Bgujftyfurcation 3
Bgujftyfurcation 3
Abstract We study the competition of two species for a single resource in a chemostat.
In the simplest space-homogeneous situation, it is known that only one species survives,
namely the best competitor. In order to exhibit coexistence phenomena, where the two
competitors are able to survive, we consider a space dependent situation: we assume
that the two species and the resource follow a diffusion process in space, on top of
the competition process. Besides, and in order to consider the most general case, we
assume each population is associated with a distinct diffusion constant. This is a key
difficulty in our analysis: the specific (and classical) case where all diffusion constants
are equal, leads to a particular conservation law, which in turn allows to eliminate
the resource in the equations, a fact that considerably simplifies the analysis and the
qualitative phenomena.
Using the global bifurcation theory, we prove that the underlying 2-species, sta-
tionary, diffusive, chemostat-like model, does possess coexistence solutions, where both
species survive. On top of that, we identify the domain, in the space of the identified
bifurcation parameters, for which the system does have coexistence solutions.
Keywords Global bifurcation · Elliptic systems · Heterogeneous environment ·
Coexistence · Chemostat
Mathematics Subject Classification (2000) 35Q92 · 35K58 · 92D25 · 92D30
1 Introduction
The present paper is devoted to the study of coexistence solutions in some chemostat-
like systems, where various species compete for a single resource. The starting point
Sten Madec
Institut de Mathematiques de Bordeaux Universite Victor Segalen Bordeaux 2, 3ter place de
la victoire, 33000 Bordeaux Cedex, France
E-mail: sten.madec@univ-rennes1.fr
François Castella
Université de Rennes 1, UMR CNRS 6625 Irmar, Campus de Beaulieu, 35042 Rennes cedex,
France
E-mail: francois.castella@univ-rennes1.fr
2
of our analysis is the fact that in the simplest models, i.e. in the space-homogeneous
situation, only one species survives, namely the best competitor. Therefore, and in
order to observe situations where all species are able to survive, we readily consider
the space-inhomogeneous situation, where the various species and the single resource
follow a diffusion process in space. Technically speaking, and in order to tackle the most
general situation, we assume that each population possesses its own distinct diffusion
coefficient. This is a major difficulty and originality in the present text, as we discuss
later in this introduction.
The main result of this paper is that the underlying 2-species chemostat-like model,
does possess coexistence solutions, i.e. solutions where all species survive. Besides,
we are able to identify a domain in the space of the relevant parameters, for which
coexistence holds.
Our construction relies on global bifurcations in elliptic systems. Although we con-
jecture that our analysis may be generalized to the case of N competing species for
any N ≥ 2, our results can only be proved in the case N = 2 for the time being.
Let us come to technical statements.
We study the nonnegative steady-state solutions of the reaction-diffusion system
∂t R = a0 ∆R − F1 (x, R)U − F2 (x, R)V − m0 (x)R + I,
∂t U = a1 ∆U + (F1 (x, R) − m1 (x)) U, (x ∈ Ω, t > 0),
∂t V = a2 ∆V + (F2 (x, R) − m2 (x)) V,
where Ω is a bounded region in Rn with smooth boundary. The above system is
supplemented with Neumann1 boundary conditions
∂n R(t, x) = ∂n U (t, x) = ∂n V (t, x) = 0 (x ∈ ∂Ω, t > 0),
where ∂n is the normal derivative on the boundary ∂Ω.
The above system describes a situation where two species with density U = U (t, x)
and V = V (t, x) respectively, compete for the same resource with density R = R(t, x),
through the nonlinear terms Fi (x, R)U and Fi (x, R)V (i = 1, 2). Besides, the space
dependent resource R, as well as the two species U , V , follow a diffusion process in
space, with the distinct diffusion constants a0 > 0, a1 > 0, a2 > 0 respectively2 . The
space dependent functions mi (x) > 0 on Ω (i = 0, 1, 2), are death rates, while the
space dependent functions Fi (x, R) = Fi (x, R(t, x)) ≥ 0 are the consumption rates.
The given, time-independent function I = I(x) ≥ 0 is the nutrient input. All these
data are assumed smooth.
In order to implement a bifurcation method, we normalize the consumption rates
as follows. We readily choose given, smooth, functions f1 = f1 (x, R), f2 = f2 (x, R),
and introduce two bifurcation parameters c1 > 0 and c2 > 0, which somehow measure
the strength of the interaction between the species and the resource, through
F1 (x, R) ≡ c1 f1 (x, R), F2 (x, R) ≡ c2 f2 (x, R). (1.1)
1 Robin boundary conditions, of the form a ∂ R + b (x)R = g(x), a ∂ U + b (x)U =
0 n 0 1 n 1
a2 ∂n V + b1 (x)V = 0 on ∂Ω, with g(x) ≥ 0 and bi (x) ≥ 0 (i = 0, 1, 2), would do as well, as we
discuss later in this text.
2 Our analysis is valid when the various constant coefficients diffusion operators a ∆ become
i
div ai (x)∇ for some smooth, space-dependent coefficients ai (x) > 0 on Ω, provided all coef-
ficients ai (x) are proportional, i.e. ai (x) = λi a0 (x) (i = 1, 2) for some constants λ1 > 0 and
λ2 > 0. This easy extension is discussed later in the text. Needless to say, in that case, Robin
boundary conditions become a0 (x)∂n R + b0 (x)R = g(x) on ∂Ω, and so on, with g(x) ≥ 0 and
bi (x) ≥ 0 on ∂Ω (i = 0, 1, 2).
3
Note that, since we are only interested in nonnegative solutions (R, U, V ), the only
important data is the value of fi (x, R) for R ≥ 0: as shown by our analysis, any
smooth extension of fi (x, R) may be retained for values R ≤ 0, provided fi (x, R) ≤ 0
whenever R ≤ 0.
With the above notations, in this paper we look for stationary solutions U = U (x),
V = V (x), R = R(x) to the above system, namely3
(m0 (x) − a0 ∆)R + c1 f1 (x, R) U + c2 f2 (x, R) V = I(x),
(m1 (x) − a1 ∆)U − c1 f1 (x, R) U = 0, (x ∈ Ω)
(m2 (x) − a2 ∆)V − c2 f2 (x, R) V = 0, (1.2)
∂n R = ∂n U = ∂n V = 0 (x ∈ ∂Ω).
More precisely, our goal is to exhibit coexistence solutions in (1.2), i.e. solutions R, U ,
V for which R > 0, U > 0, V > 0. Our approach relies on a global bifurcation method,
where c1 and c2 are used as bifurcation parameters. In that respect, we also aim at
identifying a domain in the (c1 , c2 )-plane for which coexistence holds.
Let us come to some bibliographical comments.
Bifurcation methods have been used in many texts concerning interacting species
(competition models, predator-prey systems), see [20, 21, 23, 22] and more recently in
the study of some age structured models, see [8, 9]. In that respect, we wish to stress
that the chemostat involves a fairly specific mathematical structure, a fact that plays
a crucial role below: the nonlinear coupling in (1.2), say, only involves terms of the
form fi (x, R) U or fi (x, R) V ; in other words the two species U and V in (1.2) are
only coupled through the resource R. This observation holds in any chemostat model
and allows, in some situations, to reduce the original model to a standard competition
system by eliminating the equation on the resource, see [13, 12, 11, 3, 18, 19, 10].
Steady states of unstirred chemostats have been first studied by Waltman et al.
in [11]. The authors consider two species evolving in the one-dimensional situation
Ω = [0, 1]. A generalisation in the case of two species evolving in a higher dimensional
domain Ω is studied by Wu [18] and Wu and Nie [19]. Using the index in a positive
cone (see [24]), Zheng et al. [15, 14] show coexistence results in systems with various
trophic levels. In all these texts, the heterogeneity in space, that is crucial to recover
coexistence phenomena, is introduced by imposing a gradient of the resource, which in
turn is obtained through the boundary condition, of Robin type. All other coefficients
are space independent. In the present text at variance, we allow the reaction terms
(and other less crucial coefficients) to actually depend on space.
A key point is the following. In all the above works, the authors assume that the
competing species, and the resource, have the same diffusion rate and the same death
rate. This assumption provides a specific conservation law, that links the resource and
the competing species. In our case it reads (taking a0 = a1 = a2 = a and m0 (x) =
m1 (x) = m2 (x) = m(x))
Relation (1.3) allows to eliminate the resource R from the equations, and to write a
reduced system whose semi-trivial solutions satisfy a simple, scalar, elliptic equation.
3 Recall that Robin boundary conditions are covered by our analysis, as well as variable
coefficients diffusion operators div ai (x)∇, provided ai (x) = λi a0 (x) (i = 1, 2), see footnotes 1
and 2.
4
(m0 (x) − a0 ∆)R + (m1 (x) − a1 ∆)U + (m2 (x) − a2 ∆)V = I. (1.4)
Eliminating the unknow R in (1.4) leads to nonlocal semi-trivial problems. We are able
to study these semi-trivial problems by using a lower-upper solutions technique in the
so-obtained scalar, nonlocal, elliptic equations. In an independent step, a specific use of
global bifurcation techniques then allows to construct true coexistence solutions (U >
0, V > 0), starting from the semi-trivial solutions (U > 0, V = 0) or (U = 0, V > 0).
This is a key step of our approach. We wish to stress that the lower-upper solutions
part of our analysis requires (see Assumption 2 below) the crucial hypothesis4
mi (x) m (x)
∀x ∈ Ω, ≤ 0 (i = 1, 2). (1.5)
ai a0
It means that the ratio between death rate and diffusion rate should be larger for
the resource than for the competing species, or, in other words, that the two species
should diffuse relatively faster than the resource. Since spatial heterogeneity, and the
associated diffusion processes, are the key to obtaining systems which allow coexistence,
this assumption is quite natural: diffusion of the competing species helps obtaining
coexistence situations. To be complete, let us mention that in the case when Robin
boundary conditions are retained, another crucial assumption appears, namely5
bi (x) b (x)
∀x ∈ ∂Ω, ≤ 0 (i = 1, 2). (1.6)
ai a0
Assumption (1.6) is similar to (1.5) in spirit, in that a stronger ratio between the
escape rate and the diffusion rate is required for the resource R at the boundary, in
comparison with the analogous ratio for populations U and V .
