Volume 2014, Issue 1 523656
Research Article
Open Access

A Reaction-Diffusion System with Nonlinear Nonlocal Boundary Conditions

Alexander Gladkov

Corresponding Author

Alexander Gladkov

Department of Mechanics and Mathematics, Belarusian State University, Nezavisimosti Avenue 4, 220030 Minsk, Belarus bsu.by

Search for more papers by this author
Alexandr Nikitin

Alexandr Nikitin

Department of Mathematics, Vitebsk State University, Moskovsky Avenue 33, 210038 Vitebsk, Belarus vstu.by

Search for more papers by this author
First published: 20 February 2014
Citations: 3
Academic Editor: Roberto Natalini

Abstract

We consider initial boundary value problem for a reaction-diffusion system with nonlinear and nonlocal boundary conditions and nonnegative initial data. We prove local existence, uniqueness, and nonuniqueness of solutions.

1. Introduction

In this paper we consider the following semilinear reaction-diffusion system with nonlinear nonlocal boundary conditions:
()
where p, q, m, n > 0, Ω is a bounded domain in N  for  N ≥ 1 with smooth boundary Ω, T > 0. Here, c1(x, t),   c2(x, t) are nonnegative Hölder continuous functions defined for , and t ∈ [0, T] and k1(x, y, t),   k2(x, y, t) are nonnegative continuous functions defined for xΩ, , and t ∈ [0, T]. The initial data u0(x),   v0(x) are nonnegative continuous functions satisfying the boundary conditions at t = 0.

In the past several decades, many physical phenomena have been formulated into nonlocal mathematical models. Initial boundary value problem for semilinear reaction-diffusion equations and systems with nonlocal boundary conditions has been analyzed by many authors (see, e.g., [112] and the references therein). Local and global existence, comparison principle, and various qualitative properties have been discussed.

We note that for max⁡⁡(p, q, m, n) < 1 the nonlinearities in (1) are non-Lipschitzian. The problem of uniqueness and nonuniqueness for different nonlinear parabolic equations and systems with non-Lipschitzian data has been addressed by several authors. See, for example, [1316] for equations and [1721] for systems. In particular, the authors of [16] have investigated the uniqueness of solution for a problem (1) with a single equation.

In [7] the authors have considered a problem (1) with ci(x, t) = 1 and ki(x, y, t) = ki(x, y),   i = 1,2. They have proved a comparison principal and investigated the blowup properties of the positive solutions. The aim of this paper is to study the uniqueness of the nonnegative solution of the problem (1) for any p, q, m, n > 0.

The plan of this paper is as follows. In the next section we prove a comparison principle; an existence theorem of a local solution is given in Section 3; uniqueness of solutions with nontrivial initial data, uniqueness of solution with trivial initial datum for min⁡(p, q, m, n) ≥ 1; nonuniqueness of solution with trivial initial datum for min⁡(pq, m, n) < 1 are proved in Section 4.

2. Comparison Principle

Let us introduce the definitions of a subsolution and a supersolution. For the remainder of this paper we denote QT = Ω × (0, T), ST = Ω × (0, T).

Definition 1. A pair of nonnegative functions is called a subsolution of problem (1) in QT if

()
and a pair of nonnegative functions is called a supersolution of problem (1) in QT if the reversed inequalities hold in (2).

Definition 2. A pair of functions (u, v) is called a solution of (1) in QT if it is both a subsolution and a supersolution of problem (1) in QT.

Definition 3. We say that solution (u, v) of (1) is positive in QT if u > 0 and v > 0 in QT.

To establish the uniqueness results we need a comparison principle. We prove it in a different way, not as in the work of [7].

Theorem 4. Let and be a nonnegative supersolution and a nonnegative subsolution of problem (1) in QT, respectively. Suppose that and or and in if min⁡(p, q, m, n) < 1. If and for , then and in .

Proof. Let be a nonnegative function such that . Then , where ν is the unit outward normal to the lateral boundary of QT. By the definition of a subsolution we have

()
If we multiply (3) by ξ and then integrate over Qt for 0 < t < T, we get
()

On the other hand, the supersolution satisfies (4) with reversed inequality. Set . Then we have

()
where Θ1,  Θ2 are nonnegative continuous functions if min⁡(p, q, m, n) ≥ 1 and positive continuous functions if min⁡(p, q, m, n) < 1 in which satisfy the following equalities:
()

Obviously there exists a positive constant M1 such that in . Since c1(x, t), k1(x, y, t) are nonnegative and continuous functions, then there exists a constant M2 > 0 such that 0 ≤ c1(x, t) ≤ M2, 0 ≤ k1(x, y, t) ≤ M2 in and , respectively.

