Research Article Positive Steady States of a Strongly Coupled Predator-Prey

advertisement
Hindawi Publishing Corporation
Journal of Applied Mathematics
Volume 2013, Article ID 851028, 8 pages
http://dx.doi.org/10.1155/2013/851028
Research Article
Positive Steady States of a Strongly Coupled Predator-Prey
System with Holling-(𝑛 + 1) Functional Response
Xiao-zhou Feng1,2 and Zhi-guo Wang2
1
2
College of Science, Xi’an Technological University, Xi’an 710032, China
Institute of Mathematics, Shaanxi Normal University, Xi’an 710062, China
Correspondence should be addressed to Xiao-zhou Feng; flxzfxz8@163.com
Received 25 December 2012; Accepted 2 March 2013
Academic Editor: Junjie Wei
Copyright © 2013 X.-z. Feng and Z.-g. Wang. This is an open access article distributed under the Creative Commons Attribution
License, which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly
cited.
This paper discusses a predator-prey system with Holling-(𝑛 + 1) functional response and the fractional type nonlinear diffusion
term in a bounded domain under homogeneous Neumann boundary condition. The existence and nonexistence results concerning
nonconstant positive steady states of the system were obtained. In particular, we prove that the positive constant solution (Μƒ
𝑒, ΜƒV) is
asymptotically stable when the parameter k satisfies some conditions.
1. Introduction
In this paper, we are interested in the positive steady states
of the strongly coupled predator-prey system with Holling(𝑛 + 1) functional response. The specific system is as follows:
−𝑑1 Δ𝑒 = 𝑒 (1 −
−𝑑2 Δ (V +
𝑒
𝑒𝑛 V
)−
π‘˜
π‘Ž + 𝑒𝑛
𝑑3 V
πœŽπ‘’π‘› V
) = −𝑏V +
1+𝑒
π‘Ž + 𝑒𝑛
πœ•π‘’ πœ•V
=
=0
πœ•] πœ•]
in Ω,
in Ω,
(1)
on πœ•Ω,
where 1 ≤ 𝑛 < +∞; Ω is a bounded domain in 𝑅𝑁
with smooth boundary πœ•Ω; πœ•/πœ•] is the outward directional
derivative normal to πœ•Ω; 𝑒 and V stand for the densities of
the prey and predator; the given coefficients 𝑑𝑖 (𝑖 = 1, 2),
π‘Ž, 𝑏, π‘˜, and 𝜎 are positive constants. The term 𝑒𝑛 /(π‘Ž + 𝑒𝑛 )
is named Holling-(𝑛 + 1) functional response [1, 2]. In the
second equation, the fractional type nonlinear diffusion term
Δ𝑑2 (𝑑3 V/(1 + 𝑒)) models a situation in which the population
pressure of the predator species weakens in high-density areas
of the prey species. For more precise details, we can refer
to [3, 4]. Paper [3] discusses a strongly coupled predatorprey system with nonmonotonic functional response, the
existence and nonexistence results concerning nonconstant
positive steady states of the system were proved by degree
theory. Paper [4] considers the positive steady states for a
prey-predator model with some nonlinear diffusion terms,
and the sufficient conditions for the existence of positive
steady state solutions were obtained by bifurcation theory.
In recent years, there has been considerable interest in the
dynamics of strongly coupled reaction-diffusion systems with
cross-diffusion. We point out that most efforts have concentrated on the Lotka-Volterra competition system which was
proposed first by Shigesada et al. [5]. Since their pioneering
work, many authors have studied population models with
cross-diffusion terms from various mathematical viewpoints,
for example, the global existence of time-depending solutions
[6–11], the stability analysis for steady states [12–14], and
the steady state problems [15–21]. In this paper, we mainly
consider the existence of solutions of (1). The research
method refers to [3, 22, 23].
For convenience of the research, we write (1) as the
following form:
−ΔΦ (π‘ˆ) = 𝐺 (π‘ˆ) ,
𝑒
π‘ˆ = [ ],
V
(2)
2
Journal of Applied Mathematics
where
boundary condition. Then, there exists a positive constant
𝐢 = 𝐢(Ω, ‖𝑐‖∞ ) such that maxΩ πœ” ≤ 𝐢minΩ πœ”.
Maximum Principle (see [25]). Suppose that 𝑔 ∈ 𝐢(Ω × π‘…1 ).
If 𝑀 ∈ 𝐢2 (Ω) ∩ 𝐢1 (Ω) satisfies
𝑑1 𝑒
πœ™ (π‘ˆ)
[
]
Φ (π‘ˆ) = [ 1
]=[
𝑑3 ] ,
πœ™2 (π‘ˆ)
𝑑 V (1 +
)
[ 2
1+𝑒 ]
𝐺 (π‘ˆ) = [
𝑔 (𝑒) = 𝑒 (1 −
𝑒
),
π‘˜
Δ𝑀 (π‘₯) + 𝑔 (π‘₯, 𝑀 (π‘₯)) ≥ 0 in Ω,
(3)
πœ•π‘€
≤0
πœ•]
𝑔 (𝑒) − V𝑝 (𝑒)
],
−𝑏V + 𝜎V𝑝 (𝑒)
𝑝 (𝑒) =
𝑒𝑛
,
π‘Ž + 𝑒𝑛
𝑛 ∈ 𝑁+ .
(4)
The main work of this paper is to study the effects of the
fractional type nonlinear diffusion pressures on the existence
of nonconstant positive steady states of (1). Here, a positive
solution means a smooth solution (𝑒, V) with both 𝑒 and V
being positive. We will demonstrate that the cross-diffusion
pressure 𝑑3 may help forming more patterns. Obviously, for
system (1), one notes that when 𝜎 ≤ 𝑏, there holds sup𝑠≥0 {−𝑏+
πœŽπ‘ π‘› /(π‘Ž + 𝑠𝑛 )} ≤ 0, so that the only nonnegative solutions to
(1) are π‘ˆ = (0, 0) and π‘ˆ = (π‘˜, 0). Consequently, (1) does not
have any positive solution. On the other hand, when 𝜎 > 𝑏,
the unique positive constant solution to (1) is (Μƒ
𝑒, ΜƒV); that is,
and 𝑀(π‘₯0 ) = max٠𝑀, then 𝑔(π‘₯0 , 𝑀(π‘₯0 )) ≥ 0.
Theorem 1. Let 𝑃 = (𝜎, π‘Ž, π‘˜, 𝑏) ∈ (0, ∞)4 and 𝑑, 𝑑̂ > 0 be
fixed constants. Assume that 𝑑1 , 𝑑2 ≥ 𝑑 and 0 ≤ 𝑑3 ≤ 𝑑̂
Μ‚ Ω), such that any
there exists a positive constant 𝐢 = 𝐢(𝑃, 𝑑, 𝑑,
positive solution (𝑒, V) of (1) satisfies
max 𝑒 (π‘₯) + max V (π‘₯) +
Ω
Ω
max٠𝑒 (π‘₯) maxΩ V (π‘₯)
+
≤ 𝐢.
min٠𝑒 (π‘₯) minΩ V (π‘₯)
(7)
Proof. Let πœ™1 (π‘ˆ) and πœ™2 (π‘ˆ) be defined as in (3), and denote
that
𝑔1 (π‘ˆ) = 𝑒 (1 −
𝑒̃ = √
𝑛
π‘Žπ‘
,
𝜎−𝑏
(1 − 𝑒̃/π‘˜) (π‘Ž + 𝑒̃𝑛 ) (π‘˜ − 𝑒̃) πœŽΜƒ
𝑒
ΜƒV =
=
,
π‘˜π‘
𝑒̃𝑛−1
𝑔2 (π‘ˆ) = V (
(5)
if π‘˜ > 𝑒̃.
