S.Jimbo JDE 2013
S.Jimbo JDE 2013
a r t i c l e i n f o a b s t r a c t
1. Introduction
* Corresponding author.
E-mail addresses: jimbo@math.sci.hokudai.ac.jp (S. Jimbo), morita@rins.ryukoku.ac.jp (Y. Morita).
0022-0396/$ – see front matter © 2013 Elsevier Inc. All rights reserved.
http://dx.doi.org/10.1016/j.jde.2013.05.021
1658 S. Jimbo, Y. Morita / J. Differential Equations 255 (2013) 1657–1683
∂u ∂u
= = 0 (x ∈ ∂Ω), (1.2)
∂ν ∂ν
where d, k are positive constants, g (u , v ) is sufficiently smooth function and Ω is a bounded domain
in Rn with the smooth boundary ∂Ω .
We may assume
g (s)
g (s)/s K (s > 0), lim = β > 0, g ∈ C 1 [0, ∞); R , (1.3)
s→+0 s
which ensure the nonnegativity of the solution of the initial value problem provided u (x, 0),
v (x, 0) 0 in Ω by the maximum principle applied to each equation of the system.
Since the total mass
u (x, t ) + v (x, t ) dx
Ω
is conserved in time evolution for the solution (u (x, t ), v (x, t )) of (1.1), we fix the condition as
s := u (·, 0) + v (·, 0) , (1.4)
kw
g(w) =
(aw + 1)2
in [14] and [20] as a simple model exhibiting similar dynamical behaviors to those observed in a
more complicated model describing the cell polarity. In fact, for such a function g they show the
Turing type instability for appropriate parameter values d, k and s. They also exhibit by a numerical
simulation (under the periodic boundary conditions) that after an emergence of a wave pattern and
a long transient behavior, the solution converges to a stationary solution with the shape of a single
spike. By virtue of this concentration property their model equations could be helpful for the under-
standing of the phenomenon of cell polarity. However, they don’t show why such a spike pattern is
stable while the multi-spike pattern is unstable.
The purpose of this article is to reveal some dynamical property and the stability of equilibrium
solution of (1.1)–(1.2), which justifies the numerical observation about the asymptotic profile of the
solution. More precisely, we show that the system of (1.1) allows a Lyapunov function, by which we
can assert that the ω -limit set of any bounded orbit of the solution consists of equilibrium solutions.
In addition, by investigating the linearized eigenvalue problem for any equilibrium solution, we see
that the profile of a stable equilibrium solution in one-dimensional space must be monotone (or
unimodal, i.e., of a single spike if the periodic boundary conditions are assumed).
We remark that for f (u ) instead of g (u + v ) the similar dynamical property is also found in
the same literatures [14,20]. For some mathematical result justifying it is proposed in [19,18] by
converting the equations to the equations of the phase field type system as in [4] and [9]. The present
article is thereby a successive work for a class of conserved reaction–diffusion systems.
S. Jimbo, Y. Morita / J. Differential Equations 255 (2013) 1657–1683 1659
ut = du xx − f (u , v ), v t = v xx + f (u , v ),
which arises as a simple chemical conserved reaction model, called a closed reaction–diffusion system.
The existence of traveling waves for a specific nonlinear term is shown in [5] and [8]. Moreover, this
type of the system is used in modeling of a precipitation kinetics, and the formation of Liesegang
patterns is shown in [13] and [23]. As seen in [21], by putting w = u + v , v = v, the above system
can be converted to the one belonging to a class of the systems
w t = a (u , v ) w x + b (u , v ) v x x , v t = v xx + f̃ (u , w ).
Under some conditions on a(u , v ) and b(u , v ) the instability of a spike solution of the system in the
infinite interval or a sufficiently large interval with the Neumann boundary conditions is investigated
in [21] (see also [22] on the stability of layer solutions).
Coming back to the present system, we state our results precisely. We first provide the global
boundedness of the solution for t 0 under the condition (1.3).
Proposition 1.1. Assume (1.3) and that u 0 , v 0 ∈ C 0 (Ω, R), u 0 (x), v 0 (x) 0. Let (u (x, t ), v (x, t )) be a solu-
tion of (1.1) with (1.2) and (u (x, 0), v (x, 0)) = (u 0 (x), v 0 (x)). Then u (·, t ) L ∞ , v (·, t ) L ∞ ,
∂ u (·, t )/∂ x j L ∞ , ∂ v (·, t )/∂ x j L ∞ ( j = 1, 2, . . .) are uniformly bounded in t 0. Moreover, there is a posi-
tive M > 0, independent of the initial data, such that
n
∂u ∂v
lim sup u (·, t ) L∞
+ v (·, t ) L∞
+ (·, t ) + (·, t ) Ms
t →∞ ∂xj L∞ ∂xj L∞
j =1
holds.
Next we show the system allows a Lyapunov function. Define the functional
d
E (u , v ) := ∇(u + v )2 + 1 ∇(du + v )2 + Q (u + v ) dx, (1.5)
2 2
Ω
w
dk
Q ( w ) := (1 − d) g ( z) dz + w 2.
2
Then we have
Proposition 1.2. For a smooth solution (u (·, t ), v (·, t )) (0 t < ∞) to (1.1) with (1.2) (or the periodic bound-
ary conditions)
d 2
E u (·, t ), v (·, t ) = − (1 + d)(ut + v t )2 + k∇(du + v ) dx 0 (1.6)
dt
Ω
holds. Then dE (u (·, t ), v (·, t ))/dt = 0 (∀t ∈ R) implies that (u (x, t ), v (x, t )) is an equilibrium solution.
(see for instance [11] or [12]). In other words, the asymptotic state of any solution to (1.1) with (1.2)
and (1.3) is determined by the stationary problem.
Before going to the stationary problem, we covert the system by putting w = u + v as
w t = d w + (1 − d) v ,
(x ∈ Ω) (1.7)
v t = v + g ( w ) − kv
w = w + W = s+ W,
(1.8)
v = v + V =ξ + V.