Consider the following backward problem in Qt:

()
where 0 ≤ ψ(x) ≤ 1. By the maximum principle for the heat equation 0 ≤ ξ ≤ 1, it is easy to show that −M3ξ/ν ≤ 0 on St for some M3 ≥ 0.

Let s+ = max⁡(s, 0). Then from (5), we get

()
where K1 = pM1M2,   K2 = mM1M2M3|Ω|, and |Ω| is the Lebesgue measure of Ω. Consider the sequence of functions {ϕn(x)}, , converging in L1(Ω) to ϕ(x), defined as follows:
()
Replacing ψ(x) by ϕn(x) in (8) and passing to the limit as n, we have
()
Using a similar argument for the inequality ,  xΩ,  0 < t < T, we get
()
where K3, K4 are positive constants. Adding (10) and (11), we have
()
where K5 = max⁡(K1 + K4, K2 + K3).

Applying now Gronwall′s inequality, we conclude that

()
Since w+ + z+ ≥ 0, then , .

3. Local Existence

Let {εl} be decreasing to 0 sequence such that 0 < εl < 1. For ɛ = εl let u0ɛ(x),   v0ɛ(x) be the functions with the following properties: ; for i < j; u0ɛ(x) → u0(x),   v0ɛ(x) → v0(x) as ɛ → 0; for xΩ.

Due to the nonlinearities in (1), the Lipschitz condition is not satisfied if min⁡⁡(p, q, m, n) < 1, and thus we need to consider the following auxiliary problem:
()

Theorem 5. For small values of T, (14) has a unique solution in QT.

Proof. We start the proof with the construction of a supersolution of (14). Let max⁡(supΩu0ɛ(x),   supΩv0ɛ(x)) ≤ C,   C > 0. Denote . Introduce an auxiliary function   φ(x) with the following properties:

()
where  . Let α, β be positive constants such that αqβ = βpα and
()
Note that αqβ > 0 for pq > 1 and αqβ ≤ 0 for pq ≤ 1. Obviously the pair of functions (ε, ε) is a subsolution of problem (14). We show that
()
is a supersolution of (14) in QT for T ≤ min⁡(1/α, 1/β, 1/(αqβ)) if pq > 1 and T ≤ min⁡(1/α, 1/β) if pq ≤ 1. We have
()
for (x, t) ∈ QT. On the other hand, we get that
()
for (x, t) ∈ ST. Similarly, we can show that
()
To prove the existence of a solution of the problem (14) we introduce the set
()
Clearly B is a nonempty convex subset of .

Consider the following problem:

()
where (s1, s2) ∈ B. Problem (22) has a nontrivial positive solution. Let us call A(s1, s2) = (u, v). In order to show that A has a fixed point in B we verify that A is a continuous mapping from B into itself such that AB is relatively compact. Thanks to the comparison principle for (22) we have that A maps B into itself.

Let G(x, y; t) denote the Green′s function for the heat equation given by

()
with zero boundary condition. Then (u, v) is a solution of (22) if and only if
()
()

We claim that A is continuous. In fact, let {(s1k, s2k)} be a sequence in B converging to (s1, s2) ∈ B in . Denote (uk, vk) = A(s1k, s2k). Then we see that

()
where r = max⁡(Θ1, Θ2) and
()
Choosing T so small that r < 1, we conclude that (uk, vk)→(u, v) in as k. The equicontinuity of AB follows from (24), (25) and the properties of the Green′s function (see, e.g. [22]). The Ascoli-Arzela theorem guarantees the relative compactness of AB. Thus, we are able to apply the Schauder-Tychonoff fixed point theorem and conclude that A has a fixed point in B if T is small. Since (u, v) is a fixed point of A, it is a solution of (14). Uniqueness of solution follows from a comparison principle for (14) which can be proved in a similar way as in the previous section.