The organization of our paper is as follows. In Section 2,
we establish a priori upper and lower bounds for positive
solutions of (1). In Section 3, we use a degree theory to
develop a general result to enable one to conclude the existence or nonexistence of nonconstant steady-state solutions
or patterns as the index of positive constant steady states
changes. In Section 4, we establish the existence of nonconstant positive solutions to (1) for a large range of diffusion
and cross-diffusion coefficients. Meanwhile, we prove that the
positive constant solution (Μƒ
𝑒, ΜƒV) is asymptotically stable for
different ranges of parameters.
2. Upper and Lower Bounds for Positive
Solutions
The main purpose of this section is to give a priori upper and
lower positive bounds for positive solutions of (1). Firstly, we
cite two known results.
Harnack Inequality (see [24]). Let 𝑐 ∈ 𝐢(Ω) and πœ” ∈
𝐢2 (Ω) β‹‚ 𝐢1 (Ω) be a positive classical solution to Δπœ”(π‘₯) +
𝑐(π‘₯)πœ”(π‘₯) = 0 in Ω subject to the homogeneous Neumann
(6)
on πœ•Ω,
𝑒 𝑒𝑛−1 V
),
−
π‘˜ π‘Ž + 𝑒𝑛
πœŽπ‘’π‘›
− 𝑏) ,
π‘Ž + 𝑒𝑛
(8)
𝑇
𝐺 (π‘ˆ) = (𝑔1 (π‘ˆ) , 𝑔2 (π‘ˆ)) .
Then, (1) becomes
−ΔΦ (π‘ˆ) = 𝐺 (π‘ˆ)
in Ω,
πœ•π‘ˆ
=0
πœ•]
on πœ•Ω.
(9)
For the first equation of (9), by the Maximum Principle,
we have
max 𝑒 (π‘₯) ≤
Ω
1
max πœ™ (π‘ˆ) < π‘˜.
𝑑1 Ω 1
(10)
The function πœ™2 (π‘ˆ) satisfies
Δπœ™2 (π‘ˆ) + 𝑐 (π‘₯) πœ™2 (π‘ˆ) = 0 in Ω,
πœ•πœ™2 (π‘ˆ)
=0
πœ•]
on πœ•Ω,
(11)
where 𝑐(π‘₯) = 𝑔2 (π‘ˆ)/πœ™2 (π‘ˆ) ∈ 𝐢(Ω). It is easy to verify that
the norm
σ΅„©σ΅„© V (πœŽπ‘’π‘› / (π‘Ž + 𝑒𝑛 ) − 𝑏) σ΅„©σ΅„©
πœŽπ‘˜π‘› / (π‘Ž + π‘˜π‘› ) + 𝑏
σ΅„©
σ΅„©σ΅„©
σ΅„©σ΅„© ≤
‖𝑐‖∞ = σ΅„©σ΅„©σ΅„©σ΅„©
𝑑2
σ΅„©σ΅„© 𝑑2 (V + 𝑑3 V/ (1 + 𝑒)) σ΅„©σ΅„©σ΅„©∞
πœŽπ‘˜π‘› + 𝑏 (π‘Ž + π‘˜π‘› )
;
≤
𝑑 (π‘Ž + π‘˜π‘› )
(12)
Journal of Applied Mathematics
3
then, ‖𝑐‖∞ is bounded by a constant depending only on 𝑃 and
𝑑 (𝑑2 ≥ 𝑑). By the Harnack inequality, we have
Proof. Since ∫٠𝑔2 (π‘ˆ)𝑑π‘₯ = 0, there exists π‘₯1 ∈ Ω, such that
𝑔2 (π‘ˆ(π‘₯1 )) = 0; that is,
max πœ™2 (π‘ˆ (π‘₯)) ≤ 𝐢1 min πœ™2 (π‘ˆ (π‘₯)) ,
πœŽπ‘’π‘› (π‘₯1 ) = π‘π‘Ž + 𝑏𝑒𝑛 (π‘₯1 ) .
Ω
Ω
(13)
It follows that max٠𝑒 ≥ √𝑛 a𝑏/𝜎. Consequently, by
max٠𝑒/min٠𝑒 ≤ 𝐢 in Theorem 1, we have
where 𝐢1 = 𝐢1 (Ω, ‖𝑐‖∞ ) = 𝐢1 (𝑃, 𝑑, Ω). It follows that
−1
maxΩ V maxΩ πœ™2 (π‘ˆ (π‘₯)) maxΩ {1 + 𝑑3 (1 + 𝑒) }
×
≤
minΩ V minΩ πœ™2 (π‘ˆ (π‘₯)) minΩ {1 + 𝑑3 (1 + 𝑒)−1 }
Μ‚ .
≤ 𝐢1 (1 + 𝑑3 ) ≤ 𝐢1 (1 + 𝑑)
(14)
Integrating the equations of (1) over Ω by parts and
making use of the boundary conditions, we have
𝜎 ∫ 𝑒 (1 −
Ω
𝑒
) 𝑑π‘₯ = 𝑏 ∫ V 𝑑π‘₯.
π‘˜
Ω
𝜎
𝑒
𝜎
∫ 𝑒 (1 − ) 𝑑π‘₯ < π‘˜.
π‘˜
𝑏
|Ω| 𝑏 Ω
min 𝑒 ≥
Ω
(16)
𝑒𝑖 󳨀→ 𝑒,
−Δ𝑒𝑖 =
Μ‚ 𝜎 π‘˜.
max V ≤ 𝐢1 (1 + 𝑑)
𝑏
Ω
‖𝑐‖∞
σ΅„©
σ΅„©
𝑒 𝑒𝑛−1 V σ΅„©σ΅„©σ΅„©
1 σ΅„©σ΅„©σ΅„©
σ΅„©σ΅„©1 − −
σ΅„©σ΅„©
𝑑1 σ΅„©σ΅„©σ΅„©
π‘˜ π‘Ž + 𝑒𝑛 σ΅„©σ΅„©σ΅„©∞
1
1
≤
[1 + ‖𝑒 (π‘₯)β€–∞ + π‘šβ€–V (π‘₯)β€–∞ ] ,
𝑑1
π‘˜
(18)
max٠𝑒 maxΩ πœ™1 (π‘ˆ (π‘₯))
≤ 𝐢2 .
=
min٠𝑒 minΩ πœ™1 (π‘ˆ (π‘₯))
(19)
Thus, along with (10), (14), and (17), we can complete the
proof.
Theorem 2. Assume that πœŽπ‘˜π‘› =ΜΈ 𝑏(π‘Ž + π‘˜π‘› ). Let 𝑑, 𝑑̂ > 0 be fixed
Μ‚ 𝑃, Ω) > 0, such
constants. There exists a constant 𝐢 = 𝐢(𝑑, 𝑑,
that any positive solution (𝑒, V) of (1) satisfies
min 𝑒 (π‘₯) > 𝐢,
Ω
min V (π‘₯) > 𝐢,
Ω
Μ‚
provided that 𝑑1 , 𝑑2 ≥ 𝑑 and 𝑑3 ≤ 𝑑.
(20)
as 𝑖 󳨀→ ∞.
(23)
𝑒𝑛−1 V
𝑒
1
(1 − 𝑖 − 𝑖 𝑛𝑖 ) 𝑒𝑖
𝑑1,𝑖
π‘˜ π‘Ž + 𝑒𝑖
in Ω,
(24)
on πœ•Ω.
Μ‚ 𝑃, Ω) such that
According to (7), there exists 𝐢 = 𝐢(𝑑, 𝑑,
‖𝑒𝑖 β€–∞ ≤ 𝐢, β€–(1 − 𝑒𝑖 /π‘˜) − (𝑒𝑖𝑛−1 V𝑖 /(π‘Ž + 𝑒𝑖𝑛 ))/𝑑1,𝑖 (1 + 𝑑3,𝑖 V𝑖 )β€–∞ ≤
𝐢, where 𝐢 is a positive constant which does not depend on 𝑖.
