This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Traveling wave solutions and spreading speeds for a scalar age-structured equation with nonlocal diffusionthanks: Research was partially supported by the National Natural Science Foundation of China (No. 12301259 and No. 12371169).

Arnaud Ducrota and Hao Kangb
aNormandie Univ, UNIHAVRE, LMAH, FR-CNRS-3335, ISCN, 76600 Le Havre, France
Emails: [email protected]
bCenter for Applied Mathematics, Tianjin University, Tianjin, China

Email: [email protected]
Abstract

In this paper, we study the existence of traveling wave solutions and the spreading speed for the solutions of an age-structured epidemic model with nonlocal diffusion. Our proofs make use of the comparison principles both to construct suitable sub/super-solutions and to prove the regularity of traveling waves solutions.

Key words: Age structure; Nonlocal diffusion; Traveling wave solutions; Spreading speeds.

AMS subject classifications:  35K55, 35C07, 45G10, 92D30

Dedicated to Professor Shigui Ruan for His 60th Birthday.

1 Introduction and main results

In this paper, we are concerned with the study of an age dependent epidemic model with nonlocal diffusion in space, motivated by [6], where the first author studied as similar problem with random diffusion in space. Our aim is to study a classical SI model from the point of view of the spatial spread of an epidemics. Here a(0,a+)a\in(0,a^{+}) denotes the physiological age and a+(0,)a^{+}\in(0,\infty) is the maximum age of an individual, t>0t>0 denotes the time and xx\in\mathbb{R} is the spatial position of an individual. The population can be split into two sub-populations, the susceptible and the infective. We denote by S(t,a,x)S(t,a,x), respectively I(t,a,x)I(t,a,x), the age distribution of the susceptible individuals, respectively the infective individuals, at time tt and spatial location xx\in\mathbb{R}. With these notations, the model that we consider reads as follows

{(It+Ia)(t,a,x)=(JII)(t,a,x)+S(t,a,x)0a+K(a,a)I(t,a,x)𝑑aμ(a)I(t,a,x),(St+Sa)(t,a,x)=d(JSS)(t,a,x)S(t,a,x)0a+K(a,a)I(t,a,x)𝑑aμ(a)S(t,a,x),I(t,0,x)=p0a+β2(a)I(t,a,x)𝑑a,S(t,0,x)=0a+β1(a)S(t,a,x)𝑑a+(1p)0a+β2(a)I(t,a,x)𝑑a,I(0,a,x)=I0(a,x),S(0,a,x)=S0(a,x),\begin{cases}\left(\frac{\partial I}{\partial t}+\frac{\partial I}{\partial a}\right)(t,a,x)=(J\ast I-I)(t,a,x)+S(t,a,x)\int_{0}^{a^{+}}K(a,a^{\prime})I(t,a^{\prime},x)da^{\prime}-\mu(a)I(t,a,x),\\ \left(\frac{\partial S}{\partial t}+\frac{\partial S}{\partial a}\right)(t,a,x)=d\left(J\ast S-S\right)(t,a,x)-S(t,a,x)\int_{0}^{a^{+}}K(a,a^{\prime})I(t,a^{\prime},x)da^{\prime}-\mu(a)S(t,a,x),\\ I(t,0,x)=p\int_{0}^{a^{+}}\beta_{2}(a)I(t,a,x)da,\\ S(t,0,x)=\int_{0}^{a^{+}}\beta_{1}(a)S(t,a,x)da+(1-p)\int_{0}^{a^{+}}\beta_{2}(a)I(t,a,x)da,\\ I(0,a,x)=I_{0}(a,x),\;S(0,a,x)=S_{0}(a,x),\end{cases} (1.1)

wherein we have set

(Juu)(x):=J(xy)u(y)dyu(x),uCb(),\left(J\ast u-u\right)(x)\mathrel{\mathop{\mathchar 58\relax}}=\int_{\mathbb{R}}J(x-y)u(y)dy-u(x),\quad\forall u\in C_{b}(\mathbb{R