The organization of the paper is as follows. In section 2 we present the notations and
recall some technical results used in the paper. We also state our main results, namely
Theorems 2.14 and 2.16. In section 3, we construct the above mentioned semi-trivial
solutions. Under assumption 2, the lower-upper solutions method, in conjunction with
bifurcation arguments, allows to prove existence, uniqueness, and non-degeneracy of
the semi-trivial solutions. Section 4 is the main step of our study, in that we prove the
4 In the case when the diffusion operators a ∆ become div a (x)∇ with a (x) = λ a (x)
i i i i 0
(i = 1, 2), the condition below becomes mi (x)/ai (x) ≤ m0 (x)/a0 (x) for x ∈ Ω (i = 1, 2).
5 This assumption obviously becomes b (x)/a (x) ≤ b (x)/a (x) for x ∈ ∂Ω (i = 1, 2), when
i i 0 0
the ai ’s depend on x.
5
existence of solutions (R, U, V ) to (1.2) that satisfy R > 0, U > 0, V > 0. A global
bifurcation theorem is used to construct these coexistence solutions, by joining the two
families of semi-trivial solutions. Our construction leads to define a domain Θ ⊂ R2+ in
the space of bifurcation parameters (c1 , c2 ), called the coexistence domain. This domain
is such that whenever (c1 , c2 ) ∈ Θ, a coexistence solution is at hand. In section 5, we
state some consequences of our analysis, which provide an ecological point of view.
Section 6 concludes this paper.
2.1 Generalities
For i = 0, 1, 2, the constants ai are supposed positive, and the fonctions mi (x) and
I(x) are assumed smooth, with mi (x) > 0 on Ω and I(x) ≥ 0 and I(x) 6≡ 0 on Ω.
Taking a given α ∈ (0, 1) whose value is irrelevant, we define the spaces6
Ai := mi (x) − ai ∆. (2.2)
In order to keep simple notations, the above operator will always be denoted by the
same symbol Ai for any choice of α. In the similar spirit we note
Ki := A−1
i . (2.3)
For each i = 0, 1, 2, the operator Ki is compact when seen as (more precisely : when
extended to) an operator from C 1 (Ω) to C 1 (Ω) and from L2 (Ω) to L2 (Ω). Note that
each operator Ki maps X to X compactly as well. Recall that the strong maximum
principle for elliptic operators with Neumann (or Robin) boundary conditions reads,
whenever u ∈ X,
Ai u ≥ 0
∂n u ≥ 0 =⇒ min u(x) = m > 0. (2.4)
x∈Ω
u 6≡ 0
The strong maximum principle also implies the following uniqueness
Ai u = 0
=⇒ u ≡ 0. (2.5)
∂n u = 0
We last recall the following standard Lemmas
6 with the obvious adaptation if Robin boundary conditions and/or variable coefficients a ’s
i
are retained: to each operator div ai (x) ∇−mi (x) with boundary condition ai (x)∂n ·+bi (x)· = 0
is associated the space Xi = {u ∈ C 2+α (Ω), ai (x)∂n u + bi (x)u = 0 on ∂Ω}, and the triple
(R, U, V ) then is to be exhibited in X0,+ × X1,+ × X2,+ .
6
Lemma 2.1 Take m(x) ∈ C α (Ω) and q(x) ∈ C α (Ω). Assume m(x) > 0 for all x ∈ Ω.
Take a ∈ R∗+ . Then the eigenvalue problem
Lemma 2.2 Take q(x) ∈ C α (Ω), a ∈ R∗+ such that q(x) > 0 for any x ∈ Ω. Then the
eigenvalue problem
(m(x) − a∆)φ = µq(x)φ, ∂n φ = 0,
In order to make use of a lower-upper solution technique later in this text, we readily
introduce the following assumption
∂fi
∀R > 0, fi (x, R) > 0, and (x, R) > 0.
∂R
mi (x)/ai ≤ m0 (x)/a0 .
7 Recall that we are only interested in situations with R ≥ 0, hence the way we extend f
i
for negative values of R is irrelevant.
8 See footnote 4 in the case of variable coefficients diffusion operators.
7
As we show now, this condition provides a monotonicity property that plays a key
rôle in our analysis. Whenever w ∈ X+ , define Ri (w) ∈ X as the unique solution in X
to
A0 Ri (w) + Ai w = I. (2.6)
The operator w 7→ Ri (w) is introduced for the following reason. The one-species prob-
lem (corresponding to semi-trivial solutions (U > 0, V = 0) say), reads
R = R1 (U ), A1 U − c1 f1 (x, R1 (U )) U = 0, (2.8)
and R1 (U ) may be seen as the resource at hand in the presence of the population U .
In any circumstance, the one-species problem leads to considering the above nonlinear
and nonlocal elliptic problem, with nonlinearity w 7→ f1 (R1 (w)) w.
Now, an easy computation provides the alternative formula9 .
ai 1
Ri (w) = K0 (I) − K0 A 0 w + K0 (ai m0 (x) − a0 mi (x))w . (2.9)
a0 a0
A key point is the fact that the nonlocal term K0 (ai m0 − a0 mi )w above satisfies
K0 A0 w ≤ w whenever w ≥ 0. (2.11)
This comes from the maximum principle together with the fact that, when w ≥ 0, the
function v = K0 A0 w satisfies A0 (v − w) = 0 with the boundary condition (a0 ∂n +
b0 )(v − w) = +(a0 b1 − b0 a1 )w/a1 ≤ 0.
We readily show that Assumption 2 implies the following one-sided Lipschitz con-
dition for the nonlinearity w 7→ f1 (R1 (w)) w in (2.8).
Lemma 2.3 Suppose Assumption 2 is true. Let M be a positive constant and take
i = 1, 2. Then, there exists γ = γi (M ) > 0 such that
w1 (x) fi (x, Ri (w1 ))(x) − w2 (x) fi (x, Ri (w2 ))(x) ≥ −γ(w1 (x) − w2 (x))
whenever w1 , w2 ∈ X satisfy 0 ≤ w2 ≤ w1 ≤ M .
Remark 2.4 The point is, the above estimate is pointwise in x, though it involves the
nonlocal operator Ri . ⊓
⊔
9 When the diffusion operators become div a (x)∇ with a (x) = λ a (x), see footnotes 2
i i i 0
and 4, the formula below becomes Ri (w) = K0 (I) − λi K0 A0 w + K0 (λi m0 (x) − m1 (x))w ,
with λi m0 (x) − m1 (x) ≥ 0 for all x, and our analysis is unchanged.
8
Remark 2.5 If all diffusion operators are the same, as in the previously quoted papers,
namely if Ai ≡ A0 (i = 1, 2), then the nonlocal terms of the form K0 (ai m0 − a0 mi )w
vanish in the course of the analysis. In that particular case, the method we develop
coincides with that of [18]. The nonlocal terms constitute the main difficulty we treat.
⊓
⊔
Admitting Lemma 2.3 is proved for the moment, we readily state that this result
allows us to apply a lower-upper solution method in the nonlocal elliptic system
where w ∈ X is the unknown. Indeed, using Lemma 2.3, the following definition and
Theorem are standard (see [5]).
Remark 2.8 Stricto sensu the above Theorem is not to be found in [5]. Smoller requires
the nonlinear term be Lipschitz in w, a property that we do not have at hand in
the present case. It is standard to observe that the key of the proof, which relies on
an iteration of the maximum principle, is the following. When writing the equation
Ai w = ci fi (x, Ri (w)) w =: Gi (x, w), the point is to find a (large) K > 0 and a
(large) M > 0 such that whenever 0 ≤ W1 (x) ≤ W2 (x) ≤ M for all x, we have
G(x, W1 )(x) + KW1 (x) ≤ G(x, W2 )(x) + KW2 (x) for all x as well. The one-sided
Lipschitz estimate of Lemma 2.3 is enough in that respect.
Note that Pao [7, 6] establishes variants of the above techniques for systems, in
the case where the nonlinear terms, which are vector-valued, satisfy so-called quasi-
monotonicity properties. ⊓ ⊔
M
kRi (w)kL∞ ≤ kK0 (I)kL∞ + M kK0 (m0 )kL∞ + kK0 (ai m0 − a0 mi )kL∞ =: M∞ .
a0
10 With the obvious extension in the case of Robin boundary conditions.
9
ai 1
Ri (w1 ) − Ri (w2 ) = − K0 A0 (w1 − w2 ) + K0 (ai m0 (x) − a0 mi (x)) (w1 − w2 )
a0 a0
a
≥ − i K0 A0 (w1 − w2 )
a0
a
≥ − i (w1 − w2 ).
a0
where the first lower bound uses Assumption 2 while the second uses the observation
(2.10). Hence, writing
we distinguish two cases. If x is such that Ri (w1 )(x) ≥ Ri (w2 )(x), then fi being an
increasing function of R, we recover
We state the two bifurcation theorems we use in the sequel; for equations of the form,
T (c, W ) = W,
∀c ∈ R, T (c, 0) = 0.
10
Then, there exists ε > 0 and a map (c(s), X(s)) ∈ C 0 ((−ε, ε); R × Y ), with c(0) = c0 ,
X(0) = 0, such that close to (c0 , 0) in R × Y , the only nontrivial solution to T (c, W ) =
W is given by
T (c, W ) = W
⇐⇒ ∃s ∈ (−ε, ε) such that (c, W ) = c(s), sW0 + s X(s) .