For each 𝑖 given in problem (24), it follows from 𝐿𝑝 estimate
that ‖𝑒𝑖 β€–π‘Š2,𝑝 (Ω) ≤ 𝐢2 |Ω|1/𝑝 , where 𝑝 > 1. Let 𝑝 > 𝑁; then, by
Sobolev Imbedding Theorems, we get
σ΅„©σ΅„© σ΅„©σ΅„©
∗σ΅„© σ΅„©
∗ 2
1/𝑝
󡄩󡄩𝑒𝑖 󡄩󡄩𝐢1,𝛼 (Ω) ≤ 𝐢 󡄩󡄩󡄩𝑒𝑖 σ΅„©σ΅„©σ΅„©π‘Š2,𝑝 (Ω) ≤ 𝐢 𝐢 |Ω| ,
(𝑛−1)/𝑛
where π‘š = [(𝑛 − 1)π‘Ž]
/π‘›π‘Ž. Then, ‖𝑐‖∞ is bounded by
a constant depending only on 𝑃 and 𝑑 (𝑑1 ≥ 𝑑). By the
Harnack Inequality, there exists a positive constant 𝐢2 =
Μ‚ Ω) such that max πœ™ (π‘ˆ(π‘₯)) ≤ 𝐢 min πœ™ (π‘ˆ(π‘₯)).
𝐢2 (𝑃, 𝑑, 𝑑,
2
Ω 1
Ω 1
Then,
Ω
πœ•π‘’π‘–
=0
πœ•]
(17)
Similarly, consider the equation of πœ™1 as follows:
minV𝑖 (π‘₯) 󳨀→ 0,
Μ‚ and
By (14), we have maxΩ V𝑖 (π‘₯)/minΩ V𝑖 (π‘₯) ≤ 𝐢1 (1 + 𝑑),
then, maxΩ V𝑖 (π‘₯) → 0; furthermore, V𝑖 (π‘₯) → 0 uniformly
holds as 𝑖 → ∞. Then, the first equation of (1) becomes the
following:
Along with (14), we have
=
√𝑛 π‘Žπ‘/𝜎
1
Μ‚ 𝑃, Ω) , (22)
max 𝑒 ≥
:= 𝐢3 = 𝐢3 (𝑑, 𝑑,
𝐢 Ω
𝐢
where 𝑑3 ≤ 𝑑̂ and 𝑑1 , 𝑑2 ≥ 𝑑.
In the following, we mainly prove that minΩ V(π‘₯) > 𝐢
as πœŽπ‘˜π‘› =ΜΈ 𝑏(π‘Ž + π‘˜π‘› ). Now, we suppose that claim is not true;
then, there exists a sequence (𝑑1,𝑖 , 𝑑2,𝑖 , 𝑑3,𝑖 ) → (𝑑1 , 𝑑2 , 𝑑3 )
Μ‚ And the positive solution
with 𝑑1,𝑖 , 𝑑2,𝑖 ≥ 𝑑 and 𝑑3,𝑖 ≤ 𝑑.
(𝑒𝑖 , V𝑖 ) of (1) corresponding to (𝑑1,𝑖 , 𝑑2,𝑖 , 𝑑3,𝑖 ) is such that
(15)
For any 𝑓 ∈ 𝐿1 (Ω), we denote that 𝑓 = (1/|Ω|) ∫٠𝑓(π‘₯)𝑑π‘₯.
Then, from (15) and (10), we have
V=
(21)
(25)
where 𝛼 ∈ (0, 1). We choose 𝛼 < 1−𝑁/𝑝 such that Imbedding
is compact, and along with elliptic equation regularity theory,
there exists the subsequence of {𝑒𝑖 }, which is still denoted
by {𝑒𝑖 }, and exists 𝑒 such that 𝑒𝑖 → 𝑒 uniformly holds in
𝐢2,𝛼 (Ω). On the other hand, V𝑖 → 0, and when 𝑑1,𝑖 → 𝑑1 ∈
[𝑑, ∞), the limit of (24) becomes the following problem:
−Δ𝑒 =
𝑒
1
𝑒 (1 − )
𝑑1
π‘˜
in Ω,
πœ•π‘’
=0
πœ•]
on πœ•Ω. (26)
Applying Maximum Principle to problem (26) and noting
that min٠𝑒𝑖 ≥ 𝐢3 , we have 𝑒 = π‘˜.
However, when 𝑑1,𝑖 → ∞, the limit of (24) becomes the
following problem:
−Δ𝑒 = 0 in Ω,
πœ•π‘’
=0
πœ•]
on πœ•Ω,
(27)
Μ‚ for some nonnegative constant 𝐢.
Μ‚
which implies that 𝑒 = 𝐢
𝑛−1
𝑛
Since ∫٠𝑒𝑖 (1−𝑒𝑖 /π‘˜−𝑒𝑖 V𝑖 /(π‘Ž+𝑒𝑖 ))𝑑π‘₯ = 0, by letting 𝑖 → ∞
4
Journal of Applied Mathematics
Μ‚ we also
and noting that min٠𝑒𝑖 ≥ 𝐢3 , V𝑖 → 0, and 𝑒𝑖 → 𝐢,
have 𝑒 = π‘˜.
By a similar argument as that in (24), for the second
equation of (1), we can prove that there exists a subsequence in 𝐢2,𝛼 (Ω), such that V𝑖 (π‘₯)/maxΩ V𝑖 (π‘₯) → V0 . Since
Μ‚ then
V𝑖 /maxΩ V𝑖 (π‘₯) ≥ minΩ V𝑖 (π‘₯)/maxΩ V𝑖 (π‘₯) ≥ 1/𝐢(1 + 𝑑),
Μ‚
V0 ≥ 1/𝐢(1 + 𝑑). Dividing the second equation of (1) by
maxΩ V𝑖 (π‘₯), and integrating over Ω, we have
πœŽπ‘’π‘–π‘›
V𝑖
(
− 𝑏) 𝑑π‘₯ = 0.
maxΩ V𝑖 (π‘₯) π‘Ž + 𝑒𝑖𝑛
∫
Ω
(28)
Let 𝑖 → ∞, and note that V𝑖 /maxΩ V𝑖 (π‘₯) → V0 and 𝑒𝑖 → π‘˜;
then, πœŽπ‘˜π‘› /(π‘Ž + π‘˜π‘› ) = 𝑏. This contradiction to the assumption
completes the proof.
3. A Result on Degree Theory
In this section, we obtain nonexistence of nonconstant positive solutions to (1) as 𝑑3 = 0. Meanwhile, by degree theory,
a general result to establish the existence of nonconstant
positive solutions to (1) in the next section is proved.
Denote d = (𝑑1 , 𝑑2 , 𝑑3 ) and 𝑃 = (𝜎, π‘Ž, π‘˜, 𝑏). We will fix
𝑃 ∈ (0, ∞)4 and take d ∈ (0, ∞)2 × [0, ∞) as bifurcation
parameters, the dependence of 𝑃 will often be suppressed.
Define
πœ•π‘ˆ
𝑋 = {π‘ˆ ∈ [𝐢 (Ω)] |
= 0 on πœ•Ω} ,
πœ•]
2
2
𝑋+ = {(𝑒, V) ∈ 𝑋 | 𝑒 > 0, V > 0 on Ω} ,
𝐡 (𝐢) = {(𝑒, V) ∈ 𝑋 |
does not depend on d if 𝐢 > 𝐢0 (d, 𝑃), and it also does not
depend on 𝑃, if 𝑃 changes continuously in (0, ∞)4 without
touching the surface 𝜎 = π‘π‘Ž/π‘˜π‘› + 𝑏.
For the case 𝜎 > π‘π‘Ž/π‘˜π‘› + 𝑏, by degree invariance, we need
only consider a special d; say d = (𝑑, 𝑑, 0) with large 𝑑. For
this we can use the following nonexistence result. To compare
the existence regions of (1) with and without cross-diffusion,
we give a nonexistence result stronger than what is needed
here.
Theorem 3. Let (𝑑2 , 𝑃) ∈ (0, ∞)5 be given. Then, there exists
a positive constant 𝑑̂1 (𝑑2 , 𝑃) such that when d = (𝑑1 , 𝑑2 , 0)
and 𝑑1 ≥ 𝑑̂1 (𝑑2 , 𝑃), (1) does not have any nonconstant positive
solution.