W t = d W + (1 − d) V , (1.9)
V t = V + g (s + W ) − g (s + W ) − kV , (1.10)
ξt := g (s + W ) − kξ (1.11)
with
V (·, t ) = W (·, t ) = 0, (1.12)
namely the solutions ( w (x, t ), v (x, t )) to (1.7) and ( W (x, t ), V (x, t ), ξ(t )) to (1.10)–(1.11) have the 1–1
correspondence by
w (x, t ), v (x, t ) = s + W (x, t ), ξ(t ) + V (x, t ) , ξ(t ) = v (·, t ) ,
consequently, the solution (u (x, t ), v (x, t )) of (1.1) with (1.2) and (1.4) is given by
u (x, t ), v (x, t ) = s − ξ(t ) + W (x, t ) − V (x, t ), ξ(t ) + V (x, t ) .
We notice that the equations for W and V of (1.9) and (1.10) are independent of ξ , therefore it
suffices to consider (1.9) and (1.10) with the Neumann conditions and (1.12) unless d = 1. If d = 1,
then the equations turn to be quite simple, we thereby exclude this case in the present study.
Now consider the stationary problem, that is,
d W + (1 − d) V = 0,
(x ∈ Ω), (1.13)
V + g (s + W ) − g (s + W ) − kV = 0
with
∂W ∂V
= = 0 (x ∈ Ω), W = V = 0. (1.14)
∂ν ∂ν
dW + (1 − d) V = 0
S. Jimbo, Y. Morita / J. Differential Equations 255 (2013) 1657–1683 1661
follows. Hence, we see that (1.13) with (1.14) is equivalent to the next scalar equation
W − α g (s + W ) − g (s + W ) − kW = 0 (x ∈ Ω), (1.15)
∂W
= 0 (x ∈ ∂Ω), W = 0, (1.16)
∂ν
where we put
1−d
α := (d = 1).
d
1 1 1 d
u (x), v (x) = s − g s + W ∗ +
∗ ∗
W (x), g s + W ∗ −
∗
W ∗ (x) .
k 1−d k 1−d
The next result is on the spectrum comparison between the linearized operator at an equilibrium
solution in the system and that of the above scalar equation (1.15)–(1.16). The idea of the comparison
is due to [3] for the Cahn–Hilliard equation and the phase field system. Later, some improvement for
the phase field system is proposed in [18]. We will develop the study to our system. To state the
result, we introduce function spaces. Denote by L 2 (Ω) the space of square integrable functions with
the norm u (u ∈ L 2 (Ω)) and
1/2
2
H m (Ω) = u ∈ L 2 (Ω): u H m := u 2 + ∂ a u /∂ xa <∞ ,
|a|m
n
a = (a1 , . . . , an ), a j 0, a j ∈ Z, |a| = a j.
j =1
We also introduce a subspace of L 2 (Ω) and H m (Ω) with the average zero as
2 2
L (Ω) := u ∈ L (Ω): u dx = 0 , H m (Ω) := u ∈ H m (Ω): u ∈ L 2 (Ω) .
Ω
We let W ∗ be a solution of (1.15) with (1.16) and put w ∗ = s + W ∗ . We fix s and consider the
linearized operator A for (1.7)
φ dφ + (1 − d)ψ
A := − , (1.17)
ψ ψ + g ( w ∗ )φ − kψ
L(φ) := −φ + α g w ∗ φ − g w ∗ φ + kφ, (1.18)
D(A) = (φ, ψ) ∈ H 2 (Ω) × H 2 (Ω): ∂φ/∂ ν = ∂ψ/∂ ν = 0 (x ∈ ∂Ω) ,
and
D(L) = φ ∈ H 2 (Ω): ∂φ/∂ ν = 0 (x ∈ ∂Ω) ,
respectively.
We compare the spectrum of A and L. It is easy to see that any eigenvalue is real for L. We let
{μ j } j =1,...,∞ be the set of all the eigenvalues of L arranged in increasing order with counting the
multiplicity as
μ1 μ2 · · · μm · · · . (1.19)
On the other hand there is a possibility A allows a complex eigenvalue. Fortunately, any eigenvalue λ
of A is real if the real part of λ is less that k/2 (see Section 4), in the sequel it is possible to compare
all the real eigenvalues which are related to the stability of the solutions. We thereby let {λ j } j =1,..., N
be the set of all the eigenvalues of A less than k/2, arranged in increasing order with counting the
multiplicity as
The next theorem gives the spectrum comparison between the linearized operators.
Theorem 1.3. The number of the negative eigenvalues of A and L coincides, and so the multiplicity of the zero
eigenvalue of both the operators. Moreover,
d
|λ j | |μ j | ( j = 1 , 2 , . . . , N ) (1.21)
1+d
in the space H 1 (Ω), where Q is as in (1.5). By virtue of Theorem 1.3, noticing (L(φ), φ) L 2 is the
second variation of G ( w ), we can apply the stability result for (1.22) in [10] to the system (1.1) to
obtain the following result (see also [24], and as for the periodic boundary conditions see [18]):
Corollary 1.4. Consider the one-dimensional case, namely, the case that Ω is interval. Let (u (x), v (x)) be a
stable equilibrium solution to (1.1) with (1.2) and (1.4). Then u (x) and v (x) are monotone, namely, those are
constant or strictly monotone. If the boundary conditions are periodic, then u (x) and v (x) are constant or
unimodal.
We use the terminology “stable” of the above corollary in the sense of Lyapunov (see [12]). We
note that the linearized operator always possesses zero eigenvalue.
We state the idea of the proof of Theorem 1.3 in brief. Noticing that the equations of the eigen-
value problem of A can be written in a single equation for φ , we introduce a nonlocally weighted
eigenvalue problem for L with a free parameter β 0 and establish the comparison between the
eigenvalues of the nonlocal problem and the original L, which is the case β = 0 up to multiplication
S. Jimbo, Y. Morita / J. Differential Equations 255 (2013) 1657–1683 1663
of the constant d/(1 + d). By the continuity property of the spectra in β each eigenvalue of A less
than k/2 is given by choosing an appropriate β > 0. By this we obtain the desired assertion.