(c, W ) 6= (c, 0)
We complete the picture by stating a global version of the theorem. Some additional
assumptions are required. We need the following compactness assumption
T : R × Y → Y is a compact operator, and ,
∀(c, W ), T (c, W ) = DW T (c, 0) · W + R(c, W ), (2.13)
where DW T (c, 0) is a linear compact operator.
In other words we assume that the linearized part of equation T (c, W ) = W , close to
the trivial solution W = 0, is always a compact perturbation of the identity.
Now, for those values of c such that the trivial solution W = 0 is an isolated
solution to T (c, W ) = W , i.e. typically whenever DW T (c, 0) does not admit 1 as an
eigenvalue, one may define the index of the solution W = 0, as the Leray-Schauder
degree deg(Id − T (c, ·), B, 0) (here B ⊂ Y is a ball centred at 0 such that W = 0 is
the only solution to T (c, W ) = W in B). In other words, the index of the considered
solution W = 0 is
i(T (c, ·), 0) := deg(Id − T (c, ·), B, 0). (2.14)
It has the value
i(T (c, ·), 0) = deg(Id − DW T (c, 0), B, 0) = (−1)p , (2.15)
where p is the sum of the algebraic multiplicities of all (real) eigenvalues of DW T (c, 0)
that are greater than 1.
The following theorem holds true
Theorem 2.10 (Global bifurcation from a simple eigenvalue – see [20, 18])
Under the assumptions and notation of Theorem 2.9, we suppose that T is a compact
operator such that DW T (c, 0) is linear compact for any c, as in (2.13).
We also assume11 that for some ε > 0, the index i(T (c, ·), 0) is constant on (c0 −
ε, c0 ) and on (c0 , c0 + ε), and that whenever c0 − ε < α < c0 < β < c0 + ε we have
i(T (α, ·), 0) 6= i(T (β, ·), 0).
11 This second assumption is not needed when D T (c, 0) does not depend on c. In our case –
c
see below – we shall apply this Theorem for T ’s of the form T (c, W ) = A + cB(W ) where A is
a constant and B a compact operator independent of c. This is due to our choice of bifurcation
parameters: they are only involved in the two terms c1 f1 (R)U and c2 f2 (R)V in (1.2), terms
which are proportional with c1 resp. c2 . We nevertheless describe our bifurcation method in
the present more general form, in order to keep a procedure that applies as well in the case
of a nonlinear dependence on the bifurcation parameters, as would be the case when choosing
(c1 , c2 )-dependent consumption rates for instance.
11
by starting from the 0-species problem (namely trivial solutions corresponding to R >
0, U = 0, V = 0),. Then we construct 1-species, or semi-trivial, solutions (corresponding
to R > 0, and either (U > 0, V = 0) or (U = 0, V > 0)), by using lower-upper solutions
techniques. This step is complemented with the use of bifurcations from the 0-species
problem, to prove the non-degeneracy of the so-obtained semi-trivial solutions, and to
compute the index of these solutions. This step is crucial, and makes a strong use of our
Assumption 2. It is the most difficult and technical part of our analysis. Armed with
these results, we then use bifurcations again to construct true coexistence solutions
R > 0, U > 0, V > 0. This last step uses all informations gathered on the semi-trivial
solutions.
A0 S = I. (2.17)
3
(ii) If (R, U, V ) ∈ X+ is a solution to (1.2) with U 6≡ 0 or V 6≡ 0, then13 0 < R < S.
(iii) Let w ∈ X+ . The equation
A0 R + ci fi (x, R) w = I
∗
(Ru (c1 ), U ∗ (c1 )).
∗ ∗
Ru (c1 ) −→ S, and Ru (c1 ) −→ 0,
c1 →c01 c1 →+∞
∗
where U∞ ∈ X+ is the unique solution to A1 U∞ = I.
15
In other words, Ker Id − D(R,U ) T1 (c1 , Ru ∗ (c ), U ∗ (c )) = {0}.
1 1
16 Our proof not only provides that the index of this solution (R∗ (c ), U ∗ (c )) is one, but
u 1 1
also that all eigenvalues of Id − D(R,U ) T1 (c1 , Ru ∗ (c ), U ∗ (c )) are less than one whenever
1 1
c1 > c01 is close to c01 . This implies that the so-obtained solution is stable, i.e. the associated
time-dependent parabolic problem admits (Ru ∗ (c ), U ∗ (c )) as a locally stable steady state.
1 1
17 This apparently technical statement is the key to constructing true coexistence solutions
Remark 2.13 In fact, the mere existence of semi-trivial solutions may be obtained us-
ing a simple global bifurcation argument, without making use of our Assumption 2.
Assumption 2 is required at variance to obtain uniqueness of these solutions. This as-
sumption also plays a key rôle to establish non-degeneracy, and to compute the value
of the index. ⊓
⊔
Naturally, the similar results hold in the case U ≡ 0 and V > 0. This provides a
critical value c02 , and a solution branch (Rv∗ (c2 ), V ∗ (c2 )) ∈ (X+ ∗ 2
) whenever c2 > c02 ,
which satisfies the properties similar to the ones listed before. The natural semi-trivial
∗
solutions to (1.2) are (R, U, V ) = (Ru (c1 ), U ∗ (c1 ), 0) (with c1 > c01 ), and (R, U, V ) =
(Rv (c2 ), 0, V (c2 )) (with c2 > c2 ). We define the following two subsets of R2+ × X+
∗ ∗ 0 3
,
namely
n o
∗
Cu = (c1 , c2 , Ru (c1 ), U ∗ (c1 ), 0) ; c1 > c01 ,
n o
Cv = (c1 , c2 , Rv∗ (c2 ), 0, V ∗ (c2 )) ; c2 > c02 . (2.20)
With this notation at hand, the following Theorem is the main result of the present
paper. It establishes that coexistence solutions to (1.2) may be defined using bifurca-
tions from the two sets Cu and Cv . The proof is provided in section 5.2. Figure 2.1
illustrates the situation.
∗ 3
∀c2 ∈ c2 (c1 ), c2 (c1 ) , ∃(R, U, V ) ∈ X+ coexistence solution to (1.2).
∗ 3
∀c1 ∈ c1 (c2 ), c1 (c2 ) , ∃(R, U, V ) ∈ X+ coexistence solution to (1.2).
Remark 2.15 Note that the situation wherec∗2 (c1 ) = c∗∗ 2 (c1 ), say, may very well hap-
pen. In that case the interval c1 (c2 ), c1 (c2 ) is void. Hence, as we can see, the second
statement in part (i) of the Theorem is a weak byproduct of the first one, which ex-
hibits at variance an actual branch of coexistence solutions. We refer to the conjecture
stated in paragraph 6 below for a discussion of this point. ⊓ ⊔
14
V ∗ (c∗∗
2 )
∗ c∗∗ c2
c02 c2 2
U ∗ (c1 )
With the use of the above Theorem, one may define a coexistence domain Θ, as
Θ = (c1 , c2 ) ∈ (c01 , +∞) × (c02 , +∞), s.t.
c1 ∈ c1 (c2 ), c1 (c2 ) and c2 ∈ c2 (c1 ), c2 (c1 ) . (2.21)
It corresponds to values of the parameters (c1 , c2 ) for which a coexistence solution may
be exhibited (a subset of the set of all values (c1 , c2 ) such that a coexistence solution
may be exhibited – see paragraph 6 on that point).
The following Theorem is proved in section 5.3. It explores the structure of Θ.
c∗1 (c∗∗
2 (c1 )) = c1 ,
c∗1 (c2 ) : (c02 , +∞) −→ (c01 , +∞), and c∗2 (c1 ) : (c01 , +∞) −→ (c02 , +∞)
15
V
V ∗ (c2 )
c2
1)
2 (c
c∗ Θ−
c2
)
∗ (c 2
c1
U ∗ (α)
Θ+ U ∗ (β)
U ∗ (γ)
c02
U
c01 α β γ c1
c∗∗
2 (c 1 ).
Figure (b) represents some bifurcating solutions corresponding to three values α, β and γ of
the parameter c1 > c01 . The retained values are here assumed to satisfy c∗2 (α) < c∗∗ 2 (α), resp.
c∗2 (β) = c∗∗ 0
2 (β), resp. c2 (γ) > c2 (γ). For each c1 > c1 , there is a coexistence solution joining
∗ ∗∗
are continuous and increasing. Moreover, for {i, j} = {1, 2}, we have
(iii) With the notation (2.21), whenever (c1 , c2 ) ∈ Θ, system (1.2) has a coexistence
∗ 3
solution (R, U, V ) ∈ X+ , and we have
Θ = Θ− ∪ Θ+ , with Θ− = {(c1 , c2 ), c1 < c∗1 (c2 ) and c2 < c∗2 (c1 )},
and Θ+ = {(c1 , c2 ), c1 > c∗1 (c2 ) and c2 > c∗2 (c1 )}.
The next sections are devoted to the proof of Theorem 2.11 (trivial solutions),
Theorem 2.12 (semi-trivial solutions), as well as Theorems 2.14 and 2.16 (coexistence
solutions and coexistence domain).
We prove here the various statements of Theorem 2.11. Recall that the problem with
zero species reads, shortly, A0 R = I.
Point (i). Existence and uniqueness of S is clear.
3
Point (ii). Let (R, U, V ) ∈ X+ be a solution to (1.2) with U ≥ 0 and V ≥ 0. We have
A0 R = I − c1 f1 (x, R)U − c2 f2 (x, R)V ≤ I. Hence A0 R ≤ I with A0 R 6≡ I whenever
16
A0 R + ci fi (x, R) w = I. (3.1)
In this section, we study the one species problem (2.18), corresponding to the semi-
∗ ∗
trivial solution (R, U, 0) ∈ X+ × X+ × X+ to (1.2). Recall that the one species problem
reads
A0 R + c1 f1 (x, R)U = I,
A1 U − c1 f1 (x, R)U = 0.