Proof. First, by (7), there exists a positive constant 𝐢 =
𝐢(𝑃, 𝑑̂2 ) such that a classical solution to (1) satisfies 𝑒(π‘₯),
V(π‘₯) ≤ 𝐢 provided that 𝑑2 ≥ 𝑑̂2 . Now, we write 𝑓 as average
of 𝑓 over Ω, where
𝑓 (𝑒, V) = 𝑒 (1 −
𝑒 𝑒𝑛−1 V
),
−
π‘˜ π‘Ž + 𝑒𝑛
𝑔 (𝑒, V) =
πœŽπ‘’π‘› V
.
π‘Ž + 𝑒𝑛
(32)
Multiplying the equation for 𝑒 of (1) by (𝑒−𝑒) and integrating
over Ω by parts, we have
𝑑1 ∫ ∇|𝑒 − 𝑒|2 𝑑π‘₯ = ∫ 𝑓 (𝑒, V) (𝑒 − 𝑒) 𝑑π‘₯
Ω
Ω
= ∫ {𝑓 (𝑒, V) − 𝑓 (𝑒, V)} (𝑒 − 𝑒) 𝑑π‘₯
Ω
(29)
= ∫ {𝑓𝑒 (πœ‰, πœ‚) (𝑒 − 𝑒)2
1
< 𝑒, V < 𝐢 on Ω} .
𝐢
(33)
Ω
+𝑓V (πœ‰, πœ‚) (𝑒 − 𝑒) (V − V) } 𝑑π‘₯
Since
[
Φπ‘ˆ (π‘ˆ) = [
[
𝑑1
𝑑𝑑V
− 2 3 2
[ (1 + 𝑒)
0
]
]
𝑑3
]
]
𝑑2 [1 +
(1 + 𝑒) ]
2
≤ ∫ {(𝐢1 𝑒 − 𝑒) + πœ€(V − V)2 } 𝑑π‘₯
(30)
and det Φπ‘ˆ(π‘ˆ) is positive for all nonnegative π‘ˆ, Φ−1
π‘ˆ exists.
Hence, π‘ˆ is a positive solution to (1) if and only if
=0
𝑑2 ∫ ∇|V − V|2 𝑑π‘₯ = ∫ [−𝑏V + 𝑔 (𝑒, V)] (V − V) 𝑑π‘₯
Ω
𝐹 (d; π‘ˆ) := π‘ˆ − (𝐼 − Δ)−1
× {Φπ‘ˆ(π‘ˆ)−1 [𝐺 (π‘ˆ) + ∇π‘ˆΦπ‘ˆπ‘ˆ∇π‘ˆ] + π‘ˆ}
Ω
for some positive constant 𝐢1 = 𝐢1 (𝑃, 𝑑̂2 ) and πœ€ = πœ€(𝑃, 𝑑̂2 ) β‰ͺ
1, where πœ‰ and πœ‚ lie between 𝑒 and 𝑒, V and V, respectively.
Similarly, we have
(31)
Ω
= ∫ [−𝑏 (V − V) + 𝑔 (𝑒, V) − 𝑔 (𝑒, V)]
Ω
× (V − V) 𝑑π‘₯
in 𝑋+ ,
where (𝐼 − Δ)−1 is the inverse of 𝐼 − Δ in 𝑋. As 𝐹(d; ⋅) is a
compact perturbation of the identity operator 𝐼, the LeraySchauder deg(𝐹(d; ⋅), 0, 𝐡) is well defined if 𝐹(d; π‘ˆ) =ΜΈ 0 for all
π‘ˆ ∈ πœ•π΅.
By Theorems 1 and 2, there exists a positive constant
𝐢0 (d, 𝑃) such that if 𝐹(d; π‘ˆ) = 0 in 𝑋+ , then π‘ˆ ∈ 𝐡(𝐢0 ).
Note that 𝐢0 (d, 𝑃) can be taken as a continuous function
for 𝜎 =ΜΈ π‘π‘Ž/π‘˜π‘› + 𝑏. By the invariance property of the LeraySchauder degree, we then conclude that deg(𝐹(d; ⋅), 0, 𝐡(𝐢))
= ∫ {−𝑏(V − V)2 + 𝑔V (πœ‰, πœ‚) (V − V)2
Ω
+ 𝑔𝑒 (πœ‰, πœ‚) (𝑒 − 𝑒) (V − V) } 𝑑π‘₯
≤ ∫ {(𝜎 − 𝑏) (V − V)2 + πœ€1 (V − V)2
Ω
+𝐢2 (𝑒 − 𝑒)2 } 𝑑π‘₯
(34)
Journal of Applied Mathematics
5
for some positive constant 𝐢2 = 𝐢2 (𝑃, 𝑑̂2 ) and πœ€1 =
πœ€1 (𝑃, dΜ‚2 ) β‰ͺ 1, where πœ‰ and πœ‚ are the same as in (33). Adding
(33) and (34), we obtain
∫ {𝑑1 ∇|𝑒 − 𝑒|2 + 𝑑2 ∇|V − V|2 } 𝑑π‘₯
πœ•π΅. Since for each integer 𝑖 ≥ 1, 𝑋𝑖 is invariant under
π·π‘ˆ(𝐹(d; π‘ˆ∗ )), and πœ† is an eigenvalue of π·π‘ˆπΉ on 𝑋𝑖 if and
only if πœ† is an eigenvalue of following matrix:
𝐼−
Ω
2
2
≤ ∫ {(𝐢1 + 𝐢2 ) (𝑒 − 𝑒) + (𝜎 − 𝑏 + πœ€ + πœ€1 ) (V − V) } 𝑑π‘₯.
Ω
(35)
It follows from the Poincaré Inequality that
πœ‡1 ∫ {𝑑1 (𝑒 − 𝑒)2 + 𝑑2 (V − V)2 } 𝑑π‘₯
Ω
2
1
−1
[Φπ‘ˆ(π‘ˆ∗ ) πΊπ‘ˆ (π‘ˆ∗ ) + 𝐼]
1 + πœ‡π‘–
1
−1
[πœ‡ 𝐼 − Φπ‘ˆ(π‘ˆ∗ ) πΊπ‘ˆ (π‘ˆ∗ )] .
=
1 + πœ‡π‘– 𝑖
Since 𝐻(𝑑, π‘ˆ∗ ; πœ‡) =ΜΈ 0, π·π‘ˆπΉ(d; π‘ˆ∗ ) is invertible. Therefore,
the number of eigenvalues with negative real parts of
π·π‘ˆπΉ(d; π‘ˆ∗ ) on 𝑋𝑖 is odd if and only if 𝐻(d, π‘ˆ∗ ; πœ‡π‘– ) < 0, and
therefore
index (𝐹 (d; ⋅) , π‘ˆ∗ ) = (−1)π‘Ÿ ,
2
π‘Ÿ=
∑
1.
𝑖≥1, 𝐻(d,πœ‡π‘– )<0
≤ ∫ {(𝐢1 + 𝐢2 ) (𝑒 − 𝑒) + (𝜎 − 𝑏 + πœ€ + πœ€1 ) (V − V) } 𝑑π‘₯.
Ω
(40)
(41)
(36)
Since we can choose 𝑑2 ≥ 𝑑̂2 such that πœ‡1 𝑑2 > 𝜎 − 𝑏, we
may also choose πœ€ and πœ€1 sufficiently small such that πœ‡1 𝑑2 >
𝜎 − 𝑏 + πœ€ + πœ€1 . Consequently, by (36),
πœ‡1 𝑑1 ∫ (𝑒 − 𝑒)2 𝑑π‘₯ ≤ (𝐢1 + 𝐢2 ) ∫ (𝑒 − 𝑒)2 𝑑π‘₯,
Ω
Ω
(37)
which implies that 𝑒 = 𝑒 = constant, and, hence, V = V =
constant if 𝑑1 ≥ 𝑑̂1 (𝑑2 , 𝑃) := (𝐢1 (𝑃, 𝑑̂2 ) + 𝐢2 (𝑃, 𝑑̂2 ))/πœ‡1 . Thus,
we complete the proof of the theorem.