In the next section we prove Propositions 1.1 and 1.2. Then in Section 3 we discuss the nonlocally
weighted eigenvalue problem and prove Theorem 1.3 by the key lemmas, which will be proved in
Section 4.
In Section 5 we consider a modified system which is given by g (u + γ v ) with the parameter
γ ∈ [0, 1], instead of g (u + v ). This model system covers the present case. We show that the new
system also possesses a Lyapunov function for any γ ∈ [0, 1]. We conclude the paper by Appendices A
and B in which it is shown that the Turing type instability certainly takes place for a specific g =
kw /(aw + 1)2 , and the proof of Lemma 2.2 is given.
ut = du − f (u , v ), v t = v + f (u , v ).
It is known that the solutions u (x, t ) and v (x, t ) are nonnegative and smooth in Ω × (0, ∞). We first
prove that there are c 0 , M 0 > 0 such that
Since the solutions p (t ) and q(t ) have at most the exponential growth, we obtain (2.1).
Before going to the proof of the global boundedness of the solutions, we prepare two lemmas.
1664 S. Jimbo, Y. Morita / J. Differential Equations 255 (2013) 1657–1683
φL ∞ (Ω) 1/|Ω| φ L 1 (Ω) + K (Ω)∇φL ∞ (Ω) φ ∈ C 1 (Ω; R)
holds.
Proof. From the regularity of ∂Ω , there exist a constant c > 0 and a curve γ : [0, 1] → Ω with the
conditions
γ (0) = x, γ (1) = y , γ̇ (τ ) c (0 τ 1)
1 1
d
φ(x) − φ( y ) = φ γ (τ ) dτ = ∇φ γ (τ ) · γ̇ (τ ) dτ .
dτ
0 0
Hence, we have
φ(x) φ( y ) + c ∇φL ∞ (Ω) .
Integrate both sides of this inequality in y, we easily see the desired inequality follows. 2
Next, let G 1 (x, y , t ) and G 2 (x, y , t ) be the heat kernels for d and with the Neumann boundary
condition respectively. We have the next estimate.
c1
∇x G j (x, y , t ) exp −η1 t − |x − y |2 /η2 t (x, y ∈ Ω, t > 0, j = 1, 2)
t (n+1)/2
u (x, t ) = G 1 (x, y , t )u 0 ( y ) dy
Ω
t
− G 1 (x, y , t − τ ) f u ( y , τ ), v ( y , τ ) dy dτ , (2.2)
0 Ω
v (x, t ) = G 2 (x, y , t ) v 0 ( y )
Ω
t
+ G 2 (x, y , t − τ ) f u ( y , τ ), v ( y , τ ) dy dτ . (2.3)
0 Ω
S. Jimbo, Y. Morita / J. Differential Equations 255 (2013) 1657–1683 1665
Given a small constant δ ∈ (0, 1), we consider the solutions for t δ and estimate u (t , x), v (t , x) in
[δ, ∞) × Ω . From (2.2) and (2.3) we see
t −δ
∇ u (t , x) = ∇ G 1 (x, y , t )u 0 ( y ) dy − ∇ G 1 (x, y , t − τ ) f (u , v ) dy dτ
Ω 0 Ω
t
− ∇ G 1 (x, y , t − τ ) f (u , v ) dy dτ ,
t −δ Ω
t −δ
∇ v (t , x) = ∇ G 2 (x, y , t ) v 0 ( y ) dy + ∇ G 2 (x, y , t − τ ) f (u , v ) dy dτ
Ω 0 Ω
t
+ ∇ G 2 (x, y , t − τ ) f (u , v ) dy dτ ,
t −δ Ω
w (x, t ) dx = |Ω|s (t 0).
Ω
Define a function
ρ (t ) := sup w (x, τ ): 0 τ t , x ∈ Ω . (2.4)
Lemma 2.3. There is a small number δ > 0, which is independent of the initial data, such that the function ρ
defined by (2.4) satisfies
0 ρ (t ) max Ms, ρ (δ) (t 0),
t −δ
∇ u (x, t ) ∇ G 1 (x, y , t )u 0 ( y ) dy + ∇ G 1 (t − τ , x, y ) f (u , v ) dy dτ
Ω 0 Ω
t
+ ∇ G 1 (t − τ , x, y ) f (u , v ) dy dτ
t −δ Ω
t −δ
c1 c 1 e −η1 (t −τ )
u 0 ( y ) dy + a11 u ( y , τ ) + a12 v ( y , τ ) dy dτ
δ (n+1)/2 δ(n+1)/2
Ω 0 Ω
1666 S. Jimbo, Y. Morita / J. Differential Equations 255 (2013) 1657–1683
t
c1 |x − y |2
+ exp − a11 u ( y , τ ) + a12 v ( y , τ ) dy dτ
(t − τ )(n+1)/2 η2 (t − τ )
t −δ Ω
t
c 1 u 0 L 1 c 1 s|Ω| c 1 dτ
+ max(a11 , a12 ) + (πη2 )n/2 √ · max(a11 , a12 )ρ (t )
δ (n+1)/2 δ (n+1)/2 η1 t−τ
t −δ
c1 √
u 0 L 1 + s|Ω| max(a11 , a12 )/η1 + 2c 1 (πη2 )n/2 δ max(a11 , a12 )ρ (t ).
δ (n+1)/2
Similarly, we obtain
c1
∇ v (x, t ) v 0 L 1 + s|Ω| max(a21 , a22 )/η1
δ (n+1)/2
√
+ 2c 1 (πη2 )n/2 δ max(a21 , a22 )ρ (t ).