Using Theorem 5, we can prove the following local existence theorem of a solution of problem (1).

Theorem 6. For small values of T (1) has a maximal solution in QT.

Proof. Let ε2 > ε1. It is easy to show that is a supersolution of the problem (14) with ε = ε1. Then ,  . Using these inequalities and the continuation principle of solutions we deduce that the existence time of (uɛ(x, t), vɛ(x, t)) does not decrease as ε → 0. Let ε → 0, then

()
and (umax⁡(x, t), vmax⁡(x, t)) exist in QT for some T > 0.

Moreover, by dominated convergence theorem, (umax⁡(x, t), vmax⁡(x, t)) satisfies the following equations:

()
The interior regularity of (umax⁡(x, t), vmax⁡(x, t)) follows from the continuity of (umax⁡(x, t), vmax⁡(x, t)) in QT and the properties of the Green′s function. Obviously (umax⁡(x, t), vmax⁡(x, t)) satisfies (1). Let (y1(x, t), y2(x, t)) be any other solution of (1). Then by the comparison principle uε(x, t) ≥ y1(x, t), vε(x, t) ≥ y2(x, t). Taking ε → 0, we conclude that umax⁡y1(x, t), vmax⁡(x, t) ≥ y2(x, t).

To prove the positiveness of nontrivial solutions we need the following definition.

Definition 7. We say that a function g(x, t) has the property (N) if there exist xkΩ and tk > 0,  k such that g(xk, tk) > 0,  k, and tk → 0 as k.

Remark 8. Note that if a nonnegative function g(x, t) has no the property (N) then g(x, t) ≡ 0 in Qτ for some τ > 0.

Theorem 9. Let either u0(x) or v0(x) be a nontrivial function in Ω. Supposing that k1(x, ·, t) and k2(x, ·, t) are nontrivial functions for any xΩ and t ∈ (0, T),  c1(x, t) has the property (N) if u0(x) ≡ 0,  and  c2(x, t) has the property (N) if v0(x) ≡ 0. Let (u(x, t), v(x, t)) be a supersolution of (1) in QT. Then (u(x, t), v(x, t)) is positive in for 0 < t < T.

Proof. Suppose for definiteness that u0(x) is a nontrivial function. We show at first that u(x, t) > 0 in for 0 < t < T. We have

()
then by strong maximum principle a minimum of u(x, t) should be attained in on a parabolic boundary. Thus, u(x, t) > 0 in QT, otherwise, there would be a contradiction with the initial datum. We show that u(x, t) > 0 on Ω × (0, T). Let there exist a point (x0, t0) ∈ ST such that u(x0, t0) = 0. But u(x, t0) > 0 for xΩ. By boundary conditions (1) and assumption for k1(x, y, t) we have u(x, t) > 0 for (x, t) ∈ ST. This contradiction shows that u(x, t) > 0 on ST, and therefore u(x, t) > 0 in for 0 < t < T.

Now we show the positiveness of v(x, t). If v0(x) is a nontrivial function, then v(x, t) > 0 in for 0 < t < T by previous arguments. If v0(x) ≡ 0 we suppose that there exists a constant τ > 0 such that v(x, t) ≡ 0 in Qτ since otherwise we can use the arguments from the beginning of the proof again. But this is a contradiction with the second equation in (1) since u(x, t) > 0 in Qτ and c2(x, t) has the property (N). Hence, we conclude that v(x, t) > 0 in for 0 < t < T.

4. Uniqueness and Nonuniqueness

As a simple consequence of Theorem 4 and Theorem 9 we get the first uniqueness result for problem (1).

Theorem 10. Let problem (1) have a solution in QT with nonnegative initial datum for min⁡(p, q, m, n) ≥ 1 and with positive initial datum under conditions min⁡(p, q, m, n) < 1 and k1(x, ·, t),   k2(x, ·, t) are nontrivial functions for any xΩ and t ∈ (0, T). Then solution of (1) is unique in QT.

Now we show nonuniqueness of solutions of problem (1) with trivial initial datum for min⁡(pq, m, n) < 1.

Theorem 11. Let min⁡(pq, m, n) < 1, u0(x) = v0(x) ≡ 0. Suppose that the maximal solution of problem (1) exists in QT. Assume that at least one of the following conditions is fulfilled:

()
()
()
Then the maximal solution of problem (1) is nontrivial in QT.