In the following, we only calculate deg(𝐹, 0, 𝐡) when all
solutions to 𝐹 = 0 are positive constant solutions in 𝐡(𝐢).
Let 0 = πœ‡1 < πœ‡2 < πœ‡3 < ⋅ ⋅ ⋅ be the eigenvalues of
the operator −Δ in Ω with zero flux boundary condition,
𝐸(πœ‡π‘– ) be the eigenspace corresponding to πœ‡π‘– in 𝐢1 (Ω), πœ™π‘–π‘— , 𝑗 =
1, 2, 3, . . ., dim 𝐸(πœ‡π‘– ) an orthonormal basis of 𝐸(πœ‡π‘– ), and 𝑋𝑖𝑗 =
dim 𝐸(πœ‡ )
𝑖
𝑋𝑖𝑗 .
{πΆπœ™π‘–π‘— | 𝐢 ∈ 𝑅2 }. Then, 𝑋 = ⊕∞
𝑖=1 𝑋𝑖 , where 𝑋𝑖 = ⊕𝑗=1
This decomposed method is similar to that of [22].
Let π‘ˆ∗ be a positive root to 𝐺(π‘ˆ) = 0. We can calculate
π·π‘ˆπΉ (d; π‘ˆ∗ ) = 𝐼 − (𝐼 − Δ)−1
∗
∗
× {Φ−1
π‘ˆ (π‘ˆ ) πΊπ‘ˆ (π‘ˆ ) + 𝐼}
in 𝐿 (𝑋, 𝑋) ,
(38)
where 𝐿(𝑋, 𝑋) is a linear mapping from 𝑋 to itself.
Denote that 𝐻(d, π‘ˆ∗ ; πœ‡) := det[πœ‡πΌ − Φπ‘ˆ(π‘ˆ∗ )−1 πΊπ‘ˆ(π‘ˆ∗ )].
∗
Lemma 4. Let π‘ˆ be a positive root to 𝐺(⋅) = 0, and assume
that 𝐻(d, π‘ˆ∗ ; πœ‡π‘– ) =ΜΈ 0 for all 𝑖. Then,
index (𝐹 (d; ⋅) , π‘ˆ∗ ) = (−1)π‘Ÿ ,
π‘Ÿ=
∑
𝑖≥1, 𝐻(d,πœ‡π‘– )<0
1.
(39)
Proof. If π·π‘ˆπΉ is invertible, then the index of 𝐹 at π‘ˆ∗
is defined as index(𝐹(d; ⋅), π‘ˆ∗ ) = (−1)π‘Ÿ , where π‘Ÿ is the
number of eigenvalues of π·π‘ˆπΉ with negative real parts. The
deg(𝐹(d; ⋅), 0, 𝐡) is then equal to summation of the indexes
over all solutions to 𝐹 = 0 in 𝐡, provided that 𝐹 ≠ 0 on
Theorem 5. Assume that 𝑃 = (𝜎, π‘Ž, π‘˜, 𝑏) ∈ (0, ∞)4 with
πœŽπ‘˜π‘› =ΜΈ 𝑏(π‘Ž + π‘˜π‘› ). Then, for every d ∈ (0, ∞)2 × [0, ∞), there
exists a constant 𝐢0 (d, 𝑃) such that for every 𝐢 > 𝐢0 (d, 𝑃),
{
0,
{
{
deg (𝐹 (d; ⋅) , 0, 𝐡 (𝐢)) = {
{
{
1,
{
π‘π‘Ž
+ 𝑏,
π‘˜π‘›
π‘π‘Ž
if 𝜎 < 𝑛 + 𝑏.
π‘˜
if 𝜎 <
(42)
Proof. Since (1) does not have any positive solution in 𝐡(𝐢),
when 𝜎 ≤ 𝑏, we then conclude that
deg (𝐹 (d; ⋅) , 0, 𝐡 (𝐢)) = 0,
∀𝑃 ∈ {(𝜎, π‘Ž, π‘˜, 𝑏) ∈ (0, ∞)4 | 𝜎 <
π‘π‘Ž
+ 𝑏} ,
π‘˜π‘›
(43)
𝐢 > 𝐢0 (d, 𝑃) .
Next, to complete the proof of Theorem 5. We need only
to calculate the degree for the case 𝜎 > π‘π‘Ž/π‘˜π‘› + 𝑏. In this
case, by Theorem 3, all positive solutions to 𝐹 = 0 are the
unique positive constant solution to 𝐺(⋅) = 0, denoted by
Μƒ = 𝑑𝐼 so that 𝐻(d, π‘ˆ;
Μƒ πœ‡) =
Μƒ when d = (𝑑, 𝑑, 0), Φπ‘ˆ(π‘ˆ)
π‘ˆ,
2
Μƒ
Μƒ
(1/𝑑) det(π‘‘πœ‡πΌ − πΊπ‘ˆ(π‘ˆ)), det(πΊπ‘ˆ(π‘ˆ)) = 𝑛𝑏 (𝜎 − 𝑏)ΜƒV/𝜎2 𝑒̃ > 0.
Μƒ πœ‡π‘– ) > 0
This implies that when 𝑑 is sufficiently large, 𝐻(d, π‘ˆ;
for all 𝑖 = 1, 2, 3, . . ., and thus π‘Ÿ = 0. It follows from Lemma 4
that deg(𝐹(d; ⋅), 0, 𝐡(𝐢)) = 1. This completes the proof.
Remark 6. The change of degree when 𝜎 passes the borderline
𝜎 = π‘π‘Ž/π‘˜π‘› + 𝑏 is due to the appearance (disappearance) of a
positive constant steady-state bifurcating from π‘ˆ ≡ (π‘˜, 0).
4. Existence of Nonconstant Positive Solutions
In this section, we establish the existence of positive nonconstant solutions for (1). In particular, we show that for
certain ranges of parameters where (1) does not have any
positive nonconstant steady state, our model can still produce
patterns. The idea is as follows. First we calculate the index
of 𝐹(d; ⋅) at positive constant steady states. Suppose that the
6
Journal of Applied Mathematics
sum of all these indices is not equal to the degree stated in
Theorem 5. Then, 𝐹(d; ⋅) = 0 in 𝐡(𝐢) for 𝐢 = 𝐢0 (d, 𝑃) must
have a nonconstant positive solution, which also solves (1).
In the following, we always assume that 𝜎 > 𝑏, and (1) has
Μƒ = (Μƒ
a unique positive constant solution (Μƒ
𝑒, ΜƒV), when π‘ˆ
𝑒, ΜƒV)
is a positive solution to 𝐺(π‘ˆ) = 0. We can get the following
Μƒ by simply calculating
results at π‘ˆ
1
𝑏
− 𝑏) − 2𝜎) 𝑒̃ − π‘˜ [𝑛 (𝜎 − 𝑏) − 𝜎]} −
[ πœŽπ‘˜ {(𝑛 (𝜎
𝜎]
]
Μƒ =[
πΊπ‘ˆ (π‘ˆ)
]
[
𝑛𝑏 (𝜎 − 𝑏) ΜƒV
0
]
[
πœŽΜƒ
𝑒
4.1. The Case 𝑛(𝜎−𝑏)−2𝜎>0 (𝑛>2𝜎/(𝜎−𝑏)).
In this subsection, we consider local stability of the constant steady state
Μƒ for evolution dynamics
π‘ˆ≡π‘ˆ
π‘ˆπ‘‘ = ΔΦ (π‘ˆ) + 𝐺 (π‘ˆ)
πœ•π‘ˆ
=0
πœ•]
0
𝑑1
]
[
]
[
Μƒ
Φπ‘ˆ (π‘ˆ) = [ 𝑑 𝑑 ΜƒV
𝑑3
]
2 3
]
𝑑
[1
+
−
2
2
(1 + 𝑒̃) ]
[ (1 + 𝑒̃)
Μƒ = (Μƒ
where π‘ˆ
𝑒, ΜƒV) is a positive constant solution to 𝐺(⋅) = 0 by
(5).