√
2K (Ω)c 1 (πη2 )n/2 δ max(ai j : 1 i , j 2) 1/4 (2.5)
is satisfied. Then
K (Ω)c 1 s 1
0 w (x, t ) s + η1 + 2|Ω| max(ai j : 1 i , j 2) + sup w (x, τ )
η1 δ (n+1)/2 2 x∈Ω,0τ t
for t δ . By considering the non-blowup of the solutions in a finite time, a simple argument leads us
to
2K (Ω)c 1 s
0 ρ (t ) max 2s + η1 + 2|Ω| max(ai j : 1 i , j 2) , ρ (δ)
η1 δ (n+1)/2
for t 0. It is clear by (2.5) that δ is chosen independently of the initial data. Consequently, we obtain
the desired assertion of the lemma. 2
By virtue of Lemma 2.3 we have the global boundedness of the orbit defined by the solutions
(u (t , x), v (t , x)) in L ∞ (Ω)2 . Applying this result to the estimating of |∇ u (x, t )| and |∇ v (x, t )| with the
aid of the integral equations, we easily get the global C 1 estimates for (u (t , x), v (t , x)), in consequence,
we conclude the proof of Proposition 1.1.
S. Jimbo, Y. Morita / J. Differential Equations 255 (2013) 1657–1683 1667
We prove Proposition 1.2. Let (u (·, t ), v (·, t )) be a solution of (1.1) with (1.2). A straightforward
computation shows the following:
d
E u (·, t ), v (·, t )
dt
= d∇(u + v ) · ∇(u + v )t + ∇(du + v ) · ∇(du + v )t
Ω
+ (1 − d) g (u + v ) + dk(u + v ) (u + v )t dx
= −d(u + v ) + (1 − d) g (u + v ) + dk(u + v ) (u + v )t
Ω
+ ∇(du + v ) · ∇(du + dv )t + (1 − d)∇(du + v ) · ∇ v t dx
=− d(u + v ) − (1 − d) g (u + v ) − dk(u + v ) (u + v )t
Ω
+ d(du + v )(u + v )t + (1 − d)(du + v ) v t dx.
Then we use
ut + dv t = d(u + v ) − (1 − d) g (u + v ) + k(1 − d) v ,
and
ut + v t = (du + v ),
to obtain
d
E u (·, t ), v (·, t ) = − ut + dv t − k(1 − d) v − dk(u + v ) (u + v )t
dt
Ω
+ d(ut + v t )2 + (1 − d)(ut + v t ) v t dx
=− ut + dv t − k(du + v ) + (1 − d) v t (u + v )t + d(ut + v t )2 dx
Ω
=− (1 + d)(ut + v t )2 − k(du + v )(du + v ) dx
Ω
2
=− (1 + d)(ut + v t )2 + k∇(du + v ) dx 0.
Ω
Thus
d
E u (·, t ), v (·, t ) = 0 (∀t ∈ R)
dt
1668 S. Jimbo, Y. Morita / J. Differential Equations 255 (2013) 1657–1683
implies
(u + v )t = w t = 0 and du + v = dw + (1 − d) v = constant in x.
This yields W t = V t = 0 in (1.9). Then the solution ξ(t ) of (1.11) defined for all t must be a steady
state. This concludes the proof of Proposition 1.2. 2
that is,
−1
1 −1 1 1 d 0 −g (w∗) −g (w∗) + k
A =− − ,
0 1 0 1 0 g (w∗) g (w∗) − k
the assertion of the lemma immediately follows from [15] (Theorem 6.29, Chap. 3, §§6–8). 2
dφ − (1 − d) g w ∗ φ + (1 − d)kψ = −λφ + (1 − d)λψ. (3.4)
By putting
1−d d
χ =φ+ ψ i.e. ψ = (χ − φ) ,
d 1−d
λ = λ 1 + i λ2 , χ = χ1 + i χ2 , φ = φ1 + i φ 2 ,
S. Jimbo, Y. Morita / J. Differential Equations 255 (2013) 1657–1683 1669
and
J 1 := (χ1 φ1 + χ2 φ2 ) dx, J 2 := (χ1 φ2 − χ2 φ1 ) dx,
Ω Ω
Taking the real part and imaginary part of the equalities yields
d ∇ χ1 2 + ∇ χ2 2 = λ1 J 1 − λ2 J 2 ,
λ2 J 1 + λ1 J 2 = 0,
d λ2 J 1 + (k − λ1 ) J 2 = λ2 (1 + d) φ1 2 + φ2 2 .
λ2 d ∇ χ1 2 + ∇ χ2 2 = − λ21 + λ22 J 2 ,
λ2 (1 + d) φ1 2 + φ2 2 = −d(2λ1 − k) J 2
d2 (2λ1 − k) ∇ χ1 2 + ∇ χ2 2 = (1 + d) λ21 + λ22 φ1 2 + φ2 2 0,
We define the closed operator A 0 the extension of − with the Neumann boundary condition
in L 2 (Ω). Namely, A 0 satisfies
A 0 v = − v , v ∈ u ∈ H 2 (Ω; R): ∂ u /∂ ν = 0 (x ∈ ∂Ω), u dx = 0
Ω
and A 0s stands for the fractional power of A 0 with index s. We note that A 0 is self-adjoint, positive
and invertible in L 2 (Ω). Thus A 0−1 is defined as a bounded operator in L 2 (Ω).
By using A 0−1 , we obtain the equation
d λ
ψ − ψ =− φ+ A− 1
0 φ, (3.7)
1−d 1−d
λ
φ − α g w ∗ φ − kφ + α (k − λ) ψ = − (1 + d) + (k − λ) A −
0
1
φ
d
1670 S. Jimbo, Y. Morita / J. Differential Equations 255 (2013) 1657–1683
which is decomposed to
g w ∗ φ − (k − λ) ψ = 0, (3.8)
and
λ
L(φ) = (1 + d) + (k − λ) A −
0
1
φ, (3.9)
d
where L is defined by (1.18). Given a solution (φ, λ) of (3.9), ψ is determined by (3.8). In the sequel
the eigenvalue problem of A is reduced to that of (3.9).
Lemma 3.3. The algebraic multiplicity and the geometric multiplicity of any eigenvalue λ of A coincide if
Re λ < k/2.
Proof. By Lemma 3.2 λ is real if Re λ < k/2. We let λ0 be any eigenvalue with λ0 < k/2 and (φ0 , ψ0 )
be a corresponding eigenfunction. It suffices to prove
that is,
1 d 1 −1
ψ = − (dφ + λ0 φ + φ0 ), ψ − ψ =− φ+ λ0 A − 1
0 φ + A 0 φ0 .
