A Singular Limit in a Non-Local Reaction-Diffusion Equation in Periodic Media ()
1. Introduction
We concentrate on the following equation:
(1.1)
In the above setting,
is a constant,
is a l-periodic function,
is given and the term
denotes a fractional elliptic operator which is defined as follows:
(1.2)
where
denotes the principal value and
is a l-periodic smooth function such that for all
with b and B positive constants. When
is constant, we recover the classical fractional Laplacian
.
Let
be N given real numbers. As stated below, the statement about a function
is l-periodic means that
for all
. Let
be the period cell defined by
We assume
is called the nonlocal dispersal operator with
being a
nonnegative convolution kernel, and for all
(1.3)
where
and
.
The main goal of this paper is to describe the spread of the front associated with Equation (1.1). It is demonstrated that the stable state advances into the unstable state at an exponential speed.
Equation (1) was initially presented by Fisher in 1937 [1] and by Kolmogorov, Petrovskii, and Piscunov in 1938 [2] in the specific context of a homogeneous environment (
) and standard diffusion (
), corresponding to
,
and
. Subsequently, Aronson and Weinberger[3] demonstrated a result akin to ours for the scenario proposed by Fisher and Kolmogorov, Petrovskii, and Piscunov in [1]. In this instance, the propagation occurs at a uniform speed, regardless of direction. Freidlin and Gartner, in their 1985 work [4], examined this question with a standard Laplacian in a heterogeneous environment (
periodic). Employing a probabilistic approach, they established that the propagation speed varies based on the chosen direction
, albeit remaining constant within that direction. Alternative proofs using PDE techniques are available in [5] and [6]. Turning to the fractional Laplacian and a constant environment, Cabre and Roquejoffre demonstrated in [7] that the front position grows exponentially over time (see also [8] for heuristic and numerical studies predicting this behavior, and [9] for an alternative proof). Further, in [10], Cabre, Coulon, and Roquejoffre investigated the speed of propagation in a periodic environment modeled by Equation (1.1), but with the fractional Laplacian replacing the L operator and
. Notably, in the fractional case, the speed of propagation becomes independent of direction. They proved that the propagation speed increases exponentially with time, with a specific exponent determined by a periodic principal eigenvalue.
In recent times, nonlocal population models have garnered considerable attention (refer to [11] [12]). Initially, the population’s dispersal mechanism was considered nonlocal, implying that standard Gaussian diffusion could be replaced by a process influenced by Levy flights. Such nonlocal dispersal strategies have been observed in nature (see [13]). Moreover, the potential nonlocal characteristic of the species growth rate is motivated by real-world situations where populations benefit not just from nearby resources but also from those within their “influence zone” (analogous to the “giraffe’s neck” effect) (refer to [14]). This nonlocal aspect can be modeled through convolution with an integrable kernel. It is noteworthy, from a mathematical perspective, that the two types of nonlocal operators are fundamentally distinct; the fractional Laplacian tends to decrease the differentiability of the function, whereas convolution exerts a smoothing effect.
The aim of this work is to furnish an alternative proof for this property by utilizing an asymptotic methodology known as “approximation of geometric optics.” Our focus lies in examining the long-term behavior of the solution u. We show that, within the set defined by
, as t approaches infinity, u converges to a stationary state
, whereas outside this domain, u tends towards zero. The core concept in this approach involves executing a long-time, long-range rescaling to capture the effective behavior of the solution (refer to [9] [15] [16]). This paper is closely related to [9] where the authors Meleard and Mirrahimi have introduced such an “approximation” for a model with the fractional Laplacian and a simpler reaction term
.
A significant contribution of this paper lies in the innovative rescaling techniques we introduce, which facilitate the description of the population’s asymptotic dynamics. We employ two rescaling methods. The first one, focusing on long-range and long-term perspectives, hinges on the propagation speed. The underlying concept is to observe the population from a distant vantage point, akin to homogenization, where we disregard intricate details but capture the population’s propagation. Consequently, we propose a novel asymptotic formalism tailored for reaction-diffusion equations involving the fractional Laplacian. This formalism extends the classical approach known as “geometric optics approximation,” which is extensively developed for reaction-diffusion equations with the traditional Laplacian.
Next, under the assumptions on
, the operator
admits a principal eigenpair
that is
(1.4)
We will provide a detailed proof later on. To assure the existence of a bounded, positive and periodic steady solution
for (1.1), we will assume that the principal eigenvalue
is negative:
(H1)
Note that such stationary solution is unique in the class of positive and periodic stationary solutions.
