8A3R2|x|2

1(R2−|x|2)3

1

t1(A2|x|+D2)(2t−t1

1 +t A3R2

1(R2−|x|2)+θ)t2A3R2|x|

1(R2−|x|2)2

+2(2t−t1

1 +tA3R2

1(R2−|x|2))3

2A3R2

t1(R2−|x|2)2 +t 8A3R4

1(R2−|x|2)3

1

t1(A2R+D2)(2t−t1

1 + t A3R2

1(R2−|x|2)+θ)t 2A3R3

1(R2−|x|2)2

+(t−t3

1)t A23R4

1(R2−|x|2)2 + 2t A33R6

1

t1(R2−|x|2)3,

Références

Documents relatifs

In order to give a sense to problem (1.1) in such a setting, we use the theory of viscosity solutions for equations with a measurable dependence in time which was first studied by

The problem of strong convergence of the gradient (hence the question of existence of solution) has initially be solved by the method of Minty-Browder and Leray-Lions [Min63,

The coarse crust model is finally amplified using either real-world elevation data or procedural landforms to generate a high resolution representation of the relief according to

Related topics include modeling of variability across different life-cycle stages of software systems; patterns, styles and tactics; practices for

Key-Words: Integral parabolic equations, adaptive dynamics, asymptotic behavior, Dirac concen- trations, population dynamics..

We refer the reader to [8, 17] for other works on the existence of critical traveling fronts propagating with the linear speed c ∗.. It bears mentioning that there are choices

To summarize, precise asymptotic expansions show that the large time behavior of non- negative solutions to (1) with sufficiently localized initial data is determined by the

Concentration in Lotka-Volterra parabolic or integral equations: a general convergence result.. Guy Barles, Sepideh Mirrahimi,