1. Introduction
This paper considers the following reaction-diffusion equation with Neumann boundary condition under climate change in an infinitely long straight cylinder:
\begin{equation} \begin{cases} \partial _{t}u=\Delta u+f\left (t,x_{1}-\int ^{t}_{0}c(s)ds,y,u\right ),\, t\gt 0,x_{1}\in \mathbb{R}, y\in \omega ,\\[4pt] \partial _{\nu }u(t,x_{1},y)=0,\quad t\gt 0,x_{1}\in \mathbb{R},y\in \partial \omega . \end{cases} \end{equation}
We point that in (1.1)
$\omega \subset \mathbb{R}^{N-1}$
is a bounded smooth domain. Here,
$c\,:\,\mathbb{R}\rightarrow \mathbb{R}$
is H
$\ddot {o}$
lder continuous and almost periodic in the sense that for any
$\epsilon \gt 0$
, the set
is relatively dense in
$\mathbb{R}$
. Set
Then,
$\nu$
in (1.1) denotes the exterior unit normal vector field to
$\Omega$
and
$\partial _{\nu }\,:\!=\,\nu \cdot \nabla$
. In (1.1),
$c(t)$
is the shifting speed of the climate envelope and
$\int ^{t}_{0}c(s)ds$
is the location of the climate envelope. Let
$f(t,x,u)=ug(t,x,u)$
, where the function
$g\,:\,\mathbb{R}\times \Omega \times [0,+\infty )\rightarrow \mathbb{R}$
is continuously differentiable, uniformly continuous, and bounded on
$\mathbb{R}\times \Omega \times [0,\delta ]$
for any
$\delta \gt 0$
. Then, we have the following standard assumptions for
$g(t,x,u)$
:
-
(H1)
$g(t,x,u)$
is almost periodic in
$t$
uniformly with respect to
$x\in \Omega$
and
$u$
in bounded sets of
$[0,+\infty )$
; -
(H2)
$g_{u}(t,x,u)\leq 0$
for all
$(t,x,u)\in \mathbb{R}\times \Omega \times [0,+\infty )$
, and there is a domain
$ D\subset \Omega$
with
$|D|\gt 0$
such that
$u\mapsto g(t,x,u)$
is decreasing on
$\mathbb{R}^{+}$
for each
$x\in D$
and
$t\in \mathbb{R}$
; -
(H3) there is
$M_{0}\gt 0$
such that
$\sup _{t\in \mathbb{R},x\in \Omega }g(t,x,u)\lt 0$
for all
$u\geq M_{0}$
; -
(H4)
$\lim _{r\rightarrow \infty }\sup _{|x_{1}|\gt r,t\in \mathbb{R},y\in \omega }g(t,x_{1},y,0)\lt 0$
.
Here, (H1)–(H2) say that
$f(t,x,u)$
is of logistic type and
$M_{0}$
in (H3) is an upper bound for the carrying capacity. Assumption (H4) means that the favourable zone where species can grow is bounded. A typical example of
$g(t,x,u)$
satisfying (H1)–(H4) is
$g(t,x,u)=a(t,x)-b(t,x)u$
with
$a(t,x), b(t,x)\in L^{\infty }(\mathbb{R}\times \Omega )$
such that
$b(t,x)\geq 0$
a.e. in
$\mathbb{R}\times \Omega$
,
$b(t,x)\gt 0$
in
$\mathbb{R}\times D$
with
$|D|\gt 0$
and
$D\subset \Omega$
,
where
$a(t,x)$
and
$b(t,x)$
are almost periodic in
$t$
uniformly with respect to
$x\in \Omega$
. Recall that a function
$f(t,x)\in C(\mathbb{R}\times E,\mathbb{R})$
is said to be almost periodic in
$t$
uniformly with respect to
$x\in E\subset \mathbb{R}^{N}$
, if it is uniformly continuous on
$\mathbb{R}\times E$
and for any fixed
$x\in E$
,
$f(t,x)$
is an almost periodic function of
$t$
.
In the past years, to describe the impacts of global warming on biological species, a vast amount of research has been carried out regarding the population dynamics of reaction-diffusion equation system under climate change. See, for example, [Reference Berestycki, Diekmann, Nagelkerke and Zegeling1, Reference Berestycki, Desvillettes and Diekmann4, Reference Berestycki and Fang5, Reference Du, Wei and Zhou7, Reference Fang, Lou and Wu8, Reference Hu and Li11, Reference Hu, Shang and Li13, Reference Lewis, Marculis and Shen15, Reference Li, Wang and Zhao20–Reference Potapov and Lewis22, Reference Wang, Li, Dong and Qiao28] and references therein. We also refer to see [Reference Wang, Li, Dong and Qiao28] for more results and understanding on the diffusion and propagation in shifting environments. It should be pointed out that in Berestycki and Rossi [Reference Berestycki and Rossi2], the authors proved the existence and uniqueness of travelling wave solutions of the type
$u(t,x)=U(x-cte)$
for the reaction-diffusion equation
and the large time behaviour of solutions with arbitrary nonnegative bounded initial datum depend on the sign of the generalised principle eigenvalue in
$\mathbb{R}^{N}$
of an associated linear operator. These problems with
$N=1$
have been studied in [Reference Berestycki, Desvillettes and Diekmann4] to investigate the impact of climate shift on the dynamics of a biological species. Berestycki and Rossi [Reference Berestycki and Rossi3] further established analogous results for the Neumann problem in domains which are asymptotically cylindrical, as well as for the problem in the whole space with
$f$
periodic in some space variables, orthogonal to the direction of the life. When the environment is assumed to be only globally unfavourable, with favourable pockets extending to infinity, Vo [Reference Vo27] studied the persistence versus extinction of species in reaction-diffusion equation with an infinite cylindrical domain. Boubours and Giletti [Reference Bouhours and Giletti6] also studied the spreading and vanishing situations for a monostable reaction-diffusion equation when the initial datum is compactly supported. Recently, in a time-periodic shifting environment, Fang et al. [Reference Fang, Peng and Zhao9] studied a nonautonomous reaction-diffusion equation
$u_{t}=u_{xx}+ug(t,x-ct,u)$
and showed that there exists
$c^{*} \gt 0$
such that a unique forced time-periodic wave exists if and only if
$|c|\lt c^{*}$
and it attracts other solutions in a certain sense according to the tail behaviour of initial values. Later, Zhang and Zhao [Reference Zhang and Zhao30] extended the above results to the nonlocal dispersal equation in time-periodic shifting environment.
From above works, we find that the shift environment is considered by
$x-ct$
with a constant speed. In real environments, climate change is influenced by various seasonal variations and is rather fluctuating and unpredictable, leading to fluctuations in the size, shifting speed, and location of the climate envelope. Very recently, Shen et al. [Reference Shen, Shen, Xue and Zhou26] investigated the population dynamics of a reaction-diffusion equation under climate change in a spatiotemporally heterogeneous environment, which is described by a climate envelope shifting with a time-dependent speed function. They established the persistence criterion in term of the sign of the approximate top Lyapunov exponent and the existence of a unique forced wave solution that dominates the population profile of species in the long run.
Motivated by [Reference Berestycki and Rossi3] and [Reference Shen, Shen, Xue and Zhou26], in the current paper, we try to study the persistence criterion and the existence of forced waves solution of equation (1.1) with Neumann boundary conditions in time-dependent environment. Set
$\partial _{1}\,:\!=\,\frac {\partial }{\partial x_{1}}$
. To consider (1.1) in the moving frame, we introduce the change of variable
$v(t,x_{1},y)=u(t,x_{1}+\int ^{t}_{0}c(s)ds,y)$
, and
$v(t,x_{1},y)$
satisfies the following reaction-diffusion equation with Neumann boundary condition
\begin{equation} \begin{cases} \partial _{t}v=\Delta v+c(t)\partial _{1}v+f\left (t,x,v\right ),\, t\gt 0,x\in \Omega ,\\[4pt] \partial _{\nu }v(t,x)=0,\quad t\gt 0,x\in \partial \Omega . \end{cases} \end{equation}
Note that
$v=0$
is a solution of (1.2). We first consider the stability of
$v=0$
. Then the linearisation of (1.2) at
$v=0$
is
\begin{equation} \begin{cases} \partial _{t}w=\Delta w+c(t)\partial _{1}w+g(t,x,0)w,\quad t\gt 0,\,x\in \Omega ,\\[4pt] \partial _{\nu }w=0,\quad t\gt 0,\,x\in \partial \Omega . \end{cases} \end{equation}
Our main results depend on the sign of Neumann Lyapunov exponent of Neumann problem (1.3). The results on the Neumann principal Lyapunov exponent of (1.3) are given in Section 2. Set
$L\gt 0$
. Denote by
$\lambda _{L}$
the principal Lyapunov exponent of (1.3) restricted on
$({-}L,L)\times \omega$
and equiped with Neumann boundary condition
$\partial _{\nu }w=0$
on
$({-}L,L)\times \partial \omega$
and Dirichlet boundary condition on
$\pm L$
,
\begin{equation} \begin{cases} \partial _{t}w=\Delta w+c(t)\partial _{1}w+g(t,x,0)w,\quad t\gt 0,\,x\in ({-}L,L)\times \omega ,\\[4pt] \partial _{\nu }\omega =0,\quad t\gt 0,\,x\in ({-}L,L)\times \partial \omega ,\\[4pt] w(t,-L,y)=w(t,L,y)=0,\quad t\gt 0,\,y\in \omega . \end{cases} \end{equation}
We refer the reader to Section 2 for more details on the principal Lyapunov exponent
$\lambda _{L}$
. It is shown in Lemma 2.2 that
$\lambda _{L}$
is non-decreasing in
$L$
and bounded above by
$\sup _{t\in \mathbb{R},x\in [{-}L,L]\times \omega }g(t,x,0)$
, then
is well defined and called the approximate principal Lyapunov exponent of (1.3). In fact,
$\lambda _{\infty }$
always has the same sign as that of the principal Lyapunov exponent
$\lambda$
of (1.3), see Lemma 2.3 later.
Let
with the
$C^1$
norm
$\|u\|=\sup _{x\in \mathbb{R}\times \omega }(|u(x)|+|\nabla u(x)|)$
, where
$|\nabla u(x)|=\Sigma _{i=1}^N|u_{x_i}(x)|$
. Let
For any
$u_{0}(\cdot )\in X^{+}$
, let
$u(t,x;\,u_{0})$
be the unique solution of (1.1) with
$u(0,x;\,u_{0})=u_{0}(x)$
for all
$t\geq 0$
.
Theorem 1.1. Assume that (H1)-(H4) hold true. Then the following properties hold:
-
(1) If
$\lambda _{\infty }\lt 0$
, for any
$u_{0}\in X^{+}$
, there is
uniformly with respect to
\begin{equation*} \lim _{t\rightarrow \infty }u(t,\cdot ;\,u_{0})=0 \end{equation*}
$x\in \Omega$
.