Lemma 4.1 Let c1 > 0 be fixed. There exists M0 > 0 such that each solution (R, U ) ∈
∗ 2
(X+ ) to (2.18) verifies
0 ≤ U ≤ M0 .
∗ 2
Proof of Lemma 4.1. Let (R, U ) ∈ (X+ ) be a solution to (2.18). Summing the equa-
tions on R and U provides, as already noted, A0 R + A1 U = I. As a consequence, for
some α > 0 small enough we have
(α − ∆) (a0 R + a1 U ) ≤ I.
Z
18 b0 (x) R+ − R− ≥ 0, and the conclu-
Robin boundary conditions would add a term
∂Ω
sion would remain unchanged.
17
1
The strong maximum principle19 then provides 0 ≤ a0 R + a1 U ≤ kIkL∞ . In the case
α
of variable coefficients ai (x) with ai (x) = λi a0 (x), see footnote 2, the argument is the
same, due to the bound (α − div a0 (x)∇) (R + λ1 U ) ≤ I = (m0 (x) − div a0 (x)∇) R
+ (m1 (x) − λ1 div a0 (x)∇) U. ⊓ ⊔
Lemma 4.2 The eigenvalue problem A1 φ − µf1 (x, S)φ = 0 with φ ∈ X has a prin-
cipal eigenvalue c01 > 0 and a corresponding eigenfunction φ0 ∈ X+
∗
, unique up to a
multiplicative constant. We have
Proof of Proposition 4.3. The function U > 0 verifies A1 U −c1 f1 (R)U = 0. Multiplying
by φ0 , defined in Lemma 4.2, and integrating over Ω leads to
Z Z Z
0= A1 U φ0 − c 1 f1 (x, R)U φ0 = U φ0 (c01 f1 (x, S) − c1 f1 (x, R)).
Ω Ω Ω
Since Proposition 2.11 ensures R < S hence f1 (x, R) < f1 (x, S), we recover the neces-
sary condition c1 > c01 . ⊓
⊔
4.2 Existence, uniqueness, and some properties of solutions to the one species problem
Proposition 4.4 Suppose Assumptions 1 and 2 are verified. Assume c1 > c01 .
∗ 2 ∗
Then, system (2.18) has a unique solution in (X+ ) , denoted by (Ru (c1 ), U ∗ (c1 )).
∗ 2
Proof of Proposition 4.4. Take a solution (R, U ) ∈ (X+ ) to (2.18). Defining, as in
(2.9), the quantity R1 (U ) ∈ X by the relation A0 R1 (U ) + A1 U = I we recover the
necessary condition R = R1 (U ), and system (2.18) can be rewritten (with ∂n U = 0 on
∂Ω),
A1 U − c1 f1 (x, R1 (U ))U = 0, (4.2)
Let φ0 > 0 be the eigenfunction defined in Lemma 4.2, which satisfies A1 φ0 −
c01 f1 (x, S)φ0 = 0. We claim that for ε > 0 small enough and M > 0 large enough,
the pair (εφ0 , M ) is a pair of lower-upper solutions to (4.2). Indeed, on the one hand,
choosing M > 0 large enough leads to R1 (M ) < 0 (since A0 R1 (M ) = I − A1 M =
I − m1 (x)M ). Therefore, we obtain
A1 M − c1 M f1 (x, R1 (M )) ≥ m1 M ≥ 0,
19 with the obvious adaptation in the case of Robin boundary conditions.
18
with ∂n (εφ0 ) = 0 on ∂Ω. Therefore εφ0 is a lower solution to (4.2) for ε small enough.
These considerations allow us to conclude (see Theorem 2.7) that there exists a
pair (U − , U + ) of maximal solutions to (4.2), satisfying εφ0 < U − ≤ U + < M , and
for any solution U ∈ [εφ0 , M ] to (4.2) we necessarily have U − ≤ U ≤ U + . Besides,
∗
Lemma 4.1 ensures one can choose M ≥ M0 such that any solution U ∈ X+ to (4.2)
anyhow satisfies 0 ≤ U ≤ M . Remembering that 0 is a lower-solution, we thus obtain
∗
that every solution U ∈ X+ necessarily verifies 0 ≤ U ≤ U + as well.
Let us show that U = U . We first observe that the relation 0 ≤ U ≤ U + implies
+
0 ≤ R1 (U + ) ≤ R1 (U ).
(1)
This is due to Theorem 2.11, together with the fact that R1 (U ) = RU and R1 (U + ) =
(1)
RU + in the present case (for U and U + solve the auxiliary equation A1 U = c1 f1 (. . .)U
and similarly for U + ). We deduce f1 (x, R1 (U + )) ≤ f1 (x, R1 (U )). On the other hand,
the obvious integration by parts, together with the definition of U and U + , provide
Z h i Z
0= [A1 U ] U + − A1 U + U = c1 U U + f1 (x, R1 (U )) − f1 (x, R1 (U + )) .
Ω Ω
With the above Proposition at hand, we complete the picture by stating some
∗
properties of the pair (Ru (c1 ), U ∗ (c1 )). We begin with the asymptotic behaviour as
c1 → ∞.
∗
lim kRu (c1 )k∞ + kU ∗ (c1 ) − U∞ k∞ = 0.
c1 →+∞
On the other hand, take an ε > 0 fixed. For c1 large enough, the function (1−ε)U∞
∗
is a lower-solution
to A1 U − c1 f1 (x, R1 (U )) U = 0 in X+ . Indeed, we have R1 (1 −
ε)U∞ = εK0 (I) = εS > 0 on Ω, so that
∗ a1 1
Ru (c1 ) = − K0 A0 (U ∗ (c1 ) − U∞ ) + K0 (a1 m0 − a0 m1 )(U ∗ (c1 ) − U∞ ) .
a0 a0
(with the similar formula if the coefficients ai become space-dependent, with a1 (x) =
λ1 a0 (x) and a2 (x) = λ2 a0 (x) – see footnotes 2, 4 and 9). Using the fact that U ∗ (c1 ) ≤
∗
U∞ , Assumption 2, and, more precisely, relations (2.10) and (2.11), give 0 ≤ Ru (c1 ) ≤
a1 ∗ ∗
(U∞ − U (c1 )). Using the established limiting behaviour of U (c1 ) we deduce
a0
∗
lim kRu (c1 )k∞ = 0. ⊓ ⊔
c1 →+∞
Proposition 4.6 With the notation of Proposition 4.4 the map c1 7→ U ∗ (c1 ) is in-
creasing from (c01 , +∞) to X+
∗ ∗
, while the map c1 7→ Ru (c1 ) is decreasing from (c01 , +∞)
∗
to X+ .
We observe that
A1 U ∗ (b1 ) − b2 f1 x, R1 (U ∗ (b1 )) U ∗ (b1 ) = (b1 − b2 )f1 x, R1 (U ∗ (b1 )) U ∗ (b1 ) < 0,
The previous paragraph, and more precisely Proposition 4.4 shows that two families
of solutions to the one-species problem (2.18) coexist whenever c1 > c01 , namely the
∗
trivial (c1 , S, 0) and the semi-trivial (c1 , Ru (c1 ), U ∗ (c1 )). As an immediate consequence,
0 2
it appears that (c1 , S, 0) ∈ R ×(X+ ) is a bifurcation point for system (2.18). Note that
∗
the bifurcation solution (c1 , Ru (c1 ), U ∗ (c1 )) is readily constructed for all values c1 >
c01 , without using the Crandall-Rabinowitz theorem, so that it is not even clear that
∗
the branch (c1 , Ru (c1 ), U ∗ (c1 )) actually coincides with a bifurcation in the Crandall-
Rabinowitz sense (for instance, the limit as c1 → c01 of (Ru ∗
(c1 ), U ∗ (c1 )) may well differ
from (S, 0) at this stage).
In this section, we show essentially two results. On the one hand we show that
the Crandall-Rabinowitz theorem applies, and uniqueness allows to conclude that the
∗
already constructed semi-trivial solution (c1 , Ru (c1 ), U ∗ (c1 )) coincides with the one
obtained by bifurcation. On the other hand, and as a consequence, we deduce various
properties such as the non-degeneracy of the semi-trivial branch, or we compute the
index of this branch. This part of the analysis prepares for the next section where we
construct coexistence solutions to the full 2-species problem.
We begin with the
Proposition 4.7 (Local bifurcations in the one-species problem (2.18))
∗
With the above notation, let φ0 ∈ X+ and c01 > 0 be as in Lemma 4.2. Define
0 ∗
ρ0 = c1 K0 (f1 (S)φ0 ) ∈ X+ . On the other hand, recall from (2.19) the definition
is a familly of positive solutions to (2.18). We set R(s) = S − s(ρ0 + rb(s)) and U (s) =
s(φ0 + u b(s)).
Moreover, each solution (c1 , R, U ) ∈ R ×(X+ )2 to (2.18) near (c01 , 0, 0) is either the
trivial solution (c1 , S, 0), or it coincides with (c1 (s), R(s), U (s)) for some s ∈ (−ε, ε).
In particular, for any c1 > c01 , close to c01 , there exists s > 0, such that
∗
(c1 , Ru (c1 ), U ∗ (c1 )) = (c1 (s), R(s), U (s)).
for the time-dependent parabolic problem associated with the present stationary problem.
21
∗
Remark 4.8 Point (i) establishes that the branch (c1 , Ru (c1 ), U ∗ (c1 )) constructed so
far coincides at least locally with the bifurcation branch (c1 (s), R(s), U (s)).
Point (ii) plays a crucial rôle later in the analysis, when exhibiting coexistence solu-
tions to the full 2-species system. We stress the fact that the computation of the above
index uses tools from bifurcation theory, hence relies on the identification between the
∗
bifurcation branch (c1 (s), R(s), U (s)) and the branch (c1 , Ru (c1 ), U ∗ (c1 )). ⊓
⊔
With these notations at hand, we show that the Crandall-Rabinowitz Theorem 2.9
applies at the bifurcation point (c01 , S, 0).