Proof. In the local existence theorem we constructed a maximal solution (umax⁡(x, t), vmax⁡(x, t)) of (1) in the following way: umax⁡(x, t) = limε→0uε(x, t), vmax⁡(x, t) = limε→0vε(x, t), where (uε(x, t), vε(x, t)) is some positive supersolution of (1). To prove the theorem we construct a nontrivial nonnegative subsolution of some problem with trivial initial datum. By the comparison principle we conclude that , and therefore maximal solution is a nontrivial solution.

Consider at first the case when pq < 1 and c1(x0, t0) > 0,   c2(x0, t0) > 0 for some x0Ω and 0 ≤ t0 < T. Then there exists a neighborhood U(x0) of x0 in Ω and such that c1(x, t) ≥ c0,   c2(x, t) ≥ c0,   c0 > 0 for .

Consider the following problem:

()
where w0(x) is a bounded nontrivial nonnegative continuous function which satisfies a boundary condition. By the strong maximum principle 0 < w(x, t) < M0 for , where .

Let α = (1 + p)/(1 − pq) > 1,   β = (1 + q)(1 − pq) > 1. Denote that

()
where C = min⁡(c0/(αM0), c0/(βM0)). Note that if t = t0 or xU(x0). After simple calculations we obtain
()
Similarly we can get that
()
Then and (uε(x, t), vε(x, t)) are subsolution and supersolution, respectively, of the following problem:
()
By comparison principle for (38) we conclude that , and hence ,  , for .

Now consider the case when m < 1 and k1(x, y1, t1) > 0 for any xΩ and some y1Ω,  0 ≤ t1 < T. We will consider the following problem:

()
where T0 > t1 will be defined later. Construct a subsolution of (39) using the change of variables in a neighborhood of Ω as in [23]. Let . We denote by the inner unit normal to Ω at the point . Since Ω is smooth it is well known that there exists δ > 0 such that the mapping ψ : Ω × [0, δ] → N given by defines new coordinates in a neighborhood of Ω in . A straightforward computation shows that, in these coordinates, Δ applied to a function which is independent of the variable , evaluated at a point is given by
()
where denote the principal curvatures of Ω at .

Under the made assumption there exists such that k1(x, y, t)>0 for , where V(y1) is some neighborhood of y1 in . Let α > 1/(2(1 − m)) and assume that 0 < ξ0 ≤ 1 and . For points in Ω × [0, δ]×(t1, T0] of coordinates define

()
and extend as zero to the whole . Using (40), we get that
()
for sufficiently small values of ξ0. It is clear that
()
It remains to verify the validity of the inequality
()
for xΩ and t1 < t < T0. Here is Jacobian of the change of variables. Estimating the integral I in the right-hand side of (44) we get
()
where constant C does not depend on t; we obtain that (44) is true if T0t1 is small enough.

By comparison principle for (39) we conclude that , and, respectively, ,   for (x, t) ∈ Ω × (t1, T0).

The proof in the case n < 1 and k2(x, y2, t2) > 0 for any xΩ and some y2Ω,   0 ≤ t2 < T is similar.

It is easy to get from Theorem 9 and the proof of Theorem 11 the following statement.

Corollary 12. Let the conditions of Theorem 11 hold with ti = 0,   i = 0,1, 2. Suppose also that k1(x, ·, t) and k2(x, ·, t) are nontrivial functions for any xΩ and t ∈ (0, T),   c2(x, t) has the property (N) if (32) is realized, and c1(x, t) has the property (N) if (33) is realized. Then maximal solution is positive in for 0 < t < T.

Under the conditions of Corollary 12 for some class of the coefficients ci(x, t) and ki(x, y, t),   i = 1,2, we can prove the uniqueness of solution for (1) with trivial initial datum which is positive for all positive times as long as it exists.

Theorem 13. Let the conditions of Corollary 12 hold. Suppose also that there exists t0 > 0 such that for 0 ≤ tt0 the functions c1(x, t), c2(x, t),  k1(x, y, t),   and  k2(x, y, t) are nondecreasing with respect to t.

Then there exists exactly one solution of problem (1) which is positive in for 0 < t < T.