-
(2) If
$\lambda _{\infty }\gt 0$
, then for any
$u_{0}\in X^{+}\setminus \{0\}$
and any
$L\gt 0$
, there holds
\begin{equation*} \liminf _{t\rightarrow \infty }\inf _{x_{1}\in [{-}L,L],y\in \omega }u\left (t,x_{1}+\int ^{t}_{0}c(s)ds,y;\,u_{0}\right )\gt 0. \end{equation*}
A forced wave solution
$U(t,x_{1},y)$
of (1.1) is a positive, bounded and almost-periodic entire solution
$V(t,x_{1},y)$
of (1.2) with the form
Theorem 1.2.
Assume that (H1)-(H4) hold true. If
$\lambda _{\infty }\gt 0$
, then (1.1) admits a unique forced wave solution
$U(t,x_{1},y)$
with
uniformly with respect to
$t\in \mathbb{R}$
and
$y\in \omega$
. Moreover, for any
$u_{0}\in X^{+}\setminus \{0\}$
, there holds
uniformly with respect to
$x_{1}\in \mathbb{R}$
and
$y\in \omega$
.
We point out that Theorem 1.1 follows from Theorems 3.1 and 3.2. Theorem 1.2 follows directly from Theorem 4.1. From Theorem 1.1, we know that the sign of approximate principal Lyapunov exponent
$\lambda _{\infty }$
is the sharp threshold for the persistence criterion. When
$\lambda _{\infty }\gt 0$
, the species will persist and (1.1) exists a unique, bounded and positive forced wave solution
$U(t,x_{1},y)$
in cylindrical type domain.
In order to analyse the effect of the shift speed on
$\lambda _{\infty }$
, we set
$c(t)=c+A\sigma (t)$
, where
$A$
is positive constant and
$A\sigma (t)$
is the fluctuation on the shifting speed
$c$
. Let
$\lambda ^{A}_{\infty }(c)$
be the approximate principal Lyapunov exponent depend on
$c$
and
$A$
. By using the Liouville transform, Theorem 4.2 in Section 4 shows that there is a critical speed
$c^{*}\gt 0$
such that
$\lambda ^{0}_{\infty }(c)\gt 0$
for
$c\in [0,c^{*})$
and
$\lambda ^{0}_{\infty }(c)\lt 0$
for
$c\gt c^{*}$
when
$A=0$
. This means that the species will persist when the climate envelope shifts with a slower shift speed. In this case, we also show that the force wave solution
$U(t,x_{1},y)$
is exponentially decay, see Theorem 4.3. When
$A\gt 0$
, we only show that
$\lambda ^{A}_{\infty }(c)$
has a lower bounded depend on
$c$
, see Theorem 4.2 (2). Due to the presence of cylinder domain and time almost periodic case, the characterisation of influence of the fluctuation on the shifting speed
$A\sigma (t)$
on
$\lambda ^{A}_{\infty }(c)$
tends to be much more difficult. We leave it as an open problem.
We also remark here that we only consider the propagation dynamics of time-dependent reaction-diffusion equations (1.1) under (H4). Specifically, (H4) implies that the favourable habitat for species growth is contained in a bounded region, i.e., the favourable environment is surrounded by an unfavourable one. When
the environment is favourable at
$+\infty$
and unfavourable at
$-\infty$
. Moreover, the case
implies the scenario that unfavourable environment is surrounded by favourable one. There is still few works on these two cases for time-dependent reaction-diffusion equations in an infinite cylinder and we will continue to study these two cases in the future. Very recently, what is worth mentioning is that Zhang and Zhao [Reference Zhang and Zhao31] studied spreading properties and forced travelling waves of reaction-diffusion equation in a time-heterogenous shifting environment in the case that favourable environment is surrounded by unfavourable oneenvironment is favourable at
$+\infty$
and unfavourable at
$-\infty$
.
The rest of the paper is organised as follows. In Section 2, we will study the Lyapunov exponent theory of Neumann problem (1.3) and mixed boundary conditions problem (1.4). In Section 3, we consider the long-time dynamical behaviours of (1.1) with the initial value. and prove Theorem 1.1. We will prove the existence, uniqueness, stability and exponential decay of forced wave solutions of (1.1) in Section 4.
2. Lyapunov exponent theory with mixed boundary conditions
In this section, we study the principal spectral theory for time almost periodic parabolic equation with mixed Dirichlet–Neumann boundary condition in the finite cylinders
$({-}L,L)\times \omega$
for any
$L\gt 0$
. Recall that
$\Omega =\mathbb{R}\times \omega$
. Denote
Let
$b_{i}(t,x)\, (i=1,\ldots ,N), b_{0}(t,x)$
be bounded and continuous in
$\mathbb{R}\times \Omega$
. Assume that
$b_{i}(t,x)\, (i=1,\ldots ,N)$
and
$b_{0}(t,x)$
are time almost-periodic in
$t$
uniformly for
$x\in \Omega$
. For each
$L\gt 0$
, we consider the following linear problem
\begin{equation} \begin{cases} \partial _{t}u=\Delta u+\sum ^{N}_{i=1}b_{i}(t,x)\frac {\partial u}{\partial x_{i}}+b_{0}(t,x)u,\quad t\gt 0,\,x\in ({-}L,L)\times \omega ,\\[4pt] \partial _{\nu }u=0,\quad t\gt 0,\,x\in ({-}L,L)\times \partial \omega ,\\[4pt] u(t,-L,y)=u(t,L,y)=0,\quad t\gt 0,\,y\in \omega . \end{cases} \end{equation}
The principal spectral theory of linear parabolic problem with Dirichlet, Neumann boundary conditions or periodic boundary condition has been widely studied, for example, see [Reference Huston, Shen and Vickers12, Reference Mierczyński and Shen16–Reference Mierczyński and Shen18, Reference Shen23–Reference Shen and Yi25, Reference Yi29] and so on. Note that the linear problem (2.1) is considered in the finite cylinders
$\Omega _{L}$
with Neumann boundary conditions on the sides
$({-}L,L)\times \partial w$
and Dirichlet boundary conditions on the bases
$\{\pm L\}\times \omega$
, which can be called as mixed Dirichlet–Neumann boundary conditions as that in [Reference Berestycki and Rossi3]. In this section, we will develop the principal spectral theory for time almost periodic parabolic equation in finite cylinder with mixed Dirichlet–Neumann boundary conditions.
Let
$b\,:\!=\,(b_{i},b_{0})\,:\!=\,(\{b_{i}\}^{N}_{i=1},b_{0})$
. Define
where
$b\cdot s(t,x)=\sigma _{s}b(t,x)=b(t+s,x)$
and the closure is taken under the topology of local uniform convergence. Then,
$(\mathrm{Y}(b),(\sigma _{t})_{t\in \mathbb{R}})$
is a compact flow and unique ergodic, which implies that
$\mathrm{Y}(b)$
is unique ergodic and minimal. For convenience, we write
$\mathrm{Y}$
for
$\mathrm{Y}(b)$
.
Let
be equipped with the
$C^1$
norm
$\|u\|_{X_{L}}=\max _{[{-}L,L]\times \omega }(|u(x)|+|\nabla u(x)|)$
, where
$|\nabla u(x)|=\Sigma _{i=1}^N|u_{x_i}(x)|$
. Let
Assume
$\tilde X_{L}\subset L^p(\Omega _L)$
(
$p\gt n$
) be a fractional power space of
$-\Delta \,:\,\mathcal{D}\to L^p(\Omega _L)$
satisfying
$\tilde X_{L}\hookrightarrow C^{\alpha }(\Omega _L)$
for some
$1\lt \alpha \lt 2$
, where
$\mathcal{D}=\{v\in H^{2,p}(\Omega _L)\,:\,u({-}L,y)=u(L,y)=0, \partial _{\nu } u(x_1,y)=0, y\in \partial \omega \}$
. Then, for any
$\widetilde {b}\in \mathrm{Y}$
and
$u_{0}\in \tilde X_{L}$
, there is a unique classical solution
$u(t,\cdot ;\,u_{0},\widetilde {b})$
of (2.1) with the initial condition
$u(0,\cdot ;\,u_{0},\widetilde {b})=u_{0}\in \tilde X_{L}$
. Put
For convenience of expression, in the following, we still write
$b$
for
$\widetilde {b}$
. Then, (2.1) generates a skew-product semiflow on
$\tilde X_{L}\times \mathrm{Y}(b)$
:
Let
Then, we have the following exponential separation theorem from [Reference Mierczyński and Shen17].
Theorem 2.1 (Exponential Separation). There are subspace
$\tilde X^{1}_{L}(b), \tilde X^{2}_{L}(b)$
of
$\tilde X_{L}$
such that
$\tilde X^{1}_{L}(b), \tilde X^{2}_{L}(b)$
are continuous in
$b\in \mathrm{Y}$
, and satisfy the following properties:
-
(1)
$\tilde X_{L}=\tilde X^{1}_{L}(b)\oplus \tilde X^{2}_{L}(b)$
for any
$b\in \mathrm{Y}$
. -
(2)
$\tilde X^{1}_{L}(b)=\textrm {span}\{\widetilde {\phi }_{L}(b)\}$
,
$\widetilde {\phi }_{L}(b)\in Int(\tilde X^{+}_{L})$
and is continuous in
$b$
and
$\|\widetilde {\phi }_{L}(b)\|_{\tilde X_{L}}=1$
for any
$b\in \mathrm{Y}$
. -
(3)
$\tilde X^{2}_{L}(b)\cap Int(\tilde X^{+}_{L})=\emptyset$
for any
$b\in \mathrm{Y}$
. -
(4)
$U_{L}(t,b)\tilde X^{1}_{L}(b)=\tilde X^{1}_{L}(\sigma _{t}b)$
and
$U_{L}(t,b)\tilde X^{2}_{L}(b)\subset \tilde X^{2}_{L}(\sigma _{t}b)$
for any
$b\in \mathrm{Y}$
and
$t\gt 0$
. -
(5) There are
$M,\gamma \gt 0$
such that
for any
\begin{equation*} \frac {\|U_{L}(t,b)\widetilde {\phi }_{L}(b)\|_{\tilde X_{L}}}{\|U_{L}(t,b)w\|_{\tilde X_{L}}}\leq Me^{-\gamma t} \end{equation*}
$t\gt 0$
,
$b\in \mathrm{Y}$
and
$w\in \tilde X^{2}_{L}(b)$
with
$\|w\|_{\tilde X_{L}}=1$
.
Let
and
where
$\langle \cdot ,\cdot \rangle$
is the inner product in
$L^{2}(\Omega _{L})$
. Then,
$\kappa _{L}(b)$
is continuous with respect to
$b\in \mathrm{Y}$
. Set
Then, we have
and
Then,
satisfies
\begin{equation} \begin{cases} \dfrac {\partial \phi _{L}(\sigma _{t}b)}{\partial t}=\Delta \phi _{L}(\sigma _{t}b)+\sum ^{N}_{i=1}b_{i}(t,x)\dfrac {\partial \phi _{L}(\sigma _{t}b)}{\partial x_{i}}+b_{0}(t,x)\phi _{L}(\sigma _{t}b)\\[5pt] \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad -\kappa _{L}(\sigma _{t}b)\phi _{L}(\sigma _{t}b),\quad t\gt 0,\,x\in ({-}L,L)\times \omega ,\\[5pt] \partial _{\nu }\phi _{L}(\sigma _{t}b)=0,\quad t\gt 0,\,x\in ({-}L,L)\times \partial \omega ,\\[5pt] \phi _{L}(\sigma _{t}b)({-}L,y)=\phi _{L}(\sigma _{t}b)(L,y)=0,\quad t\gt 0,\,y\in \omega . \end{cases} \end{equation}
Define
$\|U_{L}(t,b)\|=\sup _{u\in \tilde X_L, \, \|u\|_{\tilde X_L}=1}\|U_{L}(t,b)u\|_{\tilde X_L}$
. Note that
$\mathrm{Y}$
is unique ergodic, then
for all
$b\in Y$
.