Firstly, let (ρ, φ) ∈ Ker(L1 (c01 )). We have
Dc1 L1 (c01 ) · (−ρ0 , φ0 ) = t (+K0 (f1 (S)φ0 ), −K1 (f1 (S)φ0 )).
Arguing by contradiction, if Dc L1 (c01 ) · (−ρ0 , φ0 ) ∈ Im L1 (c01 ) , there exists φ and ρ
in X such that
(+K0 (f1 (S)φ0 ) , −K1 (f1 (S)φ0 )) = ρ + c01 K0 (f1 (S)φ) , φ − c01 K1 (f1 (S)φ) .
22 Here and below we abuse notation by writing f1 (R) instead of f1 (x, R) and so on.
23 Robin boundary conditions lead to the same calculation.
22
defined in the neighbourhood of s = 0, resp. c1 = c01 , such that along the trivial solution
(c1 , S, 0), we have
Id − D(R,U ) T1 (c1 , S, 0) · w0 (c1 ) = γ(c1 ) w0 (c1 ),
while along the semi-trivial solution (c1 (s), R(s), U (s)) we have
Id − D(R,U ) T1 (c1 (s), R(s), U (s)) · w(s) = µ(s)w(s),
In order to prove that i T1 (c1 (s), ·), (R(s), U (s)) = 1 for small values of s > 0,
we now show that D(R,U ) T1 (c1 (s), R(s), U (s)) has no eigenvalue greater than one (see
equation (2.15)), i.e. all eigenvalues of Id − D(R,U ) T1 (c1 (s), R(s), U (s)) are positive.
Since µ(s) is the smallest eigenvalue of Id − D(R,U ) T1 (c1 (s), R(s), U (s)) (thanks to
Lemma 2.1, and using the value of the above operator together with the fact that the
components of w(0) = (−ρ0 , φ0 ) are uniformly negative resp. positive on Ω, so that
the same property holds for the components of w(s) = (−ρ(s), φ(s)), at least for small
values of s), we therefore need to show µ(s) > 0 for small values of s > 0.
To do so we use the following known fact from local bifurcation theory, (see [5] p.
179), namely
This is the key piece of information here. There remains to study the signs of the various
terms on the right-hand-side of (4.4). Concerning c′1 (s), if s > 0 is small enough, we
have, by definition of c1 (s), the relation
b(s)) = s c1 (s) f1 S − s(ρ0 + rb(s))
s A1 (φ0 + u b(s) .
φ0 + u
d
Dividing by s and computing ds |s=0 , gives
Applying d
dc1 |c1 =c01 , multiplying by φ0 = φ0 (c01 ), using γ(c01 ) = 0, and integrating over
24
Ω leads to Z Z
− f1 (S) φ20 = +γ ′ (c01 ) f1 (S)φ20 .
Ω Ω
Hence γ ′ (c01 ) < 0. Eventually we have established that µ(s) > 0 whenever s > 0 is
small. This provides i T1 (c1 (s), ·), (R(s), U (s)) = 1 whenever s > 0 is small.
The proof is complete. ⊓ ⊔
Proof of Proposition 4.9. Using Theorem 4.7, together with the uniqueness statement
of Theorem 4.4, we have (R(s), U (s)) = (R∗ (c1 (s)), U ∗ (c1 (s))). Since lim c1 (s) = c01 ,
s→0
the result follows from the continuity of s 7→ (R(s), U (s)). ⊓
⊔
Proposition 4.10 With the notation of Proposition 4.4, for each c1 > c01 , we have
∗
Ker Id − D(R,U ) T1 (c1 , Ru (c1 ), U ∗ (c1 )) = {0}.
Hence we readily have H (U ∗ (c1 )) = U ∗ (c1 ), and the equivalence (4.6) also implies,
∗
since 0 is an eigenvalue of Id − D(R,U ) T1 (c1 , Ru (c1 ), U ∗ (c1 )), that 1 is an eigenvalue
∗
of Du H(U (c1 )) as well.
We claim that the operator H is nondecreasing, i.e. whenever U and V belong to
X, we have
This property is actually the reason for our introduction of the parameter K. It comes
from the fact that, according to Lemma 2.3, from U ≥ V ≥ 0, we deduce f1 (R1 (U )) U −
24 The computation is the same in the case of Robin boundary conditions.
24
(Note that k is a function in X). This is the key ingredient. It comes from the following
computation. We have
d
Du H U ∗ (c1 ) · U ∗ (c1 ) = H (1 + t)U ∗ (c1 )
dt t=0
d
= t=0
(A1 + K)−1 f1 R1 (1 + t)U ∗ (c1 ) + K (1 + t) U ∗ (c1 )
dt
= (A1 + K)−1 f1 R1 U ∗ (c1 ) + K U ∗ (c1 )
d
+ (A1 + K)−1 DR f1 R1 U ∗ (c1 ) U ∗ (c1 ) t=0
R1 (1 + t)U ∗ (c1 ) .
dt
while
d d
R1 (1 + t)U ∗ (c1 ) = t=0 K0 I − (1 + t) A1 U ∗ (c1 )
dt t=0 dt
d
= − t=0 (1 + t)K0 c1 f1 R1 (U ∗ (c1 )) U ∗ (c1 )
dt
= −K0 c1 f1 R1 (U ∗ (c1 )) U ∗ (c1 ) = R1 (U ∗ (c1 )) − S < 0.
k + µ < 1.
We define
provided ε is small enough. We have used relation (4.8) together with the fact that
U ∗ (c1 ) > 0.
Gathering all the above claims, let us now show that Du H (U ∗ (c1 )) cannot have 1
as an eigenvalue. Take φ ∈ X (φ 6≡ 0) such that
Du H(U ∗ (c1 )) · φ = φ.
Up to rescaling φ, we may assume that
−U ∗ (c1 ) ≤ φ ≤ U ∗ (c1 ).
For technical reasons that become clear later, we may rescale φ again, so as to ensure
that there is a point x0 ∈ Ω such that
(1 + µ)φ(x0 ) > U ∗ (c1 )(x0 ),
where µ > 0 is as before. The idea is to compute Hµ (U ∗ (c1 ) + εφ) in two different
ways, to obtain the desired contradiction.
On the one hand we have, from the relation (1 − ε)U ∗ (c1 ) ≤ U ∗ (c1 ) + εφ ≤
(1 + ε)U ∗ (c1 ), and using (4.10), the bounds
Hµ U ∗ (c1 ) + εφ ≤ Hµ (1 + ε)U ∗ (c1 ) ≤ (1 + ε)U ∗ (c1 ),
as well as Hµ (U ∗ (c1 ) + εφ) ≥ Hµ ((1 − ε)U ∗ (c1 )) ≥ (1 − ε)U ∗ (c1 ). On the other hand,
we may expand (the expansion holds in X)
Hµ U ∗ (c1 ) + εφ = Hµ U ∗ (c1 ) + εDu Hµ U ∗ (c1 ) · φ + O(ε2 )
= U ∗ (c1 ) + (1 + µ)εφ + O(ε2 )
= (1 + ε) U ∗ (c1 ) + ε((1 + µ)φ − U ∗ (c1 )) + O(ε2 ).
Hence, at the point x0 , we have Hµ (U ∗ (c1 ) + εφ) (x0 ) > (1 + ε) U ∗ (c1 )(0 ), provided ε
is small enough, which contradicts the fact that Hµ (U ∗ (c1 ) + εφ) ≤ (1 + ε)U ∗ (c1 ).
To summarize, the whole idea of our contradiction argument is that on the one
hand (1 + ε)U ∗ (c1 ) satisfies H(U ) < U in a strict fashion (as a consequence of (4.8)),
while the upper-lower solution technique, together with the fact that φ is associated
with the eigenvalue 1 of the linear part of H, imply that when perturbing U ∗ (c1 ) in the
direction φ, the function H must at the same time be almost constant in that direction
and it should decay in a strict fashion as well. ⊓ ⊔
∗
As an immediate consequence of the non-degeneracy of the solution (Ru (c1 ), U ∗ (c1 )),
together with the implicit function theorem, we deduce the
∗
Proposition 4.11 The map c1 7→ (Ru (c1 ), U ∗ (c1 )) is continuously differentiable from
(c01 , +∞) to X+
∗ ∗
× X+ .
Proof of Proposition 4.11.
∗
The pair (Ru (c1 ), U ∗ (c1 )) is defined by the equation
∗
T1 (c1 , Ru (c1 ), U ∗ (c1 )) = t (Ru
∗
(c1 ), U ∗ (c1 ).
∗
On the other hand, we have just proved that Id−D(R,U ) T1 (c1 , Ru (c1 ), U ∗ (c1 )) does not
admit 0 as an eigenvalue, while it is clear from the definition of T1 that the linearized
operator D(R,U ) T1 (c1 , R, U ) is compact for any value of (c1 , R, U ) ∈ R × X 2 . As a
∗
consequence, we have that Id − D(R,U ) T1 (c1 , Ru (c1 ), U ∗ (c1 )) is invertible, and the
local inversion Theorem applies. ⊓ ⊔
26
∗
i(T1 (c1 , ·), (Ru (c1 ), U ∗ (c1 ))) = 1.