Μƒ takes the form
Proof. The linearization of (47) at π‘ˆ
πœ™
0
=: [πœ™11 πœ™ ] ,
21
22
Μƒ Δ𝑉 + πΊπ‘ˆ (π‘ˆ)
Μƒ 𝑉
𝑉𝑑 = Φπ‘ˆ (π‘ˆ)
Μƒ πΊπ‘ˆ (π‘ˆ))
Μƒ
Μƒ πœ‡) ≡ det (πœ‡πΌ − Φ−1 (π‘ˆ)
𝐻 (d, π‘ˆ;
π‘ˆ
πœ•π‘‰
=0
πœ•]
(44)
Μƒ ⋅) = 0, we will restrict
To calculate the roots of 𝐻(d, π‘ˆ;
our attention to large |d|. Note that
Μƒ πœ‡) = πœ‡2 − Λ π‘– πœ‡,
lim 𝐻 (d, π‘ˆ;
𝑑𝑖 → ∞
∀𝑖 = 2, 3,
(45)
where
Λ 2 = Λ 2 (𝑑1 , 𝑑3 ; π‘ˆ) = (−𝑑3 π‘π‘˜ΜƒV + (1 + 𝑒̃ + 𝑑3 ) (1 + 𝑒̃)
× {(𝑛 (𝜎 − 𝑏) − 2𝜎) 𝑒̃ − π‘˜ (𝑛 (𝜎 − 𝑏) − 𝜎)} )
−1
× (πœŽπ‘˜π‘‘1 (1 + 𝑒̃ + 𝑑3 ) (1 + 𝑒̃)) ,
Λ 3 = Λ 3 (𝑑1 ; π‘ˆ)
=
(47)
on πœ•Ω,
Theorem 7. Let π‘˜ > 𝑒̃ > ((𝑛(𝜎 − 𝑏) − 2𝜎)/(𝑛(𝜎 − 𝑏) −
𝜎))Μƒ
𝑒 (𝑛 > 2𝜎/(𝜎 − 𝑏)). Then the positive constant solution
π‘ˆ(π‘₯, 𝑑) ≡ (Μƒ
𝑒, ΜƒV) is asymptotically stable with respect to the
Μƒ
dynamics (47). Consequently, in a small neighborhood of π‘ˆ,
(1) does not have any nonconstant positive solution.
𝑏
=: [𝑔11 − 𝜎 ] ,
𝑔21 0
󡄨󡄨 󡄨󡄨
σ΅„¨πœ™ 󡄨 𝑏/𝜎 − πœ™22 𝑔11 𝑔21 𝑏/𝜎
= πœ‡2 + πœ‡ 󡄨 21 󡄨
+
.
πœ™11 πœ™22
πœ™11 πœ™22
in Ω × (0, ∞) ,
−π‘π‘˜ΜƒV + {(𝑛 (𝜎 − 𝑏) − 2𝜎) 𝑒̃ − π‘˜ (𝑛 (𝜎 − 𝑏) − 𝜎)} (1 + 𝑒̃)
.
πœŽπ‘˜π‘‘1 (1 + 𝑒̃)
(46)
The sign of the trace tr(πΊπ‘ˆ) = 𝑔11 is determined by 𝛾 =
(𝑛(𝜎−𝑏)−2𝜎)Μƒ
𝑒−π‘˜(𝑛(𝜎−𝑏)−𝜎) and the sign of the determinant
det(πΊπ‘ˆ) = 𝑛𝑏2 (𝜎 − 𝑏)V/𝜎2 𝑒 > 0.
Hence, we will discuss separately the following cases:
(i) 𝑛(𝜎 − 𝑏) − 2𝜎 > 0(𝑛 > 2𝜎/(𝜎 − 𝑏)), obviously π‘˜ > 𝑒̃ >
((𝑛(𝜎 − 𝑏) − 2𝜎)/(𝑛(𝜎 − 𝑏) − 𝜎))Μƒ
𝑒, then 𝛾 < 0;
(ii) 𝑛(𝜎 − 𝑏) − 𝜎 < 0(𝑛 < 𝜎/(𝜎 − 𝑏)):
(iia) if 𝑒̃ < π‘˜ < ((2𝜎 − 𝑛(𝜎 − 𝑏))/(𝜎 − 𝑛(𝜎 − 𝑏)))Μƒ
𝑒, then
𝛾 < 0;
(iib) if π‘˜ > ((2𝜎 − 𝑛(𝜎 − 𝑏))/(𝜎 − 𝑛(𝜎 − 𝑏)))Μƒ
𝑒, then
𝛾 > 0.
in Ω × (0, ∞) ,
(48)
on πœ•Ω.
Μƒ
Denote that the corresponding linear operator 𝐿 := Φπ‘ˆ(π‘ˆ)Δ+
Μƒ
πΊπ‘ˆ(π‘ˆ). For each eigen-pair (πœ‡π‘– , πœ™π‘– ) (𝑖 = 1, 2, 3, . . .) of −Δ on
Ω with no flux boundary condition, 𝑉 = πΆπ‘’πœ†π‘‘ πœ™ is a solution
to (48) if and only if (πœ†, 𝐢) is an eigen-pair of the matrix 𝐴 𝑖 :=
Μƒ
Μƒ
−πœ‡π‘– Φπ‘ˆ(π‘ˆ)+𝐺
π‘ˆ (π‘ˆ). 𝑋𝑖 is invariant under the operator 𝐿. From
(44), we have
𝑔11 < 0
(π‘˜ >
𝑛 (𝜎 − 𝑏) − 2𝜎
𝑒̃) ,
𝑛 (𝜎 − 𝑏) − 𝜎
𝑔21 > 0
(𝜎 > 𝑏) .
(49)
Then
𝑏
𝑏
det 𝐴 𝑖 := πœ™11 πœ™22 πœ‡π‘–2 − ( πœ™21 + 𝑔11 πœ™22 ) πœ‡π‘– + 𝑔21 > 0,
𝜎
𝜎
tr 𝐴 𝑖 := − (πœ™11 + πœ™22 ) πœ‡π‘– + 𝑔11 < 0.
(50)
Hence, we conclude that the two eigenvalues πœ†+𝑖 and πœ†−𝑖 of
𝐴 𝑖 have negative real parts. More specifically, we have the
following:
2
(i) For 𝑖 = 1, πœ‡1 = 0, if 𝑔11
− 4𝑔21 𝑏/𝜎 ≤ 0, then Re πœ†±π‘– =
2
(1/2)𝑔11 < 0; if 𝑔11 − 4𝑔21 𝑏/𝜎 > 0, then
Re πœ†+𝑖 =
1[
2 − 4𝑔21 𝑏 ] ≤ 0,
𝑔11 + √ 𝑔11
2
𝜎
[
]
1
2 − 4𝑔21 𝑏 ] < 0.
Re πœ†−𝑖 = [𝑔11 − √ 𝑔11
2
𝜎
[
]
(51)
Journal of Applied Mathematics
7
(ii) For 𝑖 ≥ 2, as πœ‡π‘– is increasing with respect to 𝑖 and πœ‡π‘– →
∞ as 𝑖 → ∞, it follows that if (tr 𝐴 𝑖 )2 − 4 det 𝐴 𝑖 ≤ 0, then
Re πœ†±π‘– =
1
1
tr 𝐴 𝑖 = [− (πœ™11 + πœ™22 ) πœ‡π‘– + 𝑔11 ]
2
2
1
≤ [− (πœ™11 + πœ™22 ) πœ‡2 + 𝑔11 ] < 0;
2
Μƒ πœ‡− (d, π‘ˆ)
Μƒ = [det (Φπ‘ˆ (π‘ˆ))]
Μƒ −1 det (πΊπ‘ˆ (π‘ˆ))
Μƒ
πœ‡+ (d, π‘ˆ)
(52)
=
if (tr 𝐴 𝑖 )2 − 4 det 𝐴 𝑖 > 0, since det 𝐴 𝑖 > 0 and tr 𝐴 𝑖 < 0, then
Re πœ†−𝑖 =
≤
=
𝑑2 → ∞
2
tr 𝐴 𝑖 − √(tr 𝐴 𝑖 ) − 4 det 𝐴 𝑖
≤
𝑑3 → ∞
(53)
det 𝐴 𝑖
< −Μƒπœ€.
tr 𝐴 𝑖
for some positive constant πœ€Μƒ that does not depend on 𝑖.