Definition 2.1 (Principal Lyapunov exponent). The number
is called the principal Lyapunov exponent of (2.1).
Note that
$U_{L}(t,b)$
can act on
$X_L$
, and the principal Lyapunov exponents of
$U_{L}(t,b)$
on
$\tilde X_L$
and
$X_L$
coincide. Then, we have the following
Lemma 2.1.
There is a continuous function
$\phi _L\,:\,Y\to X_L^{+}\setminus \{0\}$
such that the following hold for any
$\tilde b\in Y$
:
-
(1)
$\|\phi _L(\tilde b)\|_{\infty }=1$
, -
(2)
$U_L(t,\tilde b)\phi _L(\tilde b)=\|U_L(t,\tilde b)\phi _L(\tilde b)\|_{\infty }\phi _L(\tilde b\cdot t)$
, -
(3)
$\lim _{t\to \infty }\dfrac {\ln \|U_L(t,\tilde b)\phi _L(\tilde b)\|_{\infty }}{t}=\lambda _L(Y)$
, -
(4) For any
$\phi _{0}\in X_{L}^{+}\setminus \{0\}$
,
$\lim _{t\to \infty }\dfrac {\ln \|U_L(t,\tilde b)\phi _{0}\|_{\infty }}{t}=\lambda _L(Y)$
.
Remark 2.1.
-
(1) Observing that
$\lambda _{L}(\mathrm{Y})$
and
$\phi _{L}(\sigma _{t}b)(x)$
are analogues of principal eigenvalue and principal eigenfunctions of elliptic problem, respectively. We call
$\textrm {span}\{\phi _{L}(\sigma _{t}b)(x)\}_{t\in \mathbb{R}}$
as the principal Floquet bundle of (2.1) associate to the principal Lyapunov exponent.
-
(2) The principal eigenvalue of the following elliptic problem with mixed Dirichlet–Neumann boundary conditions,
(2.3)has been studied in [Reference Berestycki and Rossi3 , Theorem 3.1]. Their results show that for any
\begin{equation} \begin{cases} -\Delta u-\sum ^{N}_{i=1}b_{i}(x)\frac {\partial u}{\partial x_{i}}-b_{0}(x)u=\lambda _{L}u,\quad x\in ({-}L,L)\times \omega ,\\[4pt] \partial _{\nu }u=0\quad \text{on }({-}L,L)\times \partial \omega ,\\[4pt] u({-}L,y)=u(L,y)=0\quad \text{on }y\in \overline {\omega } \end{cases} \end{equation}
$L\gt 0$
, there exists a unique principal eigenvalue
$\lambda _L$
of (2.3) corresponding to a positive eigenfunction
$\psi _{L}\in W^{2,p}(\Omega _{L})$
, for any
$p\gt 1$
. Moreover,
$\lambda _{L}$
converges to
$\lambda _{1}(\Omega )$
as
$L\rightarrow \infty$
, where
$\lambda _{1}(\Omega )$
is the generalised Neumann principal eigenvalue defined by
We point out in
\begin{align*} \lambda _{1}(\Omega )\,:\!=\,\sup \{\lambda \in \mathbb{R}\,:\,\exists \phi \gt 0,\,\Delta \phi +\sum ^{N}_{i=1}b_{i}(x)\frac {\partial \phi }{\partial x_{i}}+b_{0}(x)\phi +\lambda u&\leq 0, \text{ a.e. in } \Omega ,\\ \partial _{\nu }\phi &\geq 0 \text{ on } \partial \Omega \}. \end{align*}
$\lambda _{1}(\Omega )$
, the function
$\phi$
is understood to belong to
$W^{2,p}(({-}L,L)\times \omega )$
for some
$p\gt N$
and every
$L\gt 0$
.
Lemma 2.2.
Set
$\lambda (L)\,:\!=\,\lambda _{L}(\mathrm{Y})$
. Then, the function
$\lambda (L)\,:\,\mathbb{R}^{+}\rightarrow \mathbb{R}$
is nondecreasing in
$L$
and
$\lim _{L\rightarrow \infty }\lambda (L)$
exists.
Proof. For any
$L\gt 0$
, let
$\phi _{L}(\sigma _{t}b)(x)$
be as in Lemma 2.1. Let
$0\lt L_{1}\lt L_{2}$
be fixed. Then, there exists
$\epsilon _{0}\in (0,1)$
such that
It then follows from the comparison principle that
Then
which implies that
$\lambda (L)$
is non-decreasing in
$L$
.
Let
$b_{L}\,:\!=\,\sup _{t\in \mathbb{R},x\in \Omega _{L}}b_{0}(t,x)$
. Then,
$e^{b_{L}t}$
is a supersolution of (2.1), and thus,
for all
$t\geq 0$
. It then follows that
Hence,
$\lim _{L\rightarrow \infty }\lambda (L)$
exists. This completes the proof.
From Lemma 2.2, we know that
$\lim _{L\rightarrow \infty }\lambda _{L}(\mathrm{Y})$
exists. Set
Then
$\lambda _{\infty }(\mathrm{Y})\lt \sup _{t\in \mathbb{R},x\in \Omega }b_{0}(t,x)$
and
$\lambda _{\infty }(\mathrm{Y})$
is called the approximate principal Lyapunov exponent of the following equation
\begin{equation} \begin{cases} \partial _{t}u=\Delta u+\sum ^{N}_{i=1}b_{i}(t,x)\frac {\partial u}{\partial x_{i}}+b_{0}(t,x)u,\quad x\in \mathbb{R}\times \omega ,\\[4pt] \partial _{\nu }u=0\quad \text{on }\mathbb{R}\times \partial \omega . \end{cases} \end{equation}
Let
$X$
and
$X^{+}$
be as in Section 1. Denote by
$U(t,b)u_{0}$
the unique classical solution of (2.4) with the initial value
$u_{0}\in X$
and define
$\|U(t,b)\|=\sup _{u\in X, \, \|u\|=1}\|U(t,b)u\|$
. It follows from the almost-periodicity of
$b(t,x)$
and the subadditive ergodic theorem (see [Reference Kingman14]) that
$\lim _{t\rightarrow \infty }\frac {\ln {U(t,b)}}{t}$
exists for all
$b\in Y$
. Then, we have the following definition for the principal Lyapunov exponent of (2.4).
Definition 2.2 (Principal Lyapunov exponent). The number
is called the principal Lyapunov exponent of (2.4).
In the following, we show that the connections between the approximate principal Lyapunov exponent
$\lambda _{\infty }(\mathrm{Y})$
and the principal Lyapunov exponent
$\lambda (\mathrm{Y})$
.
Lemma 2.3.
$\lambda _{\infty }(\mathrm{Y})$
and
$\lambda (\mathrm{Y})$
have the same sign. In particular,
$\lambda _{\infty }(\mathrm{Y})=\lambda (\mathrm{Y})$
for
$\lambda _{\infty }(\mathrm{Y})\geq \lim _{r\rightarrow \infty }\sup _{|x_{1}|\gt r,t\in \mathbb{R},y\in \omega }b_{0}(t,x_{1},y)$
.
Proof. In the following proof, to emphasis the dependence of
$\mathrm{Y}$
on
$b$
, we write it as
$\mathrm{Y}(b)$
. We divide the proof into two cases:
$\lambda _{\infty }(\mathrm{Y}(b))\geq \lim _{r\rightarrow \infty }\sup _{|x_{1}|\gt r,t\in \mathbb{R},y\in \omega }b_{0}(t,x_{1},y)$
and
$\lambda _{\infty }(\mathrm{Y}(b))\lt \lim _{r\rightarrow \infty }\sup _{|x_{1}|\gt r,t\in \mathbb{R},y\in \omega }b_{0}(t,x_{1},y)$
.
Firstly, we consider the case
$\lambda _{\infty }(\mathrm{Y}(b))\geq \lim _{r\rightarrow \infty }\sup _{|x_{1}|\gt r,t\in \mathbb{R},y\in \omega }b_{0}(t,x_{1},y)$
. By definitions of
$\lambda _{\infty }(Y(b))$
and
$\lambda (Y(b))$
, we have
$\lambda _{\infty }(Y(b))\leq \lambda (Y(b))$
. Assume that
$\lambda _{\infty }(Y(b))\lt \lambda (Y(b))$
. Without loss of generality, we may assume that
In fact, we can take
$\lambda _{0}\in (\lambda _{\infty }(Y(b)),\lambda (Y(b)))$
and replace
$b_{0}(t,x)$
by
$\overline {b}_{0}(t,x)=b_{0}(t,x)-\lambda _{0}$
. Then
and
Consider
\begin{equation} \begin{cases} \partial _{t}u=\Delta u+\sum ^{N}_{i=1}a_{i}(t,x)\frac {\partial u}{\partial x_{i}}+ug(t,x,u),\quad x\in \mathbb{R}\times \omega ,\\[4pt] \partial _{\nu }u=0\quad \text{on }\mathbb{R}\times \partial \omega . \end{cases} \end{equation}
Denote
$v(t,x;\,v_{0})$
be the solution of (2.5) with initial value
$v(t,x;\,v_{0})=v_{0}\in X^{+}$
. Choose
$v^{*}_{0}\equiv 1$
, similar to that in Theorem 3.1 later, we have that
$\lim _{t\rightarrow \infty }\|v(t,\cdot ;\,v^{*}_{0})\|_{\infty }=0$
for
$\lambda _{\infty }(Y(b))\lt 0$
. Note that
$\lambda (Y(b))\gt 0$
and
$g$
is uniformly continuous on
$\mathbb{R}\times \Omega \times [0,\delta ]$
for any
$\delta \gt 0$
, then there exists large
$T\gt 0$
such that
for any
$t\geq T$
and
$x\in \mathbb{R}\times \omega$
. By comparison principle,
Due to that
$v^{*}_{0}$
is uniform positive and
$g$
is bounded on
$\mathbb{R}\times \Omega \times [0,\delta ]$
for any
$\delta \gt 0$
,
Then for any
$v_{0}\in X^{+}$
with
$\|v_{0}\|=1$
, there is a constant
$M\gt 0$
such that
Then the comparison principle implies that
which implies that
Then by the definition of
$\lambda (Y(b))$
and together with (2.6), we have that
which is a contradiction. Therefore,
$\lambda _{\infty }(Y(b))=\lambda (Y(b))$
and they have the same sign in the case
$\lambda _{\infty }(\mathrm{Y}(b))\geq \lim _{r\rightarrow \infty }\sup _{|x_{1}|\gt r,t\in \mathbb{R},y\in \omega }b_{0}(t,x_{1},y)$
.