1−d 1−d 1−d
Thus,
λ0
−φ + α g w ∗ φ + kφ − α (k − λ0 ) ψ − (1 + d)φ + (k − λ0 ) A − 1
0 φ
d
1 k − λ0 − 1
= −α ψ0 + φ0 + A 0 φ0 .
d d
g w ∗ φ − (k − λ0 ) ψ = − ψ0 ,
λ0 1 k − λ0 −1
L(φ) − (1 + d) + (k − λ0 ) A −
0
1
φ = −α ψ0 − ψ0 + φ0 + A 0 φ0 . (3.12)
d d d
By
d λ0 − 1
ψ 0 − ψ0 = − φ0 + A φ0 ,
1−d 1−d 0
(3.12) turns to be
λ0 1+d k − 2 λ0 − 1
L(φ) − (1 + d) + (k − λ0 ) A −
0
1
φ= φ0 + A 0 φ0 . (3.13)
d d d
Invoking
λ0
L(φ0 ) − (1 + d)φ0 + (k − λ0 ) A − 1
0 φ 0 = 0,
d
1
(1 + d)(φ0 , φ0 ) L 2 + (k − 2λ0 ) A −
0 φ0 , φ0 L 2
−1/2 2
= (1 + d)φ0 2 + (k − 2λ0 ) A 0 φ0 = 0.
g w ∗ φ − k ψ = −λ ψ ,
ψ̂ + g w ∗ φ − g w ∗ φ − kψ̂ = −λψ̂, ψ̂ = 0.
Thus, it suffices to consider the equation for ψ̂ since ψ follows from the solution (ψ̂, λ).
Instead of A we may consider the eigenvalue problem of the operator
φ dφ + (1 − d)ψ
A1 := − (3.14)
ψ ψ + g ( w ∗ )φ − g ( w ∗ )φ − kψ
D(A1 ) = (φ, ψ) ∈ H 2 (Ω) × H 2 (Ω): ∂φ/∂ ν = ∂ψ/∂ ν = 0 (x ∈ ∂Ω) .
Moreover, from the same computation found in the proof of the above lemma we see that the eigen-
value problem for A1 is reduced to the problem
1672 S. Jimbo, Y. Morita / J. Differential Equations 255 (2013) 1657–1683
λ
L(φ) = (1 + d) + (k − λ) A −
0
1
φ (3.15)
d
∂φ
L[φ] = Λ dc + β A −
0
1
φ (x ∈ Ω), = 0 (x ∈ ∂Ω), (3.16)
∂ν
where β is nonnegative parameter. This is an eigenvalue problem of L with the weight in L 2 (Ω).
Then we have the following comparison result for the eigenvalues of L and Λ’s of (3.16), which will
be proved in the next section.
Lemma 3.4. Let {μ j } j =1,2,... and {Λ j (β)} j =1,2,... be the sets of eigenvalues of L and (3.16) for β 0 arranged
in increasing order with counting the multiplicity respectively. If μ j > 0, then
μj μj
Λ j (β) , (3.17)
dc + β/σ2 dc
μj μj
Λ j (β) (3.18)
dc dc + β/σ2
holds, where σ2 denotes the second eigenvalue of − with the Neumann boundary condition in the domain Ω .
Compare this and (3.16) with dc = (1 + d)/d. If there is β ∗ solving the equation
k − Λ j (β)
= β, (3.19)
d
then Λ j (β ∗ ) and the corresponding eigenfunction φ j (·; β ∗ ) give solutions to (3.15). Hence, Λ j (β ∗ ) is
an eigenvalue of A1 , that is, A. The next lemma is crucial to realize this idea.
Lemma 3.5. Each Λ j (β) is continuous in β 0. Moreover, if μ j > 0, then Λ j (β) is strictly monotone de-
creasing while if μ j < 0, then Λ j (β) is strictly monotone increasing.
We will prove this key lemma in the next section. We apply Lemmas 3.4 and 3.5 to complete the
proof of Theorem 1.3.
Proof of Theorem 1.3. We first see that by Lemma 3.3 the multiplicity of zero eigenvalue for A and
L coincides. We put dc = (1 + d)/d and consider the eigenvalue problem (3.16). Assume that
Since Λ j (0) = μ j /dc , Lemmas 3.4 and 3.5 tells that, for each j p, Λ j (β) < 0 (β > 0) and k − Λ j (β)
is monotone decreasing so that a unique solution β = β ∗j of (3.19) exists. Moreover we see that
S. Jimbo, Y. Morita / J. Differential Equations 255 (2013) 1657–1683 1673
Fig. 1. The curves γ = γ j (β) := (k − Λ j (β))/d and γi (β) := (k − Λ j (β))/d for the case μ j < μi < 0 in (a) and 0 < μ j < μi
in (b). The horizontal axis is β -axis. The intersection points at the line γ = β give β ∗j and βi∗ respectively. Note Λn (0) = μn /dc .
Λ1 β1∗ Λ2 β2∗ · · · Λ p β p∗ < 0.
1+d k − λm −1
L[φm ] = λm + A 0 φm . (3.20)
d d
Put βm := (k − λm )/d. Then we easily see that λm ∈ {Λ j (βm )} j =1,2,... . The number of the negative
eigenvalues of L coincides with that of A.
If 0 < μ j < k/2, applying Lemma 3.5 similarly, we see that Λ j (β) > 0 (β > 0) and k − Λ j (β) is
monotone increasing so that there is a unique β ∗j > 0 solving (3.19). Since k − Λ j (β ∗j ) > 0, Λ j (β ∗j )
gives a positive eigenvalue of A. We also easily see that if A has q positive eigenvalues less than k/2,
then L has also so.
Finally, (1.21) immediately follows from Lemma 3.4. This completes the proof. 2
−1/2 −1/2
(u , v )β := dc (u , v ) L 2 + β A 0 u , A 0 v L 2 u , v ∈ L 2 (Ω; R) ,
−1/2 2 1/2
u β := dc u 2 + β A 0 u
σ2 A − 1/2 2
0 u u 2
holds, where σ2 be the second eigenvalue of − with the Neumann boundary condition. Thus, the
norm · β gives an equivalent norm to the usual one of L 2 .