The first result of this paper ensures the existence and the uniqueness of a positive bounded stationary state
of (1.1), the stationary problem for (1.1) is written as:
(1.5)
Theorem 1. Under the assumption (H1), there exists a unique positive and bounded solution
of (1.5). Moreover it is l-periodic.
Next, We introduce the following rescaling
(1.6)
We perform this rescaling because one expects that for t large enough, u is close to the stationary state
in the following set
and u is close to 0 in the set
. The change of variable (1.6) will therefore respect the geometries of these sets. We then rescale the solution of (1.1) as follows
and a new steady state
We prove:
Theorem 2. Assuming (H1), let u be the solution of (1.1). Then
(i)
, locally uniformly in
,
(ii)
, locally uniformly in
.
In section 2, we prove the existence and uniqueness of the solution of Equation (1.1), and introduce some main results and technical tools to prepare the following theorems. In section 3, We prove the existence and uniqueness of the non-trivial bounded stationary solution of Equation (1.1) by its eigenvalues and eigenfunction properties. In section 4, we provide a sub and a super-solution of Equation (1.1) and demonstrate Theorem 2.
2. Preliminary Results
We initially present a well-established principle concerning the fractional heat kernel.
Proposition 2.1. There exists a positive constant
larger than 1 such that the heat kernal
associated to the operator
verifies the following inequalities for
:
(2.1)
The proof of this proposition is given in [17]. Next, we give the Existence and uniqueness of the solution and comparison principle of (1.1).
In this section, we initially revisit the concept of a mild solution pertaining to the nonhomogeneous linear problem
(2.2)
where
, and
are given. The mild solution of (3.5) is:
for all
and
(
is given in Proposition 2.1). One easily checks that
. Next, the existence and uniqueness theorem and proof of solution (1.1) are given:
Proposition 2.2. Let
,
,
satisfying
(2.3)
and
. Given any
, Our focus is on the nonlinear problem
(2.4)
Then, for a given
, there is a unique unique global mild solution u to (2.4) with
.
Proof. We can obtain it by referring to reference [18].
We now claim the following comparison principle. We present a more general version of the comparison theorem. In theorem proving, we think of
as a part of
.
Lemma 2.3.(comparison principle) Let
, The function
is of class
Lipschitz-continuous with respect to u. If
are such that
,
, and
, and satisfy
on
, then
on
.
Proof. We can obtain it by referring to reference [19] Theorem 2.4.
Next, we will give a property about nonlocal dispersal operator
. We’ll use that later on when we prove theorem 2.
Proposition 2.4 Let
be positive constants, let
be a
nonnegative bounded function and satisfies (1.3). For all
,
. Then there exists a positive constant C, which does not depend on x, such that, for all
:
(i)
(2.5)
(ii) for all
,
(2.6)
Proof of (i). Let
be a positive constant and N be an enough large positive constant. By the compactness argument, we only have to prove it for
. We computer
Let us begin by approximating
.
For
, we write:
But we know that
, using that
, we deduce that
Thus, we deduce
To control
, we write
in the following form:
Next, we define
Since for all
, the map that
is
, we know that
is well defined. we deduce that the map
is continuous and so we conclude that
is bounded independently of x. Hence,
. Combining the above inequalities, we obtain that there exists a constant
such that for all
,
Proof of (ii). For convenience, we write
. Using the above inequality, we can conclude with a change of variable
(where
):
where
. Finally, we obtain
Now we use this proposition 2.1 and proposition 2.4 to demonstrate that beginning with a positive compactly supported initial date leads to a solution with algebraic tails.
Propossition 2.5. There exist two positive constants
and
depending on
, and
such that for all
, we have
(2.7)
Proof. Defining
, Thanks to the comparison principle, we have the following inequalities, for all
We begin with the proof that
.
Let
be the solution of:
(2.8)
where
. Thanks to Proposition 1, we can solve (2.8) and find
Thus for any
, we obtain
A simple substitution is made for equation (2.1), for any
, we can obtain
Thanks to the dominated convergence theorem, for any
, we have:
Therefore we conclude by a compactness argument that for any
,
, then there exist a constant
depending on
such that:
(2.9)
Using the same approach, we can obtain that for any
, then there exist a constant
depending on
such that:
(2.10)
Next, we will proceed to prove for all
, such that
(2.11)
From (2.5) of Proposition 2.4 and (2.10), we can obtain that
takeing
.
Thanks to the comparison principle and (2.11), we have the following inequalities, for all
(2.12)
where the last
in a new constant depending only on
and
.