∗
i(T1 (c1 , ·), (Ru (c1 ), U ∗ (c1 ))) = (−1)p(c1 ) ,
∗
where p(c1 ), is the number of eigenvalues of D(R,U ) T1 (c1 , Ru (c1 ), U ∗ (c1 )) that are
+ − 0
greater than 1. Now, take any c1 > c1 > c1 . By uniqueness of the solution to
∗ 2
T1 (c1 , R, U ) = (R, U ) in (X+ ) , for any c1 > c01 , we can choose a neighbourhood U of
the set {(c1 , Ru (c1 ), U (c1 )) ; c1 ∈ (c−
∗ ∗ + ∗ 2
1 , c1 )} in R × (X+ ) such that, if c1 ∈ (c1 , c1 ),
− +
5 Coexistence solutions
We now show the main result of this paper, namely we exhibit coexistence solutions to
∗ 3
the full 2-species system (1.2), i.e. solutions (R, U, V ) to (1.2) that lie in (X+ ) . Recall
that the system with 2 species reads, shortly,
A0 R + c1 f1 (x, R)U + c2 f2 (x, R)V = I,
A1 U − c1 f1 (x, R)U = 0, (5.1)
A2 V − c2 f2 (x, R)V = 0,
The following fact summarizes the work we have performed at this stage.
3
Proposition 5.1 The system (1.2) has the trivial solution (S, 0, 0) ∈ X+ . Besides,
0 ∗ ∗ 3
(i) if c1 > c1 , system (1.2) has the semi-trivial solution (Ru (c1 ), U (c1 ), 0) ∈ X+ .
0 ∗ ∗ 3
(ii) if c2 > c2 , system (1.2) has the semi-trivial solution (Rv (c2 ), 0, V (c2 )) ∈ X+ .
We denote these two families by
∗
Cu = {(c1 , c2 , Ru (c1 ), U ∗ (c1 ), 0), (c1 , c2 ) ∈ (c01 , +∞) × (c02 , +∞)},
Cv = {(c1 , c2 , Rv∗ (c2 ), 0, V ∗ (c2 )), (c1 , c2 ) ∈ (c01 , +∞) × (c02 , +∞)}.
Our first result in the direction of obtaining coexistence solutions to (5.1) is the
27
Point (ii) therefore comes as a direct consequence of the fact that R 7→ f1 (R) in-
creases with R, from which it is deduced that R 7→ λ1 (A1 − c1 f1 (R)) decreases with
∗
R (Lemma 2.1). The function R − Ru (c1 ) cannot have constant sign on Ω, unless it
vanishes identically.
Point (iii).
We use a lower-upper solution method. Whenever u and v belong to X, denote by
R(u, v) the only solution in X to A0 R + A1 u + A2 v = I. With this notation at hand,
the function u = U is seen to satisfy the following, nonlinear, nonlocal, elliptic problem
A1 u − c1 f1 (R(u, V )) u = 0. (5.2)
We first claim that U is the only positive solution to (5.2). To prove this, we observe
that whenever M > 0 is large enough, the constant function u = M is an upper-solution
to (5.2). Indeed, it is clear that R(M, V ) ≥ 0 when M is large (for A0 (R(M, V )) ≤ 0
under these circumstances), from which it follows A1 M −c1 (f1 (R(M, V ))M ≥ m1 M ≥
0. The constant function u = 0 being clearly a lower-solution to (5.2), it follows that
there exist a maximal solution 0 ≤ U + ≤ M such that any solution u to (5.2) such
that 0 ≤ u ≤ M also satisfies 0 ≤ u ≤ U + . In particular, taking M > U , we deduce
0 ≤ U ≤ U +.
To prove that U = U + , we define for convenience R+ = R(U + , V ) and R =
R(U, V ). We clearly have25
Z Z h i
0= A1 U + · U − A1 U · U + = c 1 f1 (R+ ) − f1 (R) U + U,
Ω Ω
A0 r + c1 f1 (r)U + = I − c2 f2 (r)V.
Now, since inf Ω U ∗ (c1 ) > 0, one can choose s ∈ (0, 1) small enough such that sU <
e to (5.2) such that sU < U
U ∗ (c1 ) and it follows that there exists a solution U e < U ∗ (c1 )
(the inequalities being strict because sU and U ∗ (c1 ) are not true solution). Uniqueness
of the positive solution yields U e = U hence U < U ∗ (c1 ).
The same proof shows that V < V ∗ (c2 ). ⊓ ⊔
Proof of Lemma 5.4. Let c1 > c01 be given fixed. We suppose by contradiction that
there exists a sequence of solutions (ck2 , Rk , Uk , Vk ) ∈ (c02 , +∞)×(X+
∗ 3
) with ck2 → +∞.
As in the proof of Lemma 4.1, from the relation A0 Rk + A1 Uk + A2 Vk = I we
deduce that for some α > 0 we have (α − ∆)(a0 Rk + a1 Uk + a2 Vk ) ≤ I (with the
obvious adaptation in the case of variable coefficients ai = ai (x), see the proof of
Lemma 4.1)), hence 0 ≤ a0 Rk + a1 Uk + a2 Vk ≤ M for some M ≥ 0 independent of k.
29
T2 (c2 , R, U, V ) = t (R, U, V ).
∗
We readily know that the semi-trivial solution (c2 , Ru (c1 ), U ∗ (c1 ), 0) satisfies
∗
T2 (c2 , Ru (c1 ), U ∗ (c1 ), 0) = t (Ru
∗
(c1 ), U ∗ (c1 ), 0),
for any value of c2 . We now construct coexistence solutions using bifurcations from
∗
the (family of) point(s) (c∗2 (c1 ), Ru (c1 ), U ∗ (c1 ), 0), where c∗2 (c1 ) > c02 is provided by
Lemma 5.3.
Proposition 5.5 Take c1 > c01 . Let c∗2 = c∗2 (c1 ) > c02 be the eigenvalue defined in
Lemma 5.3 and ψ ∗ = ψ ∗ (c1 ) ∈ X+ ∗
be the associated eigenfunction.
∗ ∗ ∗
Then (c2 (c1 ), Ru (c1 ), U (c1 ), 0) is a bifurcation point for T2 , in that the local bi-
furcation Theorem 2.9 applies.
In particular, there exists ρ∗ = ρ∗ (c1 ) ∈ X and ∗ ∗
φ = φ (c1 ) ∈ X, there exists
1 3
ε > 0, there exists a map (re, u e, ve) ∈ C (−ε, ε), X verifying re(0) = u e(0) = ve(0) = 0,
together with a map c2 ∈ C 1 (−ε, ε), R+ verifying c2 (0) = c∗2 (c1 ), such that the
following holds. The branch
n o
e
c2 (s), R(s), e (s), Ve (s) ; 0 < s < ε
U
30
a multiplicative constant. Equation (5.7) on (ρ, φ), together with the already proved
∗
invertibility of Id − D(R,U ) T1 (c1 , Ru (c1 ), U ∗ (c1 )) (see Proposition 4.10), then provides
∗ ∗
(ρ, φ) = (ρ (c1 ), φ (c1 )), where we have set
t
(ρ∗ (c1 ), φ∗ (c1 )) := (5.9)
∗ −1 t
Id − D(R,U ) T1 (c1 , Ru (c1 ), U ∗ (c1 )) c∗2 (c1 )K0 ∗
f2 (Ru (c1 ))ψ ∗ (c1 ) , 0 .
Hence Ker (L2 (c∗2 (c1 ))) = span(ρ∗ (c1 ), φ∗ (c1 ), ψ ∗ (c1 )) and dim (Ker (L2 (c∗2 (c1 )))) = 1.
The Fredholm alternative also provides codim (Im (L2 (c∗2 (c1 )))) = 1.
There remains to show that
Dc2 L2 (c∗2 (c1 )) · t (ρ∗ (c1 ), φ∗ (c1 ), ψ ∗ (c1 )) ∈
/ Im L2 (c∗2 (c1 )) . (5.10)
We clearly have
Dc2 L2 (c∗2 (c1 )) · t (ρ∗ (c1 ), φ∗ (c1 ), ψ ∗ (c1 )) =
∗
t
(−K0 (f2 (Ru (c1 ))ψ ∗ (c1 )), 0, −K2 (f2 (Ru
∗
(c1 ))ψ ∗ (c1 ))).
If relation (5.10) is false, we can find ψ1 such that
∗
−K2 (f2 (Ru ∗
(c1 ))ψ ∗ (c1 )) = ψ1 − c∗2 (c1 ) K2 (f2 (Ru (c1 ))ψ1 ).
Z
∗ 2
As in the proof of Proposition 4.7, we get f2 (Ru (c1 )) ψ ∗ (c1 ) = 0, which contra-
∗
dicts ψ (c1 ) > 0.
Eventually we have proved that the local bifurcation Theorem 2.9 applies, and the
Proposition follows. ⊓
⊔
31
Proposition 5.6 Let c1 > c01 be fixed. Then, equation (5.1) admits a continuum of
nontrivial solutions
Proof of Proposition 5.6. We apply the global bifurcation Theorem 2.10. It suffices to
∗
show that i(T2 (c2 , ·), (Ru (c1 ), U ∗ (c1 ), 0)) actually changes sign when crossing the value
∗
c2 = c2 (c1 ).
∗
Let µ > 1 be an eigenvalue of D(R,U,V ) T2 (c2 , Ru (c1 ), U ∗ (c1 ), 0) = Id − L2 (c2 ).
There exists (ρ, φ, ψ) 6≡ (0, 0, 0) such that (Id − L2 (c2 )) t (ρ, φ, ψ) = µ t (ρ, φ, ψ).
If ψ = 0, we recover, using relation (5.6), that
∗
Id − D(R,U ) T1 (c1 , Ru (c1 ), U ∗ (c1 )) t
(ρ, φ) = µ t (ρ, φ).
∗
Hence µ > 1 is an eigenvalue of Id − D(R,U ) T1 (c1 , Ru (c1 ), U ∗ (c1 )) . We know from
Proposition 4.12 that such µ’s are in even number.