1674 S. Jimbo, Y. Morita / J. Differential Equations 255 (2013) 1657–1683
We introduce the min–max principle for the eigenvalue problem (see [7]). Define the Rayleigh quo-
tient as
K [φ]
R [φ; β] := , K [φ] := ∇φ2 + a(·)φ, φ L2
.
φβ 2
Then we have
Λn (β) = inf sup R [φ; β]: φ ∈ X n , φ = 0 . (4.2)
Xn ∈Mn
Λn (β) = sup inf R [φ; β]: φ ∈ X n⊥−1 , φ = 0 (4.3)
X n−1 ∈Mn−1
= inf R [φ; β]: φ ∈ X , φ = 0, φ, φ j (·; β) β = 0 ( j = 1, 2, . . . , n − 1) , (4.4)
where
Y ⊥ := u: (u , v )β = 0 (∀ v ∈ Y ) .
Lemma 4.1. The zero eigenspace for (4.1) does not depend on β 0.
Proof. The zero eigenspace is the set of solutions of La [φ] = 0 which is independent of β 0. 2
φ2β2 φ2β1
φβ1 φβ2 , (φ ∈ X ),
β2 β1
and
β2
φ20 φ2β2 1 + φ20 (φ ∈ X )
σ2 d c
hold.
S. Jimbo, Y. Morita / J. Differential Equations 255 (2013) 1657–1683 1675
Proof. The first and the third inequalities immediately follows form the definitions of · β and σ2 .
For the second inequality, given φ ∈ X , we obtain
Proof. From the inf–sup type max–min principle (4.2) and the condition Λn (β2 ) > 0, there exists
δ > 0 such that
sup K [φ]/φ2β2 : φ ∈ E , φ = 0 δ E ∈ Mn .
sup K [φ]/φ2β2 : φ ∈ E , φ = 0 = sup K [φ]/φ2β2 : φ ∈ E , φ = 0, K [φ] 0 .
K [φ]
Λn (β2 ) = inf n sup : φ ∈ E, φ = 0
E ∈M φ2β2
K [φ]
= inf n sup : φ ∈ E , φ = 0, K [φ] 0
E ∈M φ2β2
K [φ]
inf n sup : φ ∈ E , φ = 0, K [φ] 0
E ∈M φ2β1
K [φ]
= inf n sup : φ ∈ E , φ = 0 = Λn (β1 ).
E ∈M φ2β1
Lemma 4.4. If 0 < β1 β2 and Λn (β1 ) > 0, then 0 < Λn (β1 ) (β2 /β1 )Λn (β2 ).
Proof. By (4.2) and Λn (β1 ) > 0 there exists δ > 0 such that
sup K [φ]/φ2β1 : φ ∈ E , φ = 0 δ E ∈ Mn .
Then we see
sup K [φ]/φ2β1 : φ ∈ E , φ = 0 = sup K [φ]/φ2β1 : φ ∈ E , φ = 0, K [φ] 0 .
K [φ]
Λn (β1 ) = inf n sup : φ ∈ E, φ = 0
E ∈M φ2β1
K [φ]
= inf n sup : φ ∈ E , φ = 0, K [φ] 0
E ∈M φ2β1
β2 K [φ]
inf n sup : φ ∈ E , φ = 0, K [φ] 0
E ∈M β1 φ2β
2
β2 K [φ]
inf sup : φ ∈ E , φ = 0 , K [φ] 0
β1 E ∈Mn φ2β2
β2 K [φ] β2
= inf sup : φ ∈ E , φ = 0 = Λn (β2 ).
β1 E ∈M n
φβ22 β1
Combining Lemmas 4.1, 4.3 and 4.4, we have the following result.
Lemma 4.5. If Λn (β0 ) > 0 for a β0 > 0, then Λn (β) > 0 for all β > 0. In addition, Λn (β0 ) < 0 for a β0 > 0
yields Λn (β) < 0 for all β > 0.
Proof. The first assertion follows directly from Lemmas 4.3 and 4.4. To see the second assertion, sup-
pose it is not true. Then there is β1 > 0 such that Λn (β1 ) 0. By Lemma 4.1 Λn (β) = 0 if Λn (β1 ) = 0,
while Λ(β) > 0 if Λn (β1 ) > 0, so we get a contradiction. 2
Lemma 4.6. Assume that there exists a number β0 > 0 for which Λn (β0 ) < 0 is met. Then
holds if 0 < β1 β2 .
Proof. We notice that the assumption implies Λn (β) < 0 for any β > 0 by Lemma 4.5. Since
Λn (β2 ) < 0, we see that arbitrarily given ∈ (0, −Λn (β2 )), there exists E ∈ Mn such that
sup K [φ]/φ2β2 : φ ∈ E , φ = 0 < Λn (β2 ) + .
Thus K [φ] < 0 for φ ∈ E . With the aid of this and φ2β1 φ2β2 , we have
Λn (β1 ) = inf n sup K [φ]/φ2β1 : φ ∈ E , φ = 0
E ∈M
sup K [φ]/φ2β1 : φ ∈ E , φ = 0
sup K [φ]/φ2β2 : φ ∈ E , φ = 0 < Λn (β2 ) + .
We prove the latter inequality. By Λn (β1 ) < 0, for any ∈ (0, −Λn (β1 )), there exists E ∈ Mn such
that
sup K [φ]/φ2β1 : φ ∈ E , φ = 0 < Λn (β1 ) + .
Using K [φ] < 0 for any φ ∈ E and 1/φ2β2 (β1 /β2 )(1/φ2β1 ), we have
Λn (β2 ) = inf n sup K [φ]/φ2β2 : φ ∈ E , φ = 0
E ∈M
sup K [φ]/φ2β2 : φ ∈ E , φ = 0
β1 K [φ] β1
sup : φ ∈ E , φ = 0 < Λn (β1 ) + .