In the same way, we can get
where the last
in a new constant depending only on
and
. By combining the above inequality and (2.12) together, we finally obtain
(2.13)
We next provide an important lemma that will be useful all along the article. Since, the same kind of result can be found in the appendix A of [8] and [9], we do not provide the proof of this lemma. Note that here, the lemma is stated with less regularity on
such than in [9]. Nevertheless, there is no difficulty to adapt the proof. Before introducing the lemma. We first introduce the computation of
of a product of functions:
(2.14)
with
The calculation of
leads to the calculation of
, we denote
, then
(2.15)
with
(2.16)
Lemma 2.6. Let
and d be two positive constants in
such that
and
. Let
be a periodic positive function and
. Then there exists a positive constant C, such that we have for all
1. for all
,
2. for all
,
3. for all
,
where
.
Proof. For the proof of (i) (ii)(iii), we refer to Appendix A of reference [9].
3. Existence and Uniqueness of the Stationary State u+
The conditions for existence, uniqueness, and the long-term behavior of solution for (1.1) are determined by the principal eigenvalue, denoted as
, of the operator B, which is defined by
with periodicity conditions. Firstly, we show that the operator B has a positive periodic first eigenfunction. We mainly prove it through Krein-Rutman Theorem and Lax-Milgram Theorem.
Proposition 3.1. The operator B has a unique positive periodic eigenfunction with eigenvalue
, that is
(3.1)
The eigenvalue
is unique both algebraically and geometrically, and it represents the lowest point in the spectrum for B.
Proof. Before we prove it, let’s do a sign change. Let
(3.2)
Obviously, K is positive, l-periodic and
in x, and symmetic in h and is singular at
and there exists
for all
(3.3)
To tackle the issue of periodicity, we have formulated this eigenfunction on the torus
. A smooth function u, which has a period l in
, can be represented in an equivalent manner as a smooth function on
. Moreover,
(3.4)
where
please note that, considering (3.3), the sum mentioned above converges for all
and
.
It also follows from (3.3) and the definition of K, for a constant
, and all
For
, we have
and we define
where
(3.5)
is a bilinear mapping. For
, and we define
where
and
denote, respectively, the homogeneous and inhomogeneous Sobolev norm, this is,
For all
, for any given
we find
where
is a constant.
For
, we show that u satisfies a Garding-type inequality, we find
Subsequently, for a sufficiently large constant
, it is evident that the operator
is bijective. Moreover, the operator
is positive compact operator, fulfilling the conditions of the Lax-Milgram theorem with in
. Consequently, due to the Sobolev embedding theorem, it possesses a compact inverse that is defined everywhere
(3.6)
According to the Krein-Rutman Theorem, there exists a positive eigenfunction
of
with eigenvalue
. Hence,
This corresponds exactly to Equation (1.4) with
. Evidently,
is the smallest eigenvalue in the spectrum of
.
Next, to prove Theorem 1.
Proof of Theorem 1. Let’s first define
for any
. We will demonstrate the existence of a positive and periodic solution in the stationary equation. It is well known that stationary solution are provided by
(3.9)
Let
represent the only positive solution of
with
. For
sufficiently small, it follows that
(3.10)
Hence, it can be concluded that
Futhermore,
serves as a subsolution of equation (3.9) with periodicity conditions. Additionally, if
for any
, then the constant M acts as an upper solution of (3.9) with periodicity conditions, and for
sufficiently small,
in
. Consequently, by employing an iterative method, it is established that there exists a periodic classical solution u to (3.9) fulfilling the inequality
in
. Define the order interval
We define
. For any given
, We may assume that
is sufficiently large, using the variational method, we see that boundary value problem of linear equation
(3.11)
admits a unique solution
which is denoted by
.
We prove that T is monotone in the order interval
, i.e.,
if
and
. We assume that
, Since
satisfies Lipschitz continuity with respect to
, we find
In fact, writing
, then
satisfies
(3.12)
The maximum principle [20] yields
, i.e.,
. Defined
By the monotonicity of T,
. We claim that the sequences
and
satisfy
(3.13)
Let
. Since
satisfies
(3.14)
we see that v satisfies
(3.15)
The Maximum principle leads to
, i.e.,
. Similary,
. Thus,
. Applying the monotonicity of T we can deduce (3.13) inductively.
As
, We have
. Take
, then the
theory asserts
. This leads to
, and
. Therefore
by the Schauder theory. In the above, positive constants
and
are all independent of i. Noting that
↪
is compact, there are sub-sequences of
and
converging to
and
in
, respectively. As these sequences are monotone and bounded, by the uniqueness of limits,
and
in
. Taking
in
and
, we see that
and
are solutions of (3.9).