Secondly, we consider the case
$\lambda _{\infty }(\mathrm{Y}(b))\lt \lim _{r\rightarrow \infty }\sup _{|x_{1}|\gt r,t\in \mathbb{R},y\in \omega }b_{0}(t,x_{1},y)$
. In this case, we show that if
$\lambda _{\infty }(Y(b))\lt 0$
, then
$\lambda (Y(b))\lt 0$
. Suppose that
$\lambda (Y(b))\geq 0$
. Then there exists
$\lambda _{0}\gt 0$
such that
Let
$a(t,x)=g(t,x,0)+\lambda _{0}$
. Then,
Similar to above, we have
which is a contradiction. Thus
$\lambda (Y(b))\lt 0$
. This completes the proof.
Proposition 2.1.
Suppose that
$b_{i}(t,x)\ (i=1,\ldots ,N)$
and
$b_{0}(t,x)$
are periodic in
$t$
. If
$\lambda _{\infty }\gt \lim _{r\rightarrow \infty }\sup _{|x_{1}|\gt r,t\in \mathbb{R},y\in \omega }b_{0}(t,x_{1},y)$
, then
$\lambda _{\infty }$
is the principal eigenvalue (i.e. the eigenvalue with large real part) of the following linear periodic equation
\begin{equation*} \begin{cases} \frac {\partial u}{\partial t}=\Delta u+\sum ^{N}_{i=1}b_{i}(t,x)\frac {\partial u}{\partial x_{i}}+b_{0}(t,x)u,\quad t\gt 0,\,x\in \mathbb{R}\times \omega ,\\[5pt] \partial _{\nu }u=0,\quad t\gt 0,\,x\in \mathbb{R}\times \partial \omega . \end{cases} \end{equation*}
Proof. Since
$b_{i}(t,x)\ (i=1,\ldots ,N)$
and
$b_{0}(t,x)$
are periodic in
$t$
, by the definition of
$\lambda _{L}$
, we know that
$\lambda _{L}$
is the principal eigenvalue of (2.1) for any
$L\gt 0$
. Thus, there exists a bounded positive time periodic eigenfunction
$\phi _{L}$
of (2.1) corresponding to
$\lambda _{L}$
, that is,
\begin{equation} \begin{cases} \frac {\partial \phi _{L}}{\partial t}=\Delta \phi _{L}+\sum ^{N}_{i=1}b_{i}(t,x)\frac {\partial \phi _{L}}{\partial x_{i}}+b_{0}(t,x)\phi _{L}-\lambda _{L}\phi _{L},\quad t\gt 0,\,x\in ({-}L,L)\times \omega ,\\[5pt] \partial _{\nu }\phi _{L}=0,\quad t\gt 0,\,x\in ({-}L,L)\times \partial \omega ,\\[5pt] \phi _{L}(t,-L,y)=\phi _{L}(t,L,y)=0,\quad t\gt 0,\,y\in \omega ,\\[5pt] \phi _{L}\text{ is periodic in }t. \end{cases} \end{equation}
Normialised
$\phi _{L}$
by
$\max _{\mathbb{R}\times ({-}L,L)\times \omega }\phi _{L}(t,x_{1},y)=1$
and let
$(t_{L},x_{1L},y_{L})\in [0,T]\times ({-}L,L)\times \omega$
be such that
$\phi _{L}(t_{L},x_{1L},y_{L})=1$
. Extending by reflection the function
$\phi _{L}$
to large cylinders, as done in [Reference Berestycki and Rossi3, Appendix A], then
$\phi _{L}$
is uniformly bounded in
$\mathbb{R}\times ({-}L,L)\times \omega$
. Hence, by standard parabolic regularity estimates, there exists
$\phi _{\infty }$
such that
$\phi _{L}(t,x)\rightarrow \phi _{\infty }(t,x)$
locally uniformly in
$\mathbb{R}\times ({-}L,L)\times \omega$
as
$L\rightarrow \infty$
, where
$\phi _{\infty }$
is a bounded, nonnegative and time periodic function satisfying
\begin{equation} \begin{cases} \frac {\partial \phi _{\infty }}{\partial t}=\Delta \phi _{\infty }+\sum ^{N}_{i=1}b_{i}(t,x)\frac {\partial \phi _{\infty }}{\partial x_{i}}+b_{0}(t,x)\phi _{\infty }-\lambda _{\infty }\phi _{\infty },\quad t\gt 0,\,x\in \mathbb{R}\times \omega ,\\[4pt] \partial _{\nu }\phi _{\infty }=0,\quad t\gt 0,\,x\in \mathbb{R}\times \partial \omega ,\\[4pt] \phi _{\infty }\text{ is periodic in }t. \end{cases} \end{equation}
Consider (2.7) at point
$(t_{L},x_{1L},y_{L})$
, we have
Since
$\lambda _{\infty }\gt \lim _{r\rightarrow \infty }\sup _{|x_{1}|\gt r,t\in \mathbb{R},y\in \omega }b_{0}(t,x_{1},y)$
, we must have that
$x_{1L}$
is bounded as
$L\rightarrow \infty$
. Thus,
$(t_{L},x_{1L},y_{L})\rightarrow (t^*,x^{*}_{1},y^{*})$
as
$L\rightarrow \infty$
, and
$\phi _{\infty }(t^*,x^{*}_{1},y^{*})=1$
. The strong maximum principal yields that
$\phi _{\infty }\gt 0$
in
$\mathbb{R}\times \Omega$
. Then,
$\lambda _{\infty }$
is the principal eigenvalue of (2.8) with principal eigenfunction
$\phi _{\infty }$
.
3. Large-time dynamical behaviour in infinite cylinder
In this section, we consider the long-time dynamical behaviour of the solution
$u(t,x_{1},y;\,u_{0})$
of (1.1) in the infinite cylinder and prove Theorem 1.1.
Let
$X$
and
$X^{+}$
be given in Section 1. By the semigroup theory, for any initial value
$v_{0}\in X^{+}$
, (1.2) admits a unique solution
$v(t,x;\,v_{0})$
with
$v(0,x;\,v_{0})=v_{0}$
. Set
$S=\max \{M_{0},\|v_{0}\|_{L^{\infty }(\Omega )}\}$
, where
$M_{0}$
is the positive constant in (H3). Due to that
$0$
and
$S$
are a sub-and supersolution of (1.2) with initial value
$v_{0}\in X^{+}$
, then
$0\leq v(t,x;\,v_{0})\leq S$
in
$\overline {\Omega }$
. By extending
$v(t,x;\,v_{0})$
to a large cylinder
$(\mathbb{R}\times \overline {\omega })\supset \supset \Omega$
through reflection (see [Reference Berestycki and Rossi3, Appendix A]) and applying the parabolic maximum principle, we find that
Take
$b_{1}(t,x)=c(t)$
,
$b_{i}(t,x)=0\, (i=2,\ldots ,N)$
and
$b_{0}(t,x)=g(t,x,0)$
in (2.1). Then for any
$L\gt 0$
, we have the following linear problem
\begin{equation} \begin{cases} \partial _{t}u=\Delta u+c(t)\partial _{1}u+g(t,x,0)u,\quad t\gt 0,\,x\in ({-}L,L)\times \omega ,\\ \partial _{\nu }u=0,\quad t\gt 0,\,x\in ({-}L,L)\times \partial \omega ,\\ u(t,-L,y)=u(t,L,y)=0,\quad t\gt 0,\,y\in \omega . \end{cases} \end{equation}
By Definition 2.1, let
$\lambda _{L}$
be the principal Lyapunov exponent of (3.1). Then, all results in Section 2 hold true for
$\lambda _{L}$
, particularly,
$\lambda _{\infty }\,:\!=\,\lim _{L\rightarrow \infty }\lambda _{L}$
.
Theorem 3.1.
Assume that (H1)-(H4) hold true. If
$\lambda _{\infty }\lt 0$
, then for any
$v_{0}\in X^{+}$
,
uniformly for
$(x_{1},y)\in \Omega$
.
Proof. Before proving (3.2), we first show that
For any fixed
$v_{0}\in X^{+}$
, we assume that (3.3) does not hold. Then there exist
$\epsilon _{0}\gt 0$
,
$\{t_{n}\}\subset (0,\infty )$
and
$\{x_{1,n}\}\subset \mathbb{R}$
with
$t_{n}\rightarrow \infty$
and
$|x_{1,n}|\rightarrow \infty$
, as
$n\rightarrow \infty$
, such that
Let
$v^{n}(t,x_{1},y)=v(t+t_{n},x_{1}+x_{1,n},y;\,v_{0})$
. Then
$v^{n}$
satisfies
\begin{equation} \begin{cases} \partial _{t}v^{n}=\Delta v^{n}+c(t+t_{n})\partial _{1}v^{n}+f(t+t_{n},x_{1}+x_{1,n},y,v^{n}), \quad t\gt 0, (x_{1},y)\in \Omega ,\\ \partial _{\nu } v^{n}(t,x_{1},y)=0, \quad t\in \mathbb{R},(x_{1},y)\in \partial \Omega . \end{cases} \end{equation}
Then parabolic estimates and the almost periodicity of
$c(t)$
and
$f(t,x,v)$
imply that (up to subsequence) there are
$v^{*}$
,
$c^{*}$
and
$f^{*}$
such that
\begin{align*} \lim _{n\rightarrow \infty }v^{n}(t,x_{1},y)=&v^{*}(t,x_{1},y)\quad \text{locally uniformly in }t\in \mathbb{R},(x_{1},y)\in \Omega ,\\ \lim _{n\rightarrow \infty }f(t+t_{n},x_{1}+x_{1,n},y,v^{n})=&f^{*}(t,x_{1},y,v^{*})\quad \text{locally uniformly in }t\in \mathbb{R},(x_{1},y)\in \Omega \end{align*}
and
where
$v^{*}(t,x_{1},y)$
is a nonnegative bounded solution of
\begin{equation} \begin{cases} \partial _{t}v^{*}=\Delta v^{*}+c(t)\partial _{1}v^{*}+f(t,x_{1},y,v^{*}),\quad t\in \mathbb{R},(x_{1},y)\in \Omega ,\\ \partial _{\nu } v^{*}=0, \quad t\in \mathbb{R},(x_{1},y)\in \partial \Omega \end{cases} \end{equation}
and
Note that
$f(t,x,u)\leq f_{u}(t,x,0)u=g(t,x,0)u$
and
$\lim _{x_{1}\rightarrow \infty }\sup _{(t,y)\in \mathbb{R}\times \omega }g(t,x_{1},y,0)\lt 0$
. Choose
$\alpha \gt 0$
such that
Recall that
$S=\max \{M_{0},\|v_{0}\|_{L^{\infty }(\Omega )}\}$
. For any fixed
$t_{0}\in \mathbb{R}$
,
$w(t,x)=e^{-\alpha (t-t_{0})}S$
is a solution of
\begin{equation} \begin{cases} \partial _{t}w= \Delta w+c(t)\partial _{1}w-\alpha w,\quad t\in \mathbb{R}, (x_{1},y)\in \Omega ,\\ \partial _{\nu } w=0, \quad t\in \mathbb{R}, (x_{1},y)\in \partial \Omega . \end{cases} \end{equation}
We also have that
$v^{*}(t,x_{1},y)$
is a subsolution of (3.6). The comparison principle implies that
particularly,
$v^{*}(0,x_{1},y)\leq e^{\alpha t_{0}S}$
for any
$(x_{1},y)\in \Omega$
and
$t_{0}\lt 0$
. Choose
$t_{0}\ll -1$
such that
$e^{\alpha t_{0}}S\leq \epsilon _{0}$
, then
$v^{*}(0,x_{1},y)\lt \epsilon _{0}$
for
$(x_{1},y)\in \Omega$
, which is a contradiction. Therefore, for any
$\epsilon \gt 0$
, there exist constants
$T_{\epsilon }\gt 0$
and
$L(\epsilon )\gt 0$
such that
and (3.3) hold true.