The previous arguments show that there exists a positive
constant πœ€, which does not depend on 𝑖, such that Re πœ† 𝑖 <
−πœ€, ∀𝑖. Consequently, the spectrum of 𝐿, which consists of
eigenvalues, lies in {Re(πœ†) < −πœ€}. It then follows from
Theorem 5.1.1 of [26, page 98] that the constant steady-state
Μƒ is asymptotically stable to (47).
π‘ˆ(π‘₯, 𝑑) ≡ π‘ˆ
4.2. The Case 𝑛(𝜎−𝑏)−𝜎<0 (𝑛<𝜎/(𝜎−𝑏))
4.2.1. (𝛾<0)Μƒ
𝑒<π‘˜<((2𝜎−𝑛(𝜎−𝑏))/(𝜎−𝑛(𝜎−𝑏)))Μƒ
𝑒
Theorem 8. Let 𝑒̃ < π‘˜ < ((2𝜎−𝑛(𝜎−𝑏))/(𝜎−𝑛(𝜎−𝑏)))Μƒ
𝑒. Then
the positive constant solution π‘ˆ(π‘₯, 𝑑) ≡ (Μƒ
𝑒, ΜƒV) is asymptotically
stable with respect to the dynamics (47).
Theorem 9 (existence with suitable 𝑑2 and 𝑑3 ). Assume that
𝑒, define
𝜎 > π‘π‘Ž/π‘˜π‘› + 𝑏 and π‘˜ > ((2𝜎 − 𝑛(𝜎 − 𝑏))/(𝜎 − 𝑛(𝜎 − 𝑏)))Μƒ
Λ 2 (𝑑1 , 𝑑3 ; π‘ˆ) and Λ 3 (𝑑1 ; π‘ˆ) as (46).
(i) Suppose that 𝑑1 and 𝑑3 are given such that
Μƒ ∈ (πœ‡π‘— , πœ‡π‘—+1 ) for some positive even integer
Λ 2 (𝑑1 , 𝑑3 ; π‘ˆ)
𝑗. There exists a positive constant 𝐷2 such that if 𝑑2 ≥ 𝐷2 , then
(1) has at least one nonconstant positive solution.
(ii) Suppose that 𝑑1 is given such that Λ 3 (𝑑1 ; π‘ˆ) ∈
(πœ‡π‘— , πœ‡π‘—+1 ) for some positive even integer 𝑗. Then, for any given
𝑑2 > 0, there exists a positive constant 𝐷3 such that if 𝑑3 ≥ 𝐷3 ,
(1) has at least one nonconstant positive solution.
Μƒ = Λ 2 (𝑑1 , 𝑑3 ; π‘ˆ)
Μƒ ,
lim πœ‡+ (d, π‘ˆ)
Μƒ = Λ 3 (𝑑1 ; π‘ˆ)
Μƒ .
lim πœ‡+ (d, π‘ˆ)
𝑑3 → ∞
(55)
Μƒ ∈ (πœ‡π‘— , πœ‡π‘—+1 ) for some positive even
Suppose that Λ 2 (𝑑1 , 𝑑3 ; π‘ˆ)
integer 𝑗; then, there exists a positive constant 𝐷2 ≫ 1 such
that 𝑑2 ≫ 𝐷2 , and we have
Μƒ < πœ‡2 ,
0 = πœ‡1 < πœ‡− (d, π‘ˆ)
Μƒ ∈ (πœ‡π‘— , πœ‡π‘—+1 ) . (56)
πœ‡+ (d, π‘ˆ)
Μƒ πœ‡π‘– ) < 0 is equivalent to 𝑖 ∈
Hence, if 𝑑2 ≫ 1, 𝐻(d, π‘ˆ;
{2, 3, . . . , 𝑗}, since 𝑗 is even. It follows from Lemma 4 that
Μƒ = (−1)𝑗−1 = −1.
index (𝐹 ( d ; ⋅) , π‘ˆ)
(57)
Consequently, 𝐹(d; π‘ˆ) = 0 has at least one nonconstant
positive solution that is different from the constant function
Μƒ Otherwise, the degree of 𝐹 = 0 in 𝐡(𝐢) would be −1
π‘ˆ = π‘ˆ.
for all large enough 𝐢, which would contradict Theorem 5.
This proves, the first assertion of the theorem, and the second
assertion is similarly proved.
Remark 10. For Λ 3 (𝑑1 ; π‘ˆ) to be positive, it is necessary and
sufficient to have
π‘˜>
Proof. Similar to the proof of Theorem 7.
4.2.2. (𝛾>0) π‘˜>((2𝜎−𝑛(𝜎−𝑏))/(𝜎−𝑛(𝜎−𝑏)))Μƒ
𝑒. In this case,
Μƒ
Μƒ
π‘ˆ = (Μƒ
𝑒, V) is the only positive constant solution to 𝐺(⋅) = 0.
By fixing the diffusion coefficients 𝑑1 (for prey) and using the
diffusion coefficients 𝑑2 and 𝑑3 (for predator) as bifurcation
parameters, we will show that (1) can create nonconstant
positive solutions. We want to emphasize that it is caused by
the presence of cross-diffusion which has a more complex role
than that of the diffusion coefficients 𝑑1 and 𝑑2 .
(54)
𝑑2 → ∞
Μƒ = 0,
lim πœ‡− (d, π‘ˆ)
1
1
tr 𝐴 𝑖 ≤ [− (πœ™11 + πœ™22 ) πœ‡2 + 𝑔11 ] < 0;
2
2
2 det 𝐴 𝑖
𝑔21 𝑏/𝜎
> 0.
πœ™11 πœ™22
From (45), we see that
Μƒ = 0,
lim πœ‡− (d, π‘ˆ)
1
2
(tr 𝐴 𝑖 − √(tr 𝐴 𝑖 ) − 4 det 𝐴 𝑖 )
2
1
2
Re πœ†+𝑖 = (tr 𝐴 𝑖 + √(tr 𝐴 𝑖 ) − 4 det 𝐴 𝑖 )
2
Μƒ with Re(πœ‡− ) ≤ Re(πœ‡+ ), the two
Proof. Denote by πœ‡± (d, π‘ˆ),
Μƒ
roots to 𝐻(d, π‘ˆ; πœ‡) = 0, and then
𝑒
(π‘˜ + 1) πœŽΜƒ
+ 𝑒̃.
(1 + 𝑒̃) [𝜎 − 𝑛 (𝜎 − 𝑏)]
(58)
Μƒ is also positive
When this inequality holds, Λ 2 (𝑑1 , 𝑑3 ; π‘ˆ)
provided that 𝑑3 is large, and we then can adjust to 𝑑1 and
make the assumptions in (i) or (ii) of theorem hold.
Remark 11. In fact, if 𝜎/(𝜎 − 𝑏) < 𝑛 < 2𝜎/(𝜎 − 𝑏), then 𝛾 <
0. This case is similar to case (i). Furthermore, the positive
constant solution U(π‘₯, 𝑑) ≡ (Μƒ
𝑒, ΜƒV) is also asymptotically stable
with respect to the dynamics (47).
Acknowledgments
This project is supported by the NSF of China under Grant
11001160, the Scientific Research Plan Projects of Shaanxi
Education Department (no. 12JK0865), and the President
Fund of Xi’an Technological University (XAGDXJJ1136).