Proof. Suppose that there exists different from (umax⁡(x, t), vmax⁡(x, t)) solution (u(x, t), v(x, t)) of (1) with trivial initial datum which is a positive in for 0 < t < T. Denote t* = min⁡(t0, T). Due to the conditions of the theorem it is easy to see that u(x, t + τ), v(x, t + τ) is positive supersolution of (1) with trivial initial datum in for any τ ∈ (0, t*). By Theorem 4 we have umax⁡(x, t) ≤ u(x, t + τ), vmax⁡(x, t) ≤ v(x, t + τ) for every 0 ≤ tt*τ. Passing to the limit as τ → 0 we get umax⁡(x, t) ≤ u(x, t),   vmax⁡(x, t) ≤ v(x, t) for 0 ≤ tt*. Hence umax⁡(x, t) = u(x, t),   vmax⁡(x, t) = v(x, t) in QT.

Note that by Theorem 10, the solution (u(x, t), v(x, t)) of (1) is unique if min⁡(p, q, m, n) ≥ 1. Now we specify our uniqueness result in the case min⁡(pq, m, n) < 1.

Theorem 14. Let the conditions of Corollary 12 fulfill only u0(x)≢0 or v0(x)≢0. Then the solution of (1) is unique.

Proof. To prove the uniqueness of the solution if min⁡(pq, m, n) < 1, it suffices to show that if (u(x, t), v(x, t)) is any solution of (1), then

()
where (umax⁡(x, t), vmax⁡(x, t)) is a maximal solution of (1).

First, consider the case when 0 < p < 1,   0 < q ≤ 1,   0 < m < 1,   and  0 < n ≤ 1. Let

()
Then (w1, w2) satisfies
()
for some T1 > 0. By Theorem 13, there exists a unique solution (h1(x, t), h2(x, t)) of the following problem:
()
such that h1(x, t) > 0,   h2(x, t) > 0 for for some T2 > 0. Let T0 = min⁡(T1, T2). In a similar way as in Theorems 13 and 4 we can prove that umax⁡(x, t) ≥ h1(x, t) ≥ w1(x, t),   vmax⁡(x, t) ≥ h2(x, t) ≥ w2(x, t). Put a1 = h1w1,   a2 = h2w2. Now, use an elementary inequality, which is recalled for instance in [20],
()
where 0 < k ≤ 1,  min⁡(a, b, c) > 0, and max⁡(a, c) ≤ ba + c. Then we obtain
()
We show that a1(x, t) > 0,   a2(x, t) > 0 in . In fact, otherwise, by Theorem 9 there exists such that a1(x, t) = a2(x, t) ≡ 0 in . Suppose that c1(x0, 0) > 0 for some x0Ω. Then we get
()
for xΩ and . This is a contradiction since h2(x, t) > 0,   v(x, t) > 0 in c1(x0, t0) > 0 for some , and 0 < p < 1.

If k1(x, y0, 0) > 0 for any xΩ and some y0Ω we can obtain a contradiction by another way. Indeed,

()
for xΩ and and we get again a contradiction since h1(x, t) > 0,  u(x, t) > 0 in , and 0 < m < 1.

Since a1 > 0,   a2 > 0 in by comparison principle with arguments of Theorems 13 and 4 we conclude that a1h1, a2h2 in . This implies (46) and completes the proof for the first case.

Now suppose that max⁡(p, q, m, n) > 1. Assume, for example, that 0 < p ≤ 1,   q > 1,   0 < m < 1,   n > 1. Then, as in the first case, we introduce the functions w1(x, t), w2(x, t). We use the following relations:

()
where θ1, θ2 are nonnegative continuous functions which are between umax⁡(x, t),   u(x, t) and vmax⁡(x, t),  v(x, t), respectively, in QT. Then we have that a pair of functions (w1(x, t), w2(x, t)) satisfies relations
()

Further develop the arguments, as in the first case, only for q = 1,  n = 1. Using the linearization of terms with powers greater than 1 in the equations and boundary conditions of (1) as above we can prove the theorem for the remaining cases in a similar way.

Conflict of Interests

The authors declare that there is no conflict of interests regarding the publication of this paper.

Acknowledgments

The authors would like to thank the referees for their valuable comments and suggestions regarding the original paper.

      The full text of this article hosted at iucr.org is unavailable due to technical difficulties.