Given
$v_{0}\in X^{+}$
,
$\epsilon \gt 0$
, let
Then,
$v^{\epsilon }(t,x_{1},y)$
satisfies
\begin{equation} \begin{cases} \partial _{t}v^{\epsilon }= \Delta v^{\epsilon }+c(t)\partial _{1}v^{\epsilon }+f(t,x,v^{\epsilon }+\epsilon ),\quad t\in \mathbb{R},x\in \Omega ,\\[3pt] \partial _{\nu }v^{\epsilon }=0,\quad t\in \mathbb{R},x\in \partial \Omega . \end{cases} \end{equation}
By the assumption on
$f$
, we have
$f(t,x,v^{\epsilon }+\epsilon )\leq g(t,x,0)(v^{\epsilon }+\epsilon )$
. Then
$v^{\epsilon }(t,x_{1},y)$
satisfies
\begin{equation} \begin{cases} \partial _{t}v^{\epsilon }\leq \Delta v^{\epsilon }+c(t)\partial _{1}v^{\epsilon }+g(t,x,0)(v^{\epsilon }+\epsilon ),\quad t\in \mathbb{R},x\in \Omega ,\\[3pt] \partial _{\nu }v^{\epsilon }=0,\quad t\in \mathbb{R},x\in \partial \Omega . \end{cases} \end{equation}
Since
$\lambda _{\infty }\lt 0$
, there exists
$L\gt 0$
such that
$\lambda _{L}\lt 0$
. Let
$\eta \,:\,\mathbb{R}\rightarrow [0,1]$
be a smooth function with
$\eta (x_{1})=1$
for
$|x_{1}|\leq \frac {L}{2}$
and
$\eta (x_{1})=0$
for
$|x_{1}|\geq \frac {3L}{4}$
. Define
Clearly,
$v^{\epsilon }_{0}(x_{1},y)\in X^{+}$
. Let
$v(t,x_{1},y;\,v^{\epsilon }_{0})$
be the unique solution of the following problem
\begin{equation*} \begin{cases} \partial _{t}v= \Delta v+c(t)\partial _{1}v+g(t,x,0)(v+\epsilon ),\quad t\gt 0, x\in ({-}L,L)\times \omega ,\\ \partial _{\nu }v=0,\quad t\gt 0, x\in ({-}L,L)\times \partial \omega ,\\ v(t,-L,y)=v(t,L,y)=0,\quad t\gt 0, y\in \omega \end{cases} \end{equation*}
with initial value
$v(0,x_{1},y;\,v^{\epsilon }_{0})=v^{\epsilon }_{0}(x_{1},y)$
,
$(x_{1},y)\in [{-}L,L]\times \omega$
. Let
$U_{L}(t,c)$
be the solution operator of
\begin{equation*} \begin{cases} \partial _{t}w= \Delta w+c(t)\partial _{1}w+f_{u}(t,x,0)w,\quad t\gt 0, x\in ({-}L,L)\times \omega ,\\ \partial _{\nu }w=0,\quad t\gt 0, x\in ({-}L,L)\times \partial \omega ,\\ w(t,-L,y)=w(t,L,y)=0\quad t\gt 0, y\in \omega . \end{cases} \end{equation*}
Then by the variation of constants formula, we have
Since
$\lambda _{L}\lt 0$
, by Definition 2.1, there exists
$M\gt 0$
such that
for
$t\geq 0$
. Together with (3.10), we have
Note that
and
Then, by (3.9) and the comparison principle, we obtain that
By (3.11),
Due to
$\lambda _{L}\lt 0$
, Definition 2.1 ensures that there exists
$T_{\epsilon }\gt 0$
such that
Hence,
for all
$t\geq T_{\epsilon }$
,
$x_{1}\in ({-}L,L)$
and
$y\in \omega$
. By (3.9), choose large
$T\gt 0$
and
$L\gt 0$
,
$v(t,x_{1},y;\,v_{0})\lt \epsilon$
for
$t\geq T$
,
$|x_{1}|\gt \frac {L}{2}$
and
$y\in \omega$
. This, together with (3.12), implies that
By the arbitrariness of
$\epsilon$
, we obtain
$\lim _{t\rightarrow \infty }v(t,x_{1},y;\,v_{0})=0$
uniformly for
$(x_{1},y)\in \Omega$
. This completes the proof.
Consider the following mixed Dirichlet–Neumann boundary value problem
\begin{equation} \begin{cases} \partial _{t}w= \Delta w+c(t)\partial _{1}w+f(t,x,w),\quad t\gt 0, x\in ({-}L,L)\times \omega ,\\ \partial _{\nu }w=0,\quad t\gt 0, x\in ({-}L,L)\times \partial \omega ,\\ w(t,-L,y)=w(t,L,y)=0,\quad t\gt 0, y\in \omega . \end{cases} \end{equation}
For any
$w_{0}\in X^{+}_{L}$
, we denote by
$w(t,x_{1},y;\,w_{0})$
the unique solution of (3.13) with
$w(0,x_{1},y;\,w_{0})=w_{0}$
. Then we have the following proposition on the dynamics of (3.13).
Proposition 3.1.
Assume that (H1)-(H4) hold true and
$\lambda _{L}\gt 0$
. Then (3.13) admits a unique positive almost periodic solution
$w^{*}(t,x)$
and for any
$w_{0}\in X^{+}_{L}\setminus \{0\}$
, there holds
Proof. It can be proved by properly modifying the arguments in [Reference Zhao32, Theorem 3.1]. For completeness, we provide the proof in the following.
Let
$\mathcal{H}(c,f)$
be the closure of
$\{(c\cdot s,f\cdot s)\,:\, s\in \mathbb{R}\}$
with the compact open topology. For any
$(\widetilde {c},\widetilde {f})\in \mathcal{H}(c,f)$
, assumptions (H1)-(H4) imply that
$w=M$
,
$M\geq M_{0}$
, is an supersolution of (3.13). Hence, by the comparison principle and a priori estimates for parabolic equations, each solution
$w(t,x_{1},y;\,w_{0},\widetilde {c},\widetilde {f})$
of (3.13) exists globally on
$[0,\infty )$
, and for any
$t\gt 0$
, the set
$\{w(t,x_{1},y;\,w_{0},\widetilde {c},\widetilde {f})\,:\,t\gt 0\}$
is precompact in
$X^{+}_{L}$
. Define the skew-product semiflow
$\Pi _{t}\,:\,X^{+}_{L}\times \mathcal{H}(c,f)\rightarrow \Pi _{t}\,:\,X^{+}_{L}\times \mathcal{H}(c,f)$
by
Then, the omega limit set
$\omega (w_{0},\widetilde {c},\widetilde {f})$
of the forward orbit
$\gamma ^{+}(w_{0},\widetilde {c},\widetilde {f})\,:\!=\,\{\Pi _{t}(w_{0},\widetilde {c},\widetilde {f})\,:\,t\geq 0\}$
is well defined, compact and invariant under
$\Pi _{t}, t\geq 0$
, for each
$(w_{0},\widetilde {c},\widetilde {f})\in X^{+}_{L}\times \mathcal{H}(c,f)$
.
We first show that there exists a
$\delta \gt 0$
such that
We set
$\lambda _{L}$
of (3.1) as
$\lambda _{L}(g)$
. Since
$\lambda _{L}(g)\gt 0$
, then we can choose a sufficient small
$\varepsilon _{0}\gt 0$
such that
Since
$f$
is uniformly almost periodic in
$t$
and
$\mathcal{H}(f)$
is compact, there exists
$\delta _{0}\gt 0$
such that
Suppose on the contrary that there is
$(w_{0},\widetilde {c},\widetilde {f})\in X^{+}_{L}\times \mathcal{H}(c,f)$
, such that
Then, there exists
$t_{0}\gt 0$
such that
Since
$w(t_{0},x_{1},y;\,w_{0},\widetilde {c},\widetilde {f})\gt 0$
, then
where
$w_{*}=w(t_{0},x_{1},y;\,w_{0},\widetilde {c},\widetilde {f})$
,
$\widetilde {c}_{t_{0}}=c(t+t_{0})$
and
$\widetilde {f}_{t_{0}}=f(t+t_{0},x,w)$
. Then
$w\big(t,x_{1},y;\,w_{*},\widetilde {c}_{t_{0}},\widetilde {f}_{t_{0}}\big)$
satisfies the following differential inequality
\begin{equation*} \begin{cases} \partial _{t}w\geq \Delta w+\widetilde {c}(t)\partial _{1}w+(f_{u}(t,x_{1},y,0)-\varepsilon _{0})w,\quad t\gt 0, x\in ({-}L,L)\times \omega ,\\ \partial _{\nu }w=0,\quad t\gt 0, x\in ({-}L,L)\times \partial \omega ,\\ w(t,-L,y)=w(t,L,y)=0,\quad t\gt 0,y\in \omega . \end{cases} \end{equation*}
Similar to (2.2),
$\lambda _{L}(f_{u}-\varepsilon _{0})=\lim _{t\rightarrow \infty }\frac {1}{t}\int ^{t}_{0}k_{L}(\sigma _{\tau }f_{u})d\tau$
and
$e^{\int ^{t}_{0}k_{L}(\sigma _{\tau }f_{u})d\tau }\phi _{L}(\sigma _{t}f_{u})$
is a solution of
\begin{equation*} \begin{cases} \partial _{t}w= \Delta w+\widetilde {c}(t)\partial _{1}w+(f_{u}(t,x_{1},y,0)-\varepsilon _{0})w,\quad t\gt 0, x\in ({-}L,L)\times \omega ,\\ \partial _{\nu }w=0,\quad t\gt 0, x\in ({-}L,L)\times \partial \omega ,\\ w(t,-L,y)=w(t,L,y)=0,\quad t\gt 0,y\in \omega . \end{cases} \end{equation*}
Choose a sufficiently small
$\varepsilon \gt 0$
such that
$w_{*}\geq \varepsilon \phi _{L}$
. By the standard comparison theorem, we have
Letting
$t\rightarrow \infty$
, we have
which is a contradiction.