β2 φβ2 β2
1
Combining the above results leads to the following continuity property for the eigenvalues in β .
Proposition 4.7. Every eigenvalue Λn (β) of (4.1) is continuous in β > 0. More precisely, for each β0 > 0
(i) if Λn (β0 ) < 0, then Λn (β) < 0 for any β > 0 and
β0 β0
Λn (β0 ) Λn (β0 + t ) Λn (β0 ) t ∈ (−β0 , β0 ) ;
β0 − |t | β0 + |t |
(ii) if Λn (β0 ) > 0, then Λn (β) > 0 for any β > 0 and
β0 β0
Λn (β0 ) Λn (β0 + t ) Λn (β0 ) t ∈ (−β0 , β0 ) ;
β0 + |t | β0 − |t |
Proof. The assertion of (iii) follows from Lemma 4.1. We prove (i) by Lemma 4.6. Since Λn (β) < 0,
we have
β0 β0
Λn (β0 ) Λn (β0 ) Λn (β0 + t ) Λn (β0 ),
β0 − t β0 + t
β0 + t β0
Λn (β0 ) Λn (β0 + t ), Λn (β0 + t ) Λn (β0 ) Λn (β0 ).
β0 β0 − t
Λn (0)
Λn (0) Λn (β) (for β > 0);
1 + (β/σ2 dc )
Λn (0)
Λn (β) Λn (0) (for β > 0);
1 + (β/σ2 dc )
The proof of this proposition is carried out similarly as done for Proposition 4.7. Indeed, using the
third inequality of Lemma 4.2, we see that the assertion of Lemma 4.3 is true for the case β1 = 0.
Similarly, with the aid of the same inequality, by replacing β2 /β1 by 1 + (β2 /σ2 dc ), the assertions
of Lemmas 4.4 and 4.6 are also true for β1 = 0. Hence it is not so difficult to complete the proof of
Proposition 4.8. We leave the details of the proof to the readers.
Proof of Lemmas 3.4 and 3.5. Put dc = (1 + d)/d. By Proposition 4.7 together with Lemmas 4.4, 4.5
and 4.6, we obtain Lemma 3.5. Invoking Λn (0) = μn /dc = dμn /(1 + d) and applying Proposition 4.8
yield Lemma 3.4. 2
Remark 4.1. We are able to prove that if Λn (β0 ) is simple for some β0 , then there exists τ > 0 such
that Λn (β) is differentiable in β in (β0 − τ , β0 + τ ) and
We omit the proof, since it is not used in the proof of the main result of the present article.
In the papers [19] and [18] the authors treat the case g (u ), instead of g (u + v ), namely
ut = du − g (u ) + kv ,
(x ∈ Ω, t > 0). (5.1)
v t = v + g (u ) − kv
u
G (u ) := g (ξ ) dξ.
It is clear that the functional (5.2) differs from (1.5). However, motivated by the fact that both systems
possess the Lyapunov functions, we get an idea to consider the following modified system:
S. Jimbo, Y. Morita / J. Differential Equations 255 (2013) 1657–1683 1679
ut = du − g (u + γ v ) + kv ,
(x ∈ Ω, t > 0) (5.3)
v t = v + g (u + γ v ) − kv
∂u ∂v
= = 0 (x ∈ ∂Ω), (5.4)
∂ν ∂ν
where d, k are positive constants, γ is a constant in [0, 1]. This system coincides with (1.1) and (5.1)
when γ = 1 and γ = 0 respectively.
We show that the system possesses a Lyapunov function.
Proposition 5.1. Let 0 < d < 1 and 0 γ 1. Then the following functional gives a Lyapunov function
of (5.3) with (5.4):
d
Eγ (u , v ) := ∇(u + γ v )2 + (1 − dγ )G (u + γ v ) + dk (u + γ v )2
2 2
Ω
k(1 − γ ) γ 2
+ (du + v )2 + ∇(du + v ) dx, (5.5)
2(1 − d) 2
1 − d2 γ
d
Eγ u (·, t ), v (·, t ) = − (u + γ v )t 2 dx − k(1 − dγ ) ∇(du + v )2 dx
dt 1 − dγ 1−d
Ω Ω
γ (1 − γ )
− (du + v )t 2 dx 0
1 − dγ
Ω
holds.
We note (5.5) coincides with (5.2) and (1.5) for γ = 0 and γ = 1 respectively.
Proof. We may assume γ < 1 since the case γ = 1 is proved in Section 2. We introduce the new
variables
U = u + γ v, V = du + v .
By
1 1
u= (U − γ V ), v= (−dU + V ),
1 − dγ 1 − dγ
1 − dγ γ (1 − d)
U t = dU + V − (1 − dγ ) g (U ) − dkU + kV , (5.6)
1−γ 1−γ
1−d 1−γ
Ut + V t = V , (5.7)
1 − dγ 1 − dγ
1680 S. Jimbo, Y. Morita / J. Differential Equations 255 (2013) 1657–1683
d 2 dk 2 k(1 − γ ) 2 γ 2
Eγ ( U , V ) = |∇ U | + (1 − dγ )G (U ) + U + V + |∇ V | dx.
2 2 2(1 − d) 2
Ω
For the solution (U (x, t ), V (x, t )) of (5.6)–(5.7) with the Neumann conditions
d
Eγ U (·, t ), V (·, t )
dt
k(1 − γ )
= d∇ U · ∇ U t + (1 − dγ ) g (U )U t + kdU U t + V V t + γ ∇ V · ∇ V t dx
1−d
Ω
1 − dγ
= −dU + (1 − dγ ) g (U ) + kdU U t + kV V − U t − γ V t V dx
1−d
Ω
1 − dγ γ (1 − d)
= − Ut + V + kV U t
1−γ 1−γ
Ω
k(1 − dγ ) 1−d 1−γ
− |∇ V |2 − kV U t − γ V t Ut + Vt dx
1−d 1 − dγ 1 − dγ
1 − dγ γ (1 − d) 1 − d 1−γ
= − U t2 + Ut + Vt Ut
1−γ 1−γ 1 − dγ 1 − dγ
Ω
γ (1 − d) γ (1 − γ ) 2 k(1 − dγ )
− Vt Ut − Vt − |∇ V |2 dx
1 − dγ 1 − dγ 1−d
1 − dγ γ (1 − d)2 2 γ (1 − γ ) 2 k(1 − dγ )
=− 1− Ut + Vt + |∇ V |2 dx.