Next, we prove the uniqueness of the stationary solution.
We first prove that for all given
, (3.9) has a unique positive solution in
. It is easy to know (3.9) that there are maximal solution
and minimal solution
in
and
. We know that
, we have
for
. According to symmetry and (3.9), we can obtain
hence
(3.16)
because
and
is monotonically decreasing about u, we can obtain that
. For formula (3.16) to hold, we have
for all
. By the arbitrariness of
, we have
for all
.
Next, Let us define that
are any two bounded positive solutions of (3.9), then there exists
such that . We have
and we define the same mapping T such that
proceed in turn, we have
as these sequences are monotone and bounded, by the uniqueness of limits
,
. Here
is the only positive solution of (3.9) in
. Hence, by (3.16) we can obtain that
for
. In the same way, we can get that
for
. Additionally, if u is a positive solution to equation (3.9), then the function defined by
for each
, is also a positive solution. Consequently, u exhibits periodicity. This establishes of Theorem 1.
4. The Proof of Theorem 2
In this section, we will provide the proof of Theorem 2. Let us redefine (1.1) in accordance with the rescaling specified by (1.6).
Notation We rescale the solution of (1.1) as follows:
Next, the equation becomes
(4.1)
where
. In the same way, we define
Moreover, according to the inequalities (2.7), we can consider
as our initial data instead of
. So we can replace the
by:
Next we are going to provide sub and super-solutions to (1.1).
Theorem 3. We assume (H1) and we choose positive constants
and
and
such that
Then there exists a positive constant
such that for all
we have:
(i)
is a super-solution of (4.1),
(ii)
is a sub-solution of (4.1).
(iii) Moreover, if we assume (H1) and
and
where
and
are given by (2.13) then for all
,
Proof. Since the proofs of (i) and (ii) follow from similar arguments, we will only provide the proof of (i) and (iii). Proof of (i). We define:
(4.2)
Then, noticing that
is independent of t, we first bound
from below,
The last inequality is obtained from the definition of
and
. Actually, for such
and
, we have, for any positive non-null constant A, the following relation:
(4.3)
because,
We also compute
as a fractional Laplacian of a product of functions,
(4.4)
with
given in section 2. Similarly, we denote
, then
(4.5)
with
given in section 2. Replacing this in Equation (4.1) and using the three previous results (4.3), (4.4) and (4.5), we find:
where we have used (1.4) and (H1) for the last equality. In order to control
and
, we will utilize Lemma 2.6.
For
noticing that
, and
thanks to the point (i) of Lemma 2.6 we obtain:
But, comparing the growths, there exists
such that for
and for all
:
hence:
By the same way, comparing the growths, there exists
such that for
and for all
:
hence:
Now we deal with
in a similar fashion. Thanks to Lemma 2.6, we find:
Then, noticing that for any choice of
is strictly positive, we deduce there exists
such that for all
:
We deduce that
In the same way, We deduce there exists
such that for all
:
We set:
Then, we conclude that for
, we have:
Therefore
is a super-solution of (4.1) and this concludes the proof of the point (i).
Proof of (iii). From (2.13), since
and
, we have:
Then, according to Lemma 2.3, we obtain:
and hence
(4.6)
Thanks to the inequalities stated in (4.6), we are now in a position to prove Theorem 2. In order to accomplish this, we will adopt the approach outlined by [9] and [16].
Proof of Theorem 2. First, we perform a Hopf-Cole transformation
(4.7)
Taking the logarithm in (4.6) and multiplying by
, we find:
Define
Letting
, we obtain
We then let
and we obtain
We deduce that
converges locally uniformly in
to n since the above limits are locally uniform in
.
Proof of (i). For any compact set K in
, there exists a positive constant a such that for all
, we have
. It is thus immediate from (4.7) that
converges uniformly to 0 in
. This concludes the proof of (i).
Proof of (ii). We can obtain it by referring to reference [20].
5. Conclusion
In this paper, we provide an asymptotic analysis of a nonlocal reaction-diffusion equation in periodic media and with a nonlocal stable operator of order
. The objective of this work is to provide an alternative proof of this property using an asymptotic approach known as “approximation of geometric optics”. We will be interested in the long-time behavior of the solution u. Our focus lies in examining the long-term behavior of the solution u. We show that, within the set defined by
, as t approaches infinity, u converges to a stationary state
, whereas outside this domain, u tends towards zero. The core concept in this approach involves executing a long-time, long-range rescaling to capture the effective behavior of the solution (refer to [10] [12]).