Following from the invariance of
$\omega (w_{0},\widetilde {c},\widetilde {f})$
and
$\Pi _{t}((X^{+}_{L}\setminus \{0\})\times \mathcal{H}(c,f))\subset int{X^{+}_{L}\times \mathcal{H}(c,f)}$
,
$\forall t\gt 0$
, we have
For each
$(\widetilde {c},\widetilde {f},t)\in \mathcal{H}(c,f)\times (0,\infty )$
, for any
$w^{\prime }_{0}(\cdot ),\, w^{\prime \prime }_{0}(\cdot )\in X^{+}_{L}$
, when
$w^{\prime }_{0}(\cdot )\gt w^{\prime \prime }_{0}(\cdot )$
, by comparison principal, we have
Then,
$w(t,\cdot ;\,\cdot ,\widetilde {c},\widetilde {f})$
is strongly monotone on
$X^{+}_{L}$
for each
$(\widetilde {c},\widetilde {f},t)\in \mathcal{H}(c,f)\times (0,\infty )$
. Moreover, by assumptions (H1)–(H4),
$w(t,\cdot ;\,\cdot ,\widetilde {c},\widetilde {f})$
is also subhomongenous on
$X^{+}_{L}$
for each
$(\widetilde {c},\widetilde {f},t)\in \mathcal{H}(c,f)\times (0,\infty )$
. Let
$w^{\prime }_{0}\in X^{+}_{L}\setminus \{0\}$
be fixed and
$K_{0}=\omega (w^{\prime }_{0},\widetilde {c},\widetilde {f})$
. By [Reference Zhao32, Theorem 2.1] and (3.15), for every
$w_{0}\in X^{+}_{L}\setminus \{0\}$
,
where
$(w^{*}_{0},\widetilde {c},\widetilde {f})\in K_{0}$
. Therefore,
$w(t,\cdot ;\,w^{*}_{0},\widetilde {c},\widetilde {f})$
is a unique positive almost periodic solution of (3.13). This completes the proof.
Theorem 3.2.
Assume that (H1)-(H4) hold true. If
$\lambda _{\infty }\gt 0$
, then for any
$L\gt 0$
and
$v_{0}\in X^{+}\setminus \{0\}$
,
Proof. By the definition of
$\lambda _{\infty }$
, due to
$\lambda _{\infty }\gt 0$
, there is
$L^{*}\gt 0$
such that
$\lambda _{L}\gt 0$
for any
$L\gt L^{*}$
. Then, by Proposition 3.1, there is a unique positive almost periodic solution
$w^{*}(t,x)$
of (3.13). Note that any nonnegative solution of (1.2) is a supersolution of (3.13). Then,
Then by (3.14), we have
This completes the proof.
4. Forced wave solutions
In this section, we study the existence, uniqueness, stability, and exponential decay of forced wave solutions of (1.1). By (1.5), in the following, we need to prove that there is an entire solution
$V(t,x_{1},y)$
of (1.2), which is bounded, positive and almost periodic in
$t$
for any
$(x_{1},y)\in \Omega =\mathbb{R}\times \omega$
.
Lemma 4.1.
Assume that (H1)-(H4) holds true. When
$\lambda _{\infty }\gt 0$
, equation (1.2), that is,
\begin{equation} \begin{cases} \partial _{t}v=\Delta v+c(t)\partial _{1}v+f(t,x_{1},y,v),\quad (x_{1},y)\in \mathbb{R}\times \Omega ,\\ \partial _{\nu }v=0\quad \text{on } \partial \Omega , \end{cases} \end{equation}
admits a unique bounded positive solution
$V(t,x_{1},y)$
with
Proof. Let
$M_{0}$
be given in (H3). Then
$\overline {v}(t,x)=M_{0}$
is a supersolution of (4.1). Let
$v(t,x_{1},y;\,\overline {v},c,f)$
be a solution of (1.2) with the initial condition
$v(0,x_{1},y;\,\overline {v},c,f)=\overline {v}$
. By comparison principle,
$v(t,x_{1},y;\,\overline {v},c,f)\leq \overline {v}(t,x)$
for
$x\in \Omega$
and
$t\gt 0$
. Since
$c$
and
$f$
are almost periodic function, there exists
$t_{n}\rightarrow \infty$
such that
$c\cdot t_{n}\rightarrow c$
and
$f\cdot t_{n}\rightarrow f$
uniformly as
$n\rightarrow \infty$
. By a priori estimates for parabolic equations, there exists a function
$V(t,x)$
such that
$v(t+t_{n},x_{1},y;\,\overline {v},c,f)$
converges to
$V(t,x_{1},y)$
locally uniformly in
$(t,x_{1},y)\in \mathbb{R}\times \Omega$
as
$n\rightarrow \infty$
. Then,
$V(t,x_{1},y)$
is a bounded entire solution satisfying (4.1) for
$t\in \mathbb{R}$
and
$(x_{1},y)\in \Omega$
. By Theorem 3.2, we have
$\inf _{\mathbb{R}\times [{-}L,L]\omega }V(t,x_{1},y)\gt 0$
for all
$L\gt 0$
.
Following from the arguments for (3.3), we also have
$\lim _{|x_{1}|\rightarrow \infty }\sup _{t\in \mathbb{R},y\in \omega }V(t,x_{1},y)=0$
, that is, (4.2) holds true. In the following, we prove the uniqueness of
$V(t,x_{1},y)$
. Suppose that there are two bounded positive solutions
$V(t,x_{1},y)$
and
$V^{\prime }(t,x_{1},y)$
of (4.1). We first show that
$V(t,x_{1},y)\leq V^{\prime }(t,x_{1},y)$
. For any
$\epsilon \gt 0$
, define
By (4.2), there exists
$R(\epsilon )\gt 0$
such that
Together with
$V^{\prime }\gt 0$
in
$\mathbb{R}\times \overline {\Omega }$
, we find that the set
$K_{\epsilon }$
is nonempty. Set
$k(\epsilon )\,:\!=\,\inf \{k\,:\,k\in K_{\epsilon }\}$
. Then,
$k\,:\,\mathbb{R}^{+}\rightarrow \mathbb{R}$
is a nonincreasing function and thus
$\lim _{\epsilon \rightarrow 0^{+}}k(\epsilon )$
exists. Set
$k^{*}=\lim _{\epsilon \rightarrow 0^{+}}k(\epsilon )$
. Assume that
For any
$0\lt \epsilon \lt \sup _{\mathbb{R}\times \Omega }V$
small enough, we have
$k(\epsilon )\gt 0$
,
and there exists a sequence
$\{t^{\epsilon }_{n}\}\subset \mathbb{R}$
such that
Without loss of generality, we assume that there are
$V^{\prime }_{\epsilon }(t,x_{1},y)$
,
$V_{\epsilon }(t,x_{1},y)$
and
$(c^{\epsilon },f^{\epsilon })\in \mathcal{H}(c,f)$
such that
and
locally uniformly in
$(t,x_{1},y)\in \mathbb{R}\times \mathbb{R}\times \omega$
,
$u\gt 0$
and
Then,
$V_{\epsilon }$
and
$V^{\prime }_{\epsilon }$
are bounded positive solution of (1.2) with
$c$
and
$f$
replaced by
$c^{\epsilon }$
and
$f^{\epsilon }$
, respectively. Moreover, there hold
$k_{\epsilon }V^{\prime }_{\epsilon }\geq V_{\epsilon }-\epsilon$
and
By (4.2) again, there is
$(x^{\epsilon }_{1},y^{\epsilon })\in \mathbb{R}\times \omega$
such that
Then, the function
$t\rightarrow k_{\epsilon }V^{\prime }_{\epsilon }(t,x^{\epsilon }_{1},y^{\epsilon })-V(t,x^{\epsilon }_{1},y^{\epsilon })+\epsilon$
attains its minimum value at
$0$
and
$k_{\epsilon }V^{\prime }_{\epsilon }(0,x_{1},y)-V(0,x_{1},y)+\epsilon$
attains its minimum value at
$(x^{\epsilon }_{1},y^{\epsilon })$
.
Then, at point
$\left(0,x^{\epsilon }_{1},y^{\epsilon }\right)$
, by (H1)-(H4), we have
\begin{align*} 0 & = k_{\epsilon }\frac {\partial V^{\prime }_{\epsilon }}{\partial t}\left(0,x^{\epsilon }_{1},y^{\epsilon }\right)-\frac {\partial V_{\epsilon }}{\partial t}\left(0,x^{\epsilon }_{1},y^{\epsilon }\right)\\ &=\left[k_{\epsilon }\Delta V^{\prime }_{\epsilon }\left(0,x^{\epsilon }_{1},y^{\epsilon }\right)+c^{\epsilon }(0)\partial _{1}\left(k_{\epsilon }V^{\prime }_{\epsilon }\left(0,x^{\epsilon }_{1},y^{\epsilon }\right)\right)+f^{\epsilon }\left(0,x^{\epsilon }_{1},y^{\epsilon },k_{\epsilon }V^{\prime }_{\epsilon }\right)\right]\\ & \quad -\left[\Delta V_{\epsilon }\left(0,x^{\epsilon }_{1},y^{\epsilon }\right)+c^{\epsilon }(0)\partial _{1}V_{\epsilon }\left(0,x^{\epsilon }_{1},y^{\epsilon }\right)+f^{\epsilon }\left(0,x^{\epsilon }_{1},y^{\epsilon },V_{\epsilon }\right)\right]\\ &=\left[k_{\epsilon }\Delta V^{\prime }_{\epsilon }\left(0,x^{\epsilon }_{1},y^{\epsilon }\right)-\Delta V_{\epsilon }\left(0,x^{\epsilon }_{1},y^{\epsilon }\right)\right]+c^{\epsilon }(0)\left[\partial _{1}\left(k_{\epsilon }V^{\prime }_{\epsilon }\left(0,x^{\epsilon }_{1},y^{\epsilon }\right)\right)-\partial _{1}V_{\epsilon }\left(0,x^{\epsilon }_{1},y^{\epsilon }\right)\right]\\ &\quad +f^{\epsilon }(0,x^{\epsilon }_{1},y^{\epsilon },k_{\epsilon }V^{\prime }_{\epsilon })-f^{\epsilon }(0,x^{\epsilon }_{1},y^{\epsilon },V_{\epsilon })\\ &\geq f^{\epsilon }(0,x^{\epsilon }_{1},y^{\epsilon },k_{\epsilon }V^{\prime }_{\epsilon })-f^{\epsilon }(0,x^{\epsilon }_{1},y^{\epsilon },V_{\epsilon })\\ & \geq \left[k_{\epsilon } V^{\prime }_{\epsilon }\left(0,x^{\epsilon }_{1},y^{\epsilon }\right)-V_{\epsilon }\left(0,x^{\epsilon }_{1},y^{\epsilon }\right)\right]\left[g^{\epsilon }(0,x^{\epsilon }_{1},y^{\epsilon },k_{\epsilon } V^{\prime }_{\epsilon })+V_{\epsilon }\left(0,x^{\epsilon }_{1},y^{\epsilon }\right)g^{\epsilon }_{u}\left(0,x^{\epsilon }_{1},y^{\epsilon },\eta _{\epsilon }\right)\right], \end{align*}
where
$\eta _{\epsilon }$
is between
$k_{\epsilon } V^{\prime }_{\epsilon }(0,x^{\epsilon }_{1},y^{\epsilon })$
and
$V_{\epsilon }(0,x^{\epsilon }_{1},y^{\epsilon })$
. Since
$k_{\epsilon } V^{\prime }_{\epsilon }(0,x^{\epsilon }_{1},y^{\epsilon })-V_{\epsilon }(0,x^{\epsilon }_{1},y^{\epsilon })\lt 0$
, we obtain that
From (H2),
$g^{\epsilon }_{u}\leq 0$
. Then
By (H4),
$\{x^{\epsilon }_{1}\}$
must be bounded. From
$k_{\epsilon }V^{\prime }_{\epsilon }(0,x^{\epsilon }_{1},y^{\epsilon })-V(0,x^{\epsilon }_{1},y^{\epsilon })+\epsilon =0$
, we get that
$\{k_{\epsilon }\}$
is bounded and then
$k^{*}\in [1,\infty )$
. Choose a sequence
$\{\epsilon _{n}\}$
in
$\mathbb{R}^{+}$
such that
and
locally uniformly in
$(t,x)\in \mathbb{R}\times \Omega$
,
and
Then
$V^{*}$
and
$V^{\prime *}$
are bounded positive entire solutions of (4.1) with
$c$
and
$f$
are replaced by
$c^{*}$
and
$f^{*}$
. Moreover, we have
$k^{*}V^{\prime *}\geq V^{*}$
and
$k^{*}V^{\prime *}(0,\xi ,\eta )-V^{*}(0,\xi ,\eta )=0$
. Set
$W\,:\!=\,k^{*}V^{\prime *}-V^{*}$
satisfies
$W(0,\xi ,\eta )=0$
. Then,
and
Since
$W$
is nonnegative in
$\mathbb{R}\times \Omega$
, vanishes at
$(\xi ,\eta )$
, the strong maximum principle and the Hopf lemma yield
$W=0$
. Then
which implies that
$k^{*}=1$
and then
$V(t,x_{1},y)\leq V^{\prime }(t,x_{1},y)$
. Similarly, we can also prove
$V(t,x_{1},y)\geq V^{\prime }(t,x_{1},y)$
. Therefore,
$V(t,x_{1},y)= V^{\prime }(t,x_{1},y)$
and
$V(t,x_{1},y)$
is a unique bounded positive solution of (4.1). This completes the proof.