∗
If ψ 6= 0, we recover using relation (5.6), that A2 ψ − c2 f2 (Ru (c1 ))ψ = (1 − µ)A2 ψ,
which means,
c
A2 ψ − 2 f2 (Ru ∗
(c1 ))ψ = 0. (5.11)
µ
Thanks to Lemma 5.3, it becomes clear that the above problem has no nontrivial
solution ψ 6≡ 0 whenever c2 ≤ c∗2 (c1 ), while it has exactly one nontrivial solution (up
to a multiplicative constant), namely ψ ∗ (c1 ), whenever c2 > c∗2 (c1 ) is close enough to
c∗2 (c1 ). This establishes
∗
i(T2 (c2 , ·), (Ru (c1 ), U ∗ (c1 ), 0)) = 1, if c2 < c∗2 (c1 ),
∗
i(T2 (c2 , ·), (Ru (c1 ), U ∗ (c1 ), 0)) = −1, if c2 > c∗2 (c1 ).
an alternative similar to the one satisfied by C0 , namely, one of the three following
situations occur:
∗
(c2 , R, U, V ) = (c2 , Ru (c1 ) + r, U ∗ (c1 ) + u, v),
∗
with (r, u, v) 6= (0, 0, 0), the symmetric point (c2 , Ru (c1 )) − r, U ∗ (c1 ) − u, −v)
+
belongs to C0 as well.
In the present contradiction argument, case (iii) cannot occur, nor can case (i)
occur. On top of that, take a point (c2 , R, U, V ) ∈ C0+ . Lemma 5.4 asserts that we
necessarily have c02 < c2 < cmax
2 (c1 ). Hence c2 remains in a fixed bounded subset of
R. Besides, the proof of Lemma 5.4 also asserts that (R, U, V ) necessarily belong to a
fixed compact subset of X 3 . Hence situation (ii) cannot occur.
This ends the proof. ⊓⊔
The above proposition asserts that C0+ necessarily leaves the positive cone. The
following Lemma provides information on the points where C0+ leaves the positive cone.
Lemma 5.8 Take c1 > c01 . Let (c2 , R, U, V ) ∈ R × (X+ )3 be the limit, in R × X 3 , of
∗ 3
a sequence of positive solutions (ck2 , Rk , Uk , Vk ) ∈ R × (X+ ) to (5.1). Then, we have
∗
C0+ joins (c∗2 (c1 ), Ru (c1 ), U ∗ (c1 ), 0)
to (c∗∗ ∗ ∗∗ ∗ ∗∗
2 (c1 ), Rv (c2 (c1 )), 0, V (c2 (c1 ))).
33
Theorem 2.14 states that two families of coexistence solutions may be obtained, namely
the first one is constructed by freezing c1 > c01 and seeing c2 as a bifurcation parameter
to bifurcate from the semi-trivial (c∗2 (c1 ), Ru∗
(c1 ), U ∗ (c1 ), 0) where c∗2 (c1 ) > c02 , while
the second one is constructed by freezing c2 > c02 and seeing c1 as a bifurcation param-
eter to bifurcate from the semi-trivial (c∗1 (c2 ), Rv∗ (c1 ), 0, V ∗ (c2 )) where c∗1 (c2 ) > c01 .
This construction leads to defining the quantities c∗∗ 0 ∗∗ 0
2 (c1 ) > c2 and c1 (c2 ) > c1 . Note
∗∗ ∗ ∗∗ ∗ ∗∗ ∗
that the three situations c2 (c1 ) > c2 (c1 ), c2 (c1 ) < c2 (c1 ), c2 (c1 ) = c2 (c1 ) may very
well occur, and similarly for c∗∗ ∗
1 (c2 ) and c1 (c2 ).
Let us now exhibit some properties of the c∗i (cj )’s and c∗∗ i (cj )’s.
Lemma 5.10 For each c1 > c01 and c2 > c02 , we define
µ(c1 , c2 ) := λ1 (A1 − c1 f1 (Rv∗ (c2 ))), ∗
ν(c1 , c2 ) := λ1 (A2 − c2 f2 (Ru (c1 ))).
We have the relation (where sgn(s) = +1 if s > 0, = −1 if s < 0 and = 0 if s = 0)
sgn (µ(c1 , c2 )) = sgn c∗1 (c2 ) − c1 = −sgn c∗∗
2 (c1 ) − c2 ,
sgn (ν(c1 , c2 )) = sgn c∗2 (c1 ) − c2 = −sgn c∗∗
1 (c2 ) − c1 .
µ(c1 , c2 ) is increasing. This shows that sgn (µ(c1 , c2 )) = sgn (c∗1 (c2 ) − c1 ).
Take c1 > c01 . The construction of the point (c∗∗ ∗ ∗∗ ∗ ∗∗
2 (c1 ), Rv (c2 (c1 )), 0, V (c2 (c1 ))),
together with Lemma 5.8, provide µ(c1 , c∗∗ 2 (c 1 )) = λ 1 (A 1 − c f
1 1 (R ∗ ∗∗
(c
v 2 (c 1 )))) = 0.
By Theorem 2.12, the map c2 7→ Rv∗ (c2 ) is decreasing, hence by Lemma 2.1, the map
c2 7→ µ(c1 , c2 ) is increasing. This shows sgn (µ(c1 , c2 )) = −sgn (c∗∗ 2 (c1 ) − c2 ). ⊓ ⊔
34
Proposition 5.11 Let be {i, j} = {1, 2}. For all cj > c0j , the scalar c∗∗
i (cj ) is charac-
terized by 26
c∗i (c∗∗
j (ci )) = ci .
Proposition 5.12 (i) The function c1 7→ c∗2 (c1 ) is continuous and increasing from
(c01 , +∞) to (c02 , +∞). The similar statement holds for c2 7→ c∗1 (c2 ).
(ii) We have lim c∗2 (c1 ) = +∞ and lim c∗1 (c2 ) = +∞.
c1 →∞ c2 →∞
(iii) We have lim c∗2 (c1 ) = c02 and lim c∗1 (c2 ) = c01 .
c1 →c01 c2 →c02
At this level of the analysis, one may define the three open sets
Θ+ = {c1 , c2 ) ∈ (c01 , +∞) × (c02 , +∞), c∗i (cj ) < ci < c∗∗
i (cj ), i 6= j},
Θ+ = {c1 , c2 ) ∈ (c01 , +∞) × (c02 , +∞), c1 > c∗2 (c1 ), c2 > c∗1 (c2 )},
Θ− = {c1 , c2 ) ∈ (c01 , +∞) × (c02 , +∞), c1 < c∗2 (c1 ), c2 < c∗1 (c2 )}.
6.1 A conjecture
Conjecture
(i) If (c1 , c2 ) ∈ e then there cannot exist (R, U, V ) ∈ (X+
/ Θ, ∗ 3
) solution to (5.1).
(ii) We have Θ− = ∅, or, in other words, c∗i (cj ) ≤ c∗∗ i (c j whenever i 6= j.
)
This conjecture is motivated by our numerical simulations. It states that the set
e actually characterizes those values of (c1 , c2 ) for which a coexistence solution may
Θ
be exhibited. It also states that species i survives if and only if ci ≤ c∗i (cj ). In other
words, species i survives if and only if λ1 (Ai − ci fi (R∗ (cj )) ≥ 0.
Provided the above conjecture holds, this assertion implies that as the diffusion
rate of a given species increases, its ability to survive decreases.
In other words, the coexistence domain Θ is embedded in the set of the (c1 , c2 )
∗
such that Ru (c1 ) − Rv∗ (c2 ) is neither positive nor negative. This point highlights the
importance of the spatial heterogeneity in the coexistence process. This point in further
discussed in the next subsection.
36
In the homogeneous case where the functions I(x), fi (x), mi (x), ai (x) do not depend on
∗
x, and when Neumann boundary conditions are retained, we have that Ru (c1 )(x) and
∗
Rv (c2 )(x) are constant functions. Hence, by Proposition 6.2, the coexistence is possible
∗ ∗
only if Ru (c2 ) = Ru (c1 ), which induces a degenerate solution. Moreover, the fact that
∗ ∗
Ru (c1 ) and Rv (c2 ) decrease imply that meas{(c1 , c2 ) ∈ (c01 , +∞) × (c02 , +∞), Ru ∗
(c2 ) =
∗
Ru (c1 )} = 0. In that degenerate case we have the
Θ = ∅,
n o
e = (c1 , c2 ) ∈ (c01 , +∞) × (c02 , +∞) s.t. R1∗ (c1 ) = R2∗ (c2 ) < S
Θ
m1 m2
= (c1 , c2 ) ∈ (c01 , +∞) × (c02 , +∞) s.t. f1−1 = f2−1 <S .
c1 c2
Another critical case appears when the two species possess heterogeneous but pro-
portional diffusion rates, mortality rate, and consumption rate, namely
∗ 3
Moreover, the system has a coexistence solution (R, U, V ) ∈ (X+ ) if and only if
e ∗ ∗ ∗
(c1 , c2 ) ∈ Θ. In that case {(Ru , tU , (1 − t)U ), t ∈ (0, 1)} is a family of solutions
∗
and each coexistence solution satisfies (R, U, V ) ∈ {(Ru , tU ∗ , (1 − t)U ∗ ), t ∈ (0, 1)}.
∗
Proof of Proposition 6.4. The system defining (Ru (c1 ), U ∗ (c1 )) is
A1 U ∗ (c1 ) − c1 f1 (Ru
∗
(c1 ))U ∗ (c1 ) = 0, ∗
A0 Ru (c1 ) + A1 U ∗ (c1 ) = 0,
while the system defining (Rv∗ (c2 ), V ∗ (c2 )) is in the present case
αA1 V ∗ (c2 ) − c2 f1 (Rv∗ (c2 ))V ∗ (c1 ) = 0, A0 Rv∗ (c2 ) + αA1 V ∗ (c2 ) = 0.