1−γ (1 − dγ )2 1 − dγ 1−d
Ω
Since
(1 − d)2 (1 − γ )(1 − d2 γ )
1−γ = ,
(1 − dγ )2 (1 − dγ )2
Acknowledgments
The first author was supported in part by the Grant-in-Aid for Scientific Research (C) No. 22540216,
Japan Society for the Promotion of Science. The second author was partially supported by the Grant-
in-Aid for Scientific Research (B) No. 22340022 and Challenging Exploratory Research No. 24654044,
Japan Society for the Promotion of Science. The authors express their thanks to the referee for valu-
able comments to the improvement of the manuscript.
S. Jimbo, Y. Morita / J. Differential Equations 255 (2013) 1657–1683 1681
w
g ( w ) = kg 0 ( w ), g 0 ( w ) := , (A.1)
(aw + 1)2
which is given in [14] and [20], we verify the condition for the instability of a constant steady state
of (1.1). Under the condition (1.4), we obtain a constant steady state
(u , v ) = s − g 0 (s), g 0 (s) . (A.2)
du dv
= −kg 0 (u + v ) + kv , = kg 0 (u + v ) − kv (A.3)
dt dt
under the constraint u + v = s because the equations are reduced to a scalar equation
du
= k − g 0 (s) + s − u .
dt
− ψ + kg 0 (s)(φ + ψ) − kψ = λψ (x ∈ Ω),
dσ j + kg 0 (s) kg 0 (s) − k
A j := ( j = 1, 2, . . .).
−kg 0 (s) σ j − kg0 (s) + k
Hence
λ2 − (d + 1)σ j + k λ + dkσ j σ j /k + 1 − (1 − 1/d) g 0 (s) = 0.
Considering
1 − aw
g0 ( w ) = ,
(aw + 1)3
we know that for s > 1/a there are unstable modes j’s if d is sufficiently small.
1682 S. Jimbo, Y. Morita / J. Differential Equations 255 (2013) 1657–1683
Lemma 2.2 provides a point wise estimate of the gradient of the heat kernel (fundamental solution
of the Laplacian with the Neumann boundary condition) for all t > 0. It seems that this kind of the
estimate is known to the experts who are familiar to the Green function. Unfortunately, there is no
suitable literature for this result. We therefore show how this lemma can be proved in this section.
Since G 1 (x, y , t ) = G 2 (x, y , dt ), we deal with only G 2 , and put G (x, y , t ) = G 2 (t , x, y ) for simplicity
of notation.
Proof of Lemma 2.2. Short time estimate: We consider the interval 0 < t 1. From the process of the
construction of the fundamental solution for the initial boundary value problem of the heat equation
as in [16], we see that there exist constants c 1 > 0, c 2 > 0 such that
2
∇x G (x, y , t ) c 1 exp − |x − y | (0 < t 1, x, y ∈ Ω).
n +1
t 1/2 c2t
Long time estimate: We consider the interval [1, ∞). The following expressions are well known:
∞
G (x, y , t ) = exp(−σk t )Φk (x)Φk ( y ),
k =1
∞
∇x G (x, y , t ) = exp(−σk t ) ∇x Φk (x) Φk ( y ),
k =1
where σk and Φk are the k-th eigenvalue and the corresponding eigenfunction of with the Neu-
mann boundary condition. Here {Φk } is orthonormalized in L 2 (Ω). Note that σ1 = 0 and Φ1 (x) =
1/|Ω|1/2 (constant function), therefore the first term of ∇x G (x, y , t ) vanishes. The convergence of the
above infinite series is assured by the following estimates:
σk c3k2/n − c4 , (B.1)
n/2 (n/2)+1
Φk L ∞ (Ω) c 5 k ,
σ ∇Φk L ∞ (Ω) c 6 σk (k 1), (B.2)
where c 3 , c 4 , c 5 and c 6 are positive constants depending only on Ω . Those inequalities in (B.1)–(B.2)
are verified as follows: The inequality (B.1) follows from the famous Weyl formula for the asymptotic
distribution of the eigenvalues (see [6])
|Ω|
N (s) := max{k ∈ N | σk < s} = n/2 Γ ((n/2) + 1)
sn/2 + o sn/2 (s → ∞),
2n π
while the first inequality of (B.2) is due to the result in [17]. For the proof of the second inequality
in (B.2) we use the Sobolev inequality (see [1]) and the elliptic estimate (see [2]) which applies to
Φk = −σk Φk in Ω , ∂Φk /∂ ν = 0 on ∂Ω , in the manner
∇Φk L ∞ (Ω) c 7 ∇Φk W 1, p (Ω) c 8 Φk W 2, p (Ω) c 9 σk Φk L p (Ω) + Φk L p (Ω)
(n/2)+1
c 9 (σk + 1)|Ω|1/ p Φk L ∞ (Ω) c 9 c 5 |Ω|1/ p σk ,
∞
∇x G (x, y , t ) exp(−σk t )∇x Φk (x)Φk ( y )
k =2
∞
exp(−σ2 t ) exp −(σk − σ2 )t ∇x Φk L ∞ (Ω) Φk L ∞ (Ω)
k =2
∞
exp(−σ2 t ) exp −(σk − σ2 ) c 5 c 6 σkn+1 (t 1, x, y ∈ Ω).
k =2
n+1 ∞
Due to the estimates (B.1) the infinite series k=2 σk exp(−(σk − σ2 )) converges. Since Ω is
bounded, the above estimate for ∇x G (x, y , t ) is bounded by the right hand side of the inequality
of Lemma 2.2 in t 1 if c 1 , η1 and η2 in the lemma are appropriately chosen.
Combining the short time estimate and the long time estimate, we obtain the desired inequality
of Lemma 2.2. 2
References