References
[1] W. Wang and J. H. Sun, “On the predator-prey system with
Holling-(𝑛 + 1) functional response,” Acta Mathematica Sinica,
vol. 23, no. 1, pp. 1–6, 2007.
8
[2] J. A. Johnson, “Banach spaces of Lipschitz functions and vectorvalued Lipschitz functions,” Transactions of the American Mathematical Society, vol. 148, pp. 147–169, 1970.
[3] X. F. Chen, Y. W. Qi, and M. X. Wang, “A strongly coupled predator-prey system with non-monotonic functional
response,” Nonlinear Analysis. Theory, Methods & Applications,
vol. 67, no. 6, pp. 1966–1979, 2007.
[4] T. Kadota and K. Kuto, “Positive steady states for a preypredator model with some nonlinear diffusion terms,” Journal
of Mathematical Analysis and Applications, vol. 323, no. 2, pp.
1387–1401, 2006.
[5] N. Shigesada, K. Kawasaki, and E. Teramoto, “Spatial segregation of interacting species,” Journal of Theoretical Biology, vol.
79, no. 1, pp. 83–99, 1979.
[6] H. Amann, “Dynamic theory of quasilinear parabolic systems.
III. Global existence,” Mathematische Zeitschrift, vol. 202, no. 2,
pp. 219–250, 1989.
[7] Y. S. Choi, R. Lui, and Y. Yamada, “Existence of global solutions
for the Shigesada-Kawasaki-Teramoto model with strongly
coupled cross-diffusion,” Discrete and Continuous Dynamical
Systems A, vol. 10, no. 3, pp. 719–730, 2004.
[8] L. Dung, “Cross diffusion systems on n spatial dimensional
domains,” Indiana University Mathematics Journal, vol. 51, no.
3, pp. 625–643, 2002.
[9] L. Dung, L. V. Nguyen, and T. T. Nguyen, “Shigesada-KawasakiTeramoto model on higher dimensional domains,” Electronic
Journal of Differential Equations, vol. 2003, pp. 1–12, 2003.
[10] Y. Lou, W.-M. Ni, and Y. Wu, “On the global existence of a crossdiffusion system,” Discrete and Continuous Dynamical Systems,
vol. 4, no. 2, pp. 193–203, 1998.
[11] A. Yagi, “Global solution to some quasilinear parabolic system
in population dynamics,” Nonlinear Analysis. Theory, Methods
& Applications, vol. 21, no. 8, pp. 603–630, 1993.
[12] Y. Kan-on, “Stability of singularly perturbed solutions to
nonlinear diffusion systems arising in population dynamics,”
Hiroshima Mathematical Journal, vol. 23, no. 3, pp. 509–536,
1993.
[13] K. Kuto, “Stability of steady-state solutions to a prey-predator
system with cross-diffusion,” Journal of Differential Equations,
vol. 197, no. 2, pp. 293–314, 2004.
[14] Y. Wu, “The instability of spiky steady states for a competing
species model with cross diffusion,” Journal of Differential
Equations, vol. 213, no. 2, pp. 289–340, 2005.
[15] K. Kuto and Y. Yamada, “Multiple coexistence states for a preypredator system with cross-diffusion,” Journal of Differential
Equations, vol. 197, no. 2, pp. 315–348, 2004.
[16] Y. Lou and W.-M. Ni, “Diffusion vs cross-diffusion: an elliptic
approach,” Journal of Differential Equations, vol. 154, no. 1, pp.
157–190, 1999.
[17] Y. Lou, W.-M. Ni, and S. Yotsutani, “On a limiting system in the
Lotka-Volterra competition with cross-diffusion,” Discrete and
Continuous Dynamical Systems A, vol. 10, no. 1-2, pp. 435–458,
2004.
[18] M. Mimura, Y. Nishiura, A. Tesei, and T. Tsujikawa, “Coexistence problem for two competing species models with densitydependent diffusion,” Hiroshima Mathematical Journal, vol. 14,
no. 2, pp. 425–449, 1984.
[19] K. Nakashima and Y. Yamada, “Positive steady states for preypredator models with cross-diffusion,” Advances in Differential
Equations, vol. 1, no. 6, pp. 1099–1122, 1996.
Journal of Applied Mathematics
[20] W. H. Ruan, “Positive steady-state solutions of a competing
reaction-diffusion system with large cross-diffusion coefficients,” Journal of Mathematical Analysis and Applications, vol.
197, no. 2, pp. 558–578, 1996.
[21] K. Ryu and I. Ahn, “Positive steady-states for two interacting
species models with linear self-cross diffusions,” Discrete and
Continuous Dynamical Systems A, vol. 9, no. 4, pp. 1049–1061,
2003.
[22] P. Y. H. Pang and M. X. Wang, “Strategy and stationary pattern
in a three-species predator-prey model,” Journal of Differential
Equations, vol. 200, no. 2, pp. 245–273, 2004.
[23] P. Y. H. Pang and M. Wang, “Qualitative analysis of a ratiodependent predator-prey system with diffusion,” Proceedings of
the Royal Society of Edinburgh A, vol. 133, no. 4, pp. 919–942,
2003.
[24] C.-S. Lin, W.-M. Ni, and I. Takagi, “Large amplitude stationary
solutions to a chemotaxis system,” Journal of Differential Equations, vol. 72, no. 1, pp. 1–27, 1988.
[25] Y. Lou and W.-M. Ni, “Diffusion, self-diffusion and crossdiffusion,” Journal of Differential Equations, vol. 131, no. 1, pp.
79–131, 1996.
[26] D. Henry, Geometric Theory of Semilinear Parabolic Equations,
vol. 840 of Lecture Notes in Mathematics, Springer, Berlin,
Germany, 3rd edition, 1981.
Advances in
Operations Research
Hindawi Publishing Corporation
http://www.hindawi.com
Volume 2014
Advances in
Decision Sciences
Hindawi Publishing Corporation
http://www.hindawi.com
Volume 2014
Mathematical Problems
in Engineering
Hindawi Publishing Corporation
http://www.hindawi.com
Volume 2014
Journal of
Algebra
Hindawi Publishing Corporation
http://www.hindawi.com
Probability and Statistics
Volume 2014
The Scientific
World Journal
Hindawi Publishing Corporation
http://www.hindawi.com
Hindawi Publishing Corporation
http://www.hindawi.com
Volume 2014
International Journal of
Differential Equations
Hindawi Publishing Corporation
http://www.hindawi.com
Volume 2014
Volume 2014
Submit your manuscripts at
http://www.hindawi.com
International Journal of
Advances in
Combinatorics
Hindawi Publishing Corporation
http://www.hindawi.com
Mathematical Physics
Hindawi Publishing Corporation
http://www.hindawi.com
Volume 2014
Journal of
Complex Analysis
Hindawi Publishing Corporation
http://www.hindawi.com
Volume 2014
International
Journal of
Mathematics and
Mathematical
Sciences
Journal of
Hindawi Publishing Corporation
http://www.hindawi.com
Stochastic Analysis
Abstract and
Applied Analysis
Hindawi Publishing Corporation
http://www.hindawi.com
Hindawi Publishing Corporation
http://www.hindawi.com
International Journal of
Mathematics
Volume 2014
Volume 2014
Discrete Dynamics in
Nature and Society
Volume 2014
Volume 2014
Journal of
Journal of
Discrete Mathematics
Journal of
Volume 2014
Hindawi Publishing Corporation
http://www.hindawi.com
Applied Mathematics
Journal of
Function Spaces
Hindawi Publishing Corporation
http://www.hindawi.com
Volume 2014
Hindawi Publishing Corporation
http://www.hindawi.com
Volume 2014
Hindawi Publishing Corporation
http://www.hindawi.com
Volume 2014
Optimization
Hindawi Publishing Corporation
http://www.hindawi.com
Volume 2014
Hindawi Publishing Corporation
http://www.hindawi.com
Volume 2014
Download