1
c
The uniqueness result of Proposition 4.4 provides V ∗ (c2 ) = U ∗ 2 , and Rv∗ (c2 ) =
α α
∗ c2
Ru . Now, since c∗2 (c1 ) is defined as the unique value of the parameter c2 such that
α ∗ ∗
λ1 (A2 − c2 f2 (Ru (c1 ))) = 0, i.e. λ1 (A1 − (c2 /α)f1 (Ru (c1 ))) = 0, it comes c∗2 (c1 ) = αc1 .
∗
This together with the analogous relation for c1 (c2 ) provides
Θ = ∅, e = {(c1 , αc1 ) ; c1 > c01 }.
Θ
Take now (c1 , c2 ) such that (R, U, V ) is an associated coexistence solution. We have
c
λ1 (A1 − 2 f1 (R)) = λ1 (A1 − c1 f1 (R)) = 0,
α
and monotone dependence of the above λ1 ’s with the parameters c1 and c2 implies
c2 = αc1 . Besides, summing the last two equations of (1.2) leads to
A0 R + c1 f1 (R)(U + αV ) = I,
(A1 − c1 f1 (R))(U + αV ) = 0,
so that uniqueness provides U +αV = U ∗ (c1 ), and R = Ru ∗
(c1 ). On top of that, coming
back to the equations satisfied by U resp. V , it appears that there exists (t, y) ∈ R2+
such that U = tU ∗ (c1 ), and αV = yU ∗ (c1 ). Gathering the relations then provides the
necessary equation t + y = 1. This ends the proof. ⊓ ⊔
with Neumann boundary condition27 It can be shown (see [25]), using the central
manifold theorem, that the solution to (6.1)converges to the solution of the so-called
aggregated system
m f0 r + c1 fe1 (r)u + c2 fe2 (r)v = Ie
(mf − c1 fe1 (r))u = 0 (6.2)
1
(mf2 − c2 fe2 (r))v = 0
27 This is the only place in this text where Neumann – and not Robin – boundary conditions
are required
38
R R R
where m 1
fi = |Ω| Ω
mi (x)dx, fei (r) = |Ω|
1
f (x, r), Ie = |Ω|
Ω i
1
Ω
I(x)dx, and the un-
known r, u, v now are scalars (independent of x). System (6.2) is a homogeneous
chemostat system. Therefore, and as is easily seen on the equations, in the generic case
there are no positive solution to (6.2). As it is proved in [25], it turns out that for
ε > 0 small enough, the original system (6.1) has no positive solution in the generic
case neither.
This result allows to describe the behavior of the coexistence domain Θ when the
diffusion rates tend to +∞. Remark in passing that, if Assumption 2 is true for a given
ε > 0, then it remains true for each ε > 0. In this case, Theorem 2.16 shows that there
exists Θε ⊂ R2+ such that, for each (c1 , c2 ) ∈ Θε , the system (6.1) admits a coexistence
solution. The boundaries of Θ fε are given by the curves
We have the
Proposition 6.5 Denote Θ∞ = {(c1 , c2 ) s.t. r1∗ (c1 ) = r2∗ (c2 ) < +∞}.
/ Θ∞ , there exists ε0 > 0 such that ∀ε ∈ (0, ε0 ), we have
Then, for each (c1 , c2 ) ∈
ε
(c1 , c2 ) ∈
/Θ .
This study examines a model where two species compete for a single resource, in a spa-
tially heterogeneous domain. Our system differs from the classical unstirred chemostat
system [13, 12, 11, 3, 18] in that the the reaction terms do depend on space, and, more
importantly, we allow the diffusion rates to depend on the species under consideration.
This point leads to a new mathematical difficulty. Namely, the conservation law which
links the resource R with the two species U and V , noted A0 R + A1 U + A2 V = I
in the core of the paper, becomes a nonlocal equation (as compared to the previously
quoted papers where the analogous equation is local). We circumvent this difficulty by
introducing Assumption 2 (supplemented with Assumption (1.6) in the case of Robin
boundary conditions).
We show that coexistence occurs when the consumption parameters (c1 , c2 ) lie in
a subdomain Θ ⊂ R2+ . In addition, we study the set Θ by using a characterisation of
Θ that relies on the two functions c∗1 (c2 ) and c∗2 (c1 ) defined in the text.
Several direction may extend this study. Firstly, our numerical observations indicate
that the coexistence solution are non-degenerate, except in the particular case when the
39
two functions c∗1 (.) and c∗2 (.) coincide. When the coexistence solution is non-degenerate,
it turns out that our construction can be extended to three species, and by iteration,
to N species for any value of N . It would therefore be a key step to actually prove that
the coexistence solutions necessarily are non-degenerate, unless c∗1 (.) and c∗2 (.) coincide
Note in passing that Propositions 6.3 and 6.4 give two examples of situations where
c∗1 (.) and c∗2 (.) coincide, and a complete description of the coexistence phenomena is
provided in these situations.
Secondly, we defined Θ as the union of two subdomain Θ− and Θ+ . If (c1 , c2 ) ∈ Θ−
then c∗i > c∗∗i and the bifurcation occurs ”to the left” (see Figure 2.2). We conjecture
that Θ− = ∅ in any case. In fact, to rephrase our conjecture, if (c1 , c2 ) ∈ Θ− , then
both species are ”not invasive” in the sense that
Note that Waltmann et al. [11] formulate a similar conjecture. Namely, they conjecture
that a necessary condition for two species to coexist is that both species are ”invasive”
in the sense that λ1 (A1 − c1 f1 (Rv∗ (c2 )) ≥ 0 and λ1 (A2 − c2 f2 (Ru
∗
(c1 )) ≥ 0. Lastly,
note that if (c1 , c2 ) ∈ Θ− , then the index of both semi-trivial solutions is equal to
1. To rephrase the above considerations, Waltmann et al. in [11] conjecture that a
necessary condition for two species to coexist is that both sem-trivial solution are
unstable (for the time-dependent problem), which in our case, means that the index
of the two semi-trivial solutions is equal to −1. Note that even if the latter result is
proved, it is not clear that the coexistence solution itself is stable. Indeed, Hofbauer
and So [17] show that there exists gradostats (that is, similar models with a discrete
spatial structuration) for which an unstable coexistence solution may be exhibited. A
more precise description of Θ would be a first step to understand the situation.
Thirdly, we conjecture that if (c1 , c2 ) ∈ e then no coexistence solution can be
/ Θ,
found. Would this result be proved, we could use Θ as a geometrical indicator of the
possibility of coexistence in a given system. Numerical investigations on the relation
between Θ, spatial heterogeneity, and the biodiversity, will be published soon.
Finally, our proof uses basically Assumption 2, an assumption that allows us to
extend the analysis of the (known) case where all diffusion operators coincide. It is
to be noted, however, that our approach proves the existence of semi-trivial solutions
without using Assumption 2. This assumption is only needed to obtain uniqueness and
non-degeneracy of the so-obtained semi-trivial solutions. A natural question is: can one
extend our construction to situations where Assumption 2 is not verified?
References
7. C.V. Pao, On nonlinear reaction-diffusion equations, J. Math. Analysis applic, 87, 1982,
165-198.
8. C. Walker, Coexistence Steady States in a Predator-Prey Model, Arch. Math., 95, 2010,
87-99.
9. C. Walker, Global Bifurcation of Positive Equilibria in Nonlinear Population Models, J.
Differential Equations, 248, 2010, 1756-1776.
10. Z. Zhang, Coexistence and Stability of Solutions for a Class of Reaction-Diffusion Systems,
E. J . Diff. Eq., 137, 2005, 1-16.
11. S. B. Hsu and P. Waltman, On a system of reaction-diffusion equations arising from
competition in an unstirred chemostat., SIAM J. Appl. Math, 53, 1993, 1026-1044.
12. S.B. Hsu, H. Smith and P. Waltman, Dynamic of competition in the unstirred chemostat.,
Can. Appl. Math. Quart., 2, 1994, 461-483.
13. J. L. Dung, H. L. Smith and P. Waltman,Growth in the unstirred chemostat with different
diffusion rates, Fields institute communications, 21, 1999, 131-142.
14. S. Zheng and J. Liu, Coexistence solutions for a reaction-diffusion system of un-sirred
chemostat model, Applied Math. and Comp., 145, 2003, 579-590.
15. S. Zheng and H. Guo and J. Liu, A food chain model for two resources in unstirred
chemostat, Applied Math. Comp., 206, 2008, 389-402.
16. J. V. Baxley and S. B. Robinson, Coexistence in the unstirred chemosat,
Applied Math. And Comput., 39, 1998, 41-65.
17. J. Hofbauer and J. W. H. So, Competition in the gradostat: The global stability problem,
Nonlinear analysis, 22, 1994, 1017-1033.
18. J. H. Wu, Global bifurcation of coexistence state for the competition model in the chemo-
stat, Nonlinear Analysis, 39, 2000, 817-835.
19. H. Nie and J. Wu: Uniqueness and stability for coexistence solutions of the unstirred
chemostat model, Applicable Analysis 89, 1151–1159 (2010)
20. J. Blat and K. J. Brown, Bifurcation of steady-state solutions in predator-prey and com-
petition systems, Proc. Roy. Soc. Edin.A, 97, 1984, 21-34.
21. J. Blat and K. J. Brown, Global bifurcation of positive solutions in some systems of elliptic
equations, SIAM J. Math. Anal., 17, 1986, 1339-1353.
22. Y. Du end K. J. Brown,Bifurcation and Monotonicity in Competition Reaction-Diffusion
Systems, Nonlinear Ana. Th. Meth. & Appl., 23 No. 1, 1994, 1-13.
23. E. D. Conway, Diffusion and The Predator-Prey Interaction: Steady States with Flux at
the Boundaries, Contemporary. Math., 17, 1983, 215-234.
24. E. Dancer, On positive solutions of some partial differential equations, Trans. Amer. Math.
Soc., 284, 1984,
25. S. Madec and F. Castella, Global behavior of N competiting species with strong diffusion:
diffusion leads to exclusion, to appear