Lemma 4.2.
A function
$f(t,x)$
is almost periodic in
$t$
uniformly with respect with respect to
$x\in E\subset \mathbb{R}^{K}$
if and only if it is uniformly continuous on
$\mathbb{R}\times E$
and for every pair of sequences
$\{s_{n}\}^{\infty }_{n=1}$
,
$\{r_{m}\}^{\infty }_{m=1}$
, there are subsequences
$\{s^{\prime }_{n}\}^{\infty }_{n=1}\subset \{s_{n}\}^{\infty }_{n=1}$
,
$\{r^{\prime }_{n}\}^{\infty }_{n=1}\subset \{r_{n}\}^{\infty }_{n=1}$
such that for each
$(t,x)\in \mathbb{R}\times \mathbb{R}^{K}$
,
Proof. See [Reference Fink10, Theorems 1.17 and 2.10].
Theorem 4.1.
Assume that (H1)–(H4) hold true. If
$\lambda _{\infty }\gt 0$
,
$V(t,x_{1},y)$
is a bounded, positive, unique and almost periodic entire solution of (4.1). For any
$v_{0}\in X^{+}\setminus \{0\}$
, there holds
Proof. From Lemma 4.1, we need to show that
$V(t,x_{1},y)$
is almost periodic in
$t$
for
$(x_{1},y)$
in bounded sets. We write
$V(t,x_{1},y;\,c,f)$
for
$V(t,x_{1},y)$
to indicate the dependence on
$c$
and
$f$
. The Schauder estimates ensure that
$V(t,x_{1},y;\,c,f)$
is uniformly continuous in
$(t,x_{1},y)\in \mathbb{R}\times \Omega$
. It then suffices to show that for each
$x$
in bounded sets,
$V(t,x_{1},y;\,c,f)$
is almost periodic in
$t$
. To this end, let
$\{\alpha _{n}\}$
and
$\{\beta _{n}\}$
be any two sequence of
$\mathbb{R}$
. Without loss of generality, we may assume that
locally uniformly in
$(t,x,u)\in \mathbb{R}\times \Omega \times \mathbb{R}$
and
locally uniformly in
$x\in \Omega$
. Therefore,
locally uniformly in
$(t,x)\in \mathbb{R}\times \Omega$
, where
$V(t,x;\,v^{\prime },c^{\prime },f^{\prime })$
and
$V(t,x;\,v^{\prime \prime },c^{\prime \prime },f^{\prime \prime })$
are bounded positive entire solution of (4.1) with
$c$
and
$f$
replaced by
$c^{\prime },f^{\prime }$
and
$c^{\prime \prime },f^{\prime \prime }$
, respectively. By Lemma 4.1,
for all
$L\gt 0$
.
Let
locally uniformly in
$(t,x)\in \mathbb{R}\times \Omega$
. Since
$\lim _{n\rightarrow \infty }c^{\prime }\cdot \alpha _{n}=c^{\prime \prime }$
, then
$V^{\prime }$
is a bounded positive entire solutions of (1.2) with
$c$
and
$f$
are replaced by
$c^{\prime }$
and
$f^{\prime }$
, and
$\inf _{\mathbb{R}\times [{-}L,L]\times \omega }V^{\prime }\gt 0$
. By the uniqueness of strictly positive solution, see Lemma 4.1,
$V^{\prime }(t,x)=V(t+\alpha _{n},x;\,v^{\prime \prime },c^{\prime \prime },f^{\prime \prime })$
. It then follows from (4.5) that
\begin{align*} &\lim _{n\rightarrow \infty }\lim _{m\rightarrow \infty }V(t+\alpha _{n}+\beta _{m},x;\,c,f)\\ =&\lim _{n\rightarrow \infty }V(t+\alpha _{n},x;\,v^{\prime },c^{\prime },f^{\prime })=V(t+\alpha _{n},x;\,v^{\prime \prime },c^{\prime \prime },f^{\prime \prime })\\ =&\lim _{n\rightarrow \infty }V(t+\alpha _{n}+\beta _{n},x;\,c,f) \end{align*}
holds for all
$t\in \mathbb{R}$
and uniformly for
$x$
in bounded subsets. Hence, by Lemma 4.3,
$V(t,x;\,c,f)$
is almost periodic in
$t$
uniformly for
$x$
in bounded sets.
Next, we show that (4.4) holds true. Let
$v_{0}\in X^{+}\setminus \{0\}$
. Assume that the conclusion fails. Then there exist
$\varepsilon _{0}\gt 0$
and
$\{t_{n}\}\subset (0,\infty )$
satisfying
$\lim _{n\rightarrow \infty }t_{n}=\infty$
and
$\{(x^{n}_{1},y^{n})\}\subset \Omega$
such that
It is no loss of generality to assume that
$y^{n}$
converges to some
$\eta \in \overline {\omega }$
. We may assume without loss of generality that
$c\cdot t_{n}\rightarrow c^{\prime }$
as
$n\rightarrow \infty$
. When
$\lim _{n\rightarrow \infty }|x^{n}_{1}|=\infty$
, by (3.3), we must have
$\lim _{n\rightarrow \infty }V(t_{n},x^{n}_{1},y^{n};\,c,f)=0$
and
$\lim _{n\rightarrow \infty }v(t_{n},x^{n}_{1},y^{n};\,v_{0},c,f)=0$
, which is a contradiction with (4.6). When
$\{x^{n}_{1}\}$
is bounded, set
$x^{n}_{1}\rightarrow x^{\prime }_{1}$
as
$n\rightarrow \infty$
. By standard parabolic estimates, we may assume that
locally uniformly in
$t\in \mathbb{R}$
and
$(x_{1},y)\in \Omega$
, respectively. Then,
$v_{i}(t,x_{1},y;\,c^{\prime },f^{\prime })\,(i=1,2)$
are bounded positive entire solution of (4.1) with
$c$
and
$f$
replaced by
$c^{\prime }$
and
$f^{\prime }$
. Moreover,
$\inf _{\mathbb{R}\times [{-}L,L]\times \omega }v_{i}(\cdot ,\cdot ,\cdot ;\,c^{\prime },f^{\prime })\gt 0$
for all
$L\gt 0$
. The uniqueness in Lemma 4.1 implies that
$v_{1}(t,x_{1},y;\,c^{\prime },f^{\prime })=v_{2}(t,x_{1},y;\,c^{\prime },f^{\prime })$
for all
$t\in \mathbb{R}$
and
$(x_{1},y)\in \Omega$
. By (4.6) again, when
$n\rightarrow \infty$
, there is
$|v_{1}(0,x^{\prime }_{1},\eta ;\,c^{\prime })-v_{2}(0,x^{\prime }_{1},\eta ;\,c^{\prime })|\geq \varepsilon _{0}$
. This is a contradiction. Hence, (4.4) holds true. This completes the proof.
In the following, we will mainly study the effects of fluctuations on the shifting speed and the exponential decay of the forced wave solution
$V(t,x_{1},y)$
of (4.1). Assume the shifting speed function
$c(t)=c+\sigma (t)$
for some fluctuations
$\sigma (t)$
. We would like to see if the fluctuation
$\sigma (t)$
can drive the species to extinction and how large it needs to be in order to make this happen. Consider the following specified model
\begin{equation} \begin{cases} \partial _{t}u=\Delta u+f\left (t,x_{1}-ct-A\int ^{t}_{0}\sigma (s)ds,y,u\right ),\, t\gt 0,x\in \Omega ,\\[3pt] \partial _{\nu }u(t,x)=0,\quad t\gt 0,x\in \partial \Omega , \end{cases} \end{equation}
where
$c\gt 0$
,
$A\geq 0$
and
$\sigma \,:\,\mathbb{R}\rightarrow \mathbb{R}$
is Holder continuous. The linearisation of (4.7) in the moving frame around the extinction state
$0$
is
\begin{equation} \begin{cases} \partial _{t}w=\Delta w+[c+A\sigma (t)]\partial _{1}w+g(t,x,0)w,\quad x\in \Omega ,\\[3pt] \partial _{\nu }w=0,\quad \text{on }\partial \Omega . \end{cases} \end{equation}
Here,
$A\sigma (t)$
is the fluctuation on the shifting speed
$c$
and the parameter
$A$
characterises the amplitude if fluctuations.
Define
$\lambda ^{A}$
the principal Lyapunov exponent of (4.8) and
$\lambda ^{A}_{L}$
restricted on
$({-}L,L)$
. Then,
$\lambda ^{A}_{\infty }=\lim _{L\rightarrow \infty }\lambda ^{A}_{L}$
and
$\lambda ^{A}_{L}\leq \lambda ^{A}$
. Let
$\lambda (\Omega )$
is the generalised Neumann principal eigenvalue of the following problem
\begin{equation*} \begin{cases} -(\Delta \phi +c\partial _{1}\phi )=\lambda (\Omega )\phi ,\quad x\in \mathbb{R}\times \omega ,\\[3pt] \partial _{\nu }\phi =0,\quad x\in \mathbb{R}\times \partial \omega \end{cases} \end{equation*}
and we refer to see Remark 2.1(2) for the definition of generalised Neumann principal eigenvalue.
Theorem 4.2.
(1) If
$A=0$
, then there is a critical speed
$c^{*}\gt 0$
such that
$\lambda ^{0}_{\infty }(c)\gt 0$
for
$c\in [0,c^{*})$
and
$\lambda ^{0}_{\infty }(c)\lt 0$
for
$c\gt c^{*}$
.
(2) If
$A\neq 0$
, assume that
$\sigma (t)$
and
$f(t,x,u)$
are periodic in
$t$
. Then
Proof. (1) When
$A=0$
, by the Liouville transformation
$v(t,x_{1},y)=u(t,x_{1},y)e^{\frac {c}{2}x_{1}}$
, problem (4.1) reduces to
\begin{equation} \begin{cases} \partial _{t}v=\Delta v+f\left (t,x_{1},y,v(t,x_{1},y)e^{-\frac {c}{2}x_{1}}\right )e^{\frac {c}{2}x_{1}}-\frac {c^{2}}{4}v,\, t\gt 0,x\in \Omega ,\\[3pt] \partial _{\nu }v(t,x)=0,\quad t\gt 0,x\in \partial \Omega . \end{cases} \end{equation}
The associated linearised problem associated with (4.9) is
\begin{equation} \begin{cases} \partial _{t}w=\Delta w+g(t,x,0)w-\frac {c^{2}}{4}w,\, t\gt 0,x\in \Omega ,\\[3pt] \partial _{\nu }w(t,x)=0,\quad t\gt 0,x\in \partial \Omega . \end{cases} \end{equation}
Let
$\lambda ^{0}_{\infty }(0)$
is the principal Lyapunov exponent of
\begin{equation} \begin{cases} \partial _{t}w=\Delta w+g(t,x,0)w,\, t\gt 0,x\in \Omega ,\\[3pt] \partial _{\nu }w(t,x)=0,\quad t\gt 0,x\in \partial \Omega . \end{cases} \end{equation}
Then
Define
\begin{equation*} c^{*}= \begin{cases} 2\sqrt {\lambda ^{0}_{\infty }}\quad \text{if }\lambda ^{0}_{\infty }\gt 0,\\[4pt] 0\quad \text{otherwise }. \end{cases} \end{equation*}
It then follows that
$\lambda ^{0}_{\infty }(c)\gt 0$
for
$c\in [0,c^{*})$
and
$\lambda ^{0}_{\infty }(c)\lt 0$
for
$c\gt c^{*}$
.
(2) Fix
$A\gt 0$
. Note that
$\sigma (t)$
and
$f(t,x,u)$
are periodic in
$t$
. Then
$\lambda ^{A}_{L}$
is the principal eigenvalue of (4.10) on
$({-}L,L)\times \Omega$
and
$t\in \mathbb{R}$
. Then
where
$\Lambda _{L}$
is the principal eigenvalue of
\begin{equation*} \begin{cases} \partial _{t}w=\Delta w+[c+A\sigma (t)]\partial _{1}w,\quad x\in ({-}L,L)\times \omega ,\\ \partial _{\nu }w=0,\quad x\in ({-}L,L)\times \partial \omega ,\\ w(t,L,y)=w(t,-L,y)=0,\quad y\in \omega \end{cases} \end{equation*}
and
$\phi _{L}(t,x_{1},y)$
is the periodic eigenfunction corresponding to
$\Lambda _{L}$
. Set
Then
$\Phi _{L}(t,z,x)$
is a periodic function defined on periodic domain
and
$\Phi _{L}(t,z,x)$
also satisfies
\begin{equation*} \begin{cases} \partial _{t}\Phi _{L}=\Delta \Phi _{L}+c\partial _{1}\Phi _{L}-\Lambda _{L}\Phi _{L},\quad x\in ({-}L,L)\times \omega ,\\ \partial _{\nu }\Phi _{L}=0,\quad x\in ({-}L,L)\times \partial \omega ,\\ \Phi _{L}(t,L,y)=\Phi _{L}(t,-L,y)=0,\quad y\in \omega . \end{cases} \end{equation*}
Let
$\widetilde {L}\gt 0$
large such that
$\mathbb{R}\times [{-}\widetilde {L}-1,\widetilde {L}+1]\times \omega \subset D_{T}$
. Moreover,
$\widetilde {L}\rightarrow \infty$
as
$L\rightarrow \infty$
. Then on
$\mathbb{R}\times [{-}\widetilde {L}-1,\widetilde {L}+1]\times \omega$
, we find that
$\Phi _{L}$
is periodic in
$t$
and
$\inf \Phi _{L}\gt 0$
and
\begin{equation*} \begin{cases} \partial _{t}\Phi _{L}=\Delta \Phi _{L}+c\partial _{1}\Phi _{L}-\Lambda _{L}\Phi _{L},\quad [{-}\widetilde {L}-1,\widetilde {L}+1]\times \omega ,\\ \partial _{\nu }\Phi _{L}=0,\quad x\in ({-}L,L)\times \partial \omega . \end{cases} \end{equation*}
Set
$\widetilde {\Lambda }_{L}$
is the principal eigenvalue of
\begin{equation*} \begin{cases} \Delta \phi +c\partial _{1}\phi =\widetilde {\Lambda }\phi ,\quad x\in ({-}\widetilde {L},\widetilde {L})\times \omega ,\\ \partial _{\nu }\phi _{L}=0,\quad x\in ({-}\widetilde {L},\widetilde {L})\times \partial \omega ,\\ \phi (\widetilde {L},y)=\phi ({-}\widetilde {L},y)=0,\quad y\in \omega . \end{cases} \end{equation*}
Then the comparison principal implies that
$\Lambda _{L}\geq \widetilde {\Lambda }_{L}$
. By (4.12), we get
Following from [Reference Berestycki and Rossi3, Proposition 1], we have that
$\lim _{L\rightarrow \infty }\widetilde {\lambda }_{L}$
exists and
$\lim _{L\rightarrow \infty }\widetilde {\lambda }_{L}=-\lambda (\Omega )$
. Thus, we obtain that
$\lambda ^{A}_{\infty }\geq -\lambda (\Omega )+\inf _{t\in \mathbb{R},x\in \Omega }g(t,x,0).$
This completes the proof.
In order to use Theorem 4.2(1), let
$V(t,x_{1},y)$
be the forced wave solution of (4.1) with
$c(t)\equiv c$
, that is
$A=0$
.
Lemma 4.3.
Let
$v(t,x_{1},y)\in W^{1,2}_{N+1,loc}(\mathbb{R}\times \Omega _{L})$
with
$\partial _{\nu }v\leq 0$
on
$\mathbb{R}\times \partial \Omega$
for every
$L\gt 0$
and let
$C,\gamma \gt 0$
be such that
uniformly in
$t\in \mathbb{R}$
and
$y\in \omega$
. Then there exists a constant
$\kappa \gt 0$
such that
Proof. The proof is similar to that in [Reference Berestycki and Rossi3]. For sake of completeness, we include the proof here. By the hypotheses on
$v$
, there exist
$\varepsilon \gt 0$
and
$L\gt 0$
such that
for a.e.
$t\in \mathbb{R}$
and
$(x_{1},y)\in \Omega \setminus \overline {\Omega }_{L}$
. We list the following functions as that in [Reference Berestycki and Rossi3, Lemma 3.2]. For
$a\gt 0$
, let
$\theta _{a}\,:\,[L,L+a]\rightarrow \mathbb{R}$
be the solution to
\begin{equation} \begin{cases} \theta ^{\prime \prime }=(\gamma +\varepsilon )\theta ,\quad \text{in }(L,L+a),\\ \theta (L)=ke^{\sqrt {\gamma }L},\\ \theta (L+a)=ke^{\sqrt {\gamma }(L+a)}. \end{cases} \end{equation}
Hence,
$\theta _{a}(\rho )=A_{a}e^{-\sqrt {\gamma +\varepsilon }\rho }+B_{a}e^{\sqrt {\gamma +\varepsilon }\rho }$
,
$\rho \in [L,L+a]$
with
and
Fix
$\tau \in \mathbb{R}$
and define
where
By the similar computation in [Reference Berestycki and Rossi2, Lemma 3.2],
\begin{equation} \begin{cases} \partial _{t}\theta _{a}\geq \Delta \theta _{a}-(\gamma +2\varepsilon )\theta _{a},\quad t\in ({-}\infty ,\tau ), (x_{1},y)\in (L,L+a)\times \omega ,\\ \partial _{\nu }\theta _{a}=0,\quad t\in ({-}\infty ,\tau ), (x_{1},y)\in (L,L+a)\times \partial \omega . \end{cases} \end{equation}
Note that
$v$
is a subsolution of (4.15) and
$v\leq \theta _{a}$
on
$\{\pm L,\pm (L+a)\}\times \omega$
,
$t\in ({-}\infty ,\tau )$
. Then the comparison principle implies that
For
$|x_{1}|\gt L$
and
$y\in \omega$
, we have
Due to the arbitrarily of
$\tau$
, we obtain that (4.13) holds true. This completes the proof.
Theorem 4.3.
Assume that (H1)-(H4) hold true. Let
$V(t,x_{1},y)$
be a solution of (4.1) with
$c(t)\equiv c$
. Then there exist two constants
$h$
and
$\beta \gt 0$
such that
Proof. The function
$v(t,x_{1},y)=V(t,x_{1},y)e^{\frac {c}{2}x_{1}}$
is a solution of (4.9). Hence,
for
$t\in \mathbb{R}$
and
$(x_{1},y)\in \Omega$
, where
$\zeta (t,x_{1},y)\leq g(t,x_{1},y,0)$
.
Set
$\zeta =-\lim _{r\rightarrow \infty }\sup _{|x_{1}|\gt r,t\in \mathbb{R},y\in \omega }g(t,x_{1},y,0)$
and (H4) implies
$\zeta \gt 0$
. Then for
$\gamma \in (\frac {c^{2}}{4},\frac {c^{2}}{4}+\zeta )$
,
On the other hand,
$v(t,x_{1},y)e^{-\sqrt {\gamma }|x_{1}|}\leq v(t,x_{1},y)e^{-\frac {c}{2}|x_{1}|}$
, which is bounded. Then by Lemma 4.3, there exist two constants
$h$
and
$\beta \gt 0$
such that
This completes the proof.
Acknowledgements
The authors would like to thank the referee for valuable comments and suggestions that improved the presentation of this manuscript.
Funding support
Xiongxiong Bao was partially supported by NSF of China (12271058) and Shaanxi Fundamental Science Research Project for Mathematics and Physics (23JSY040). Dun Zhou was partially supported by NSF of China (12331006).
Competing interest
The authors declare no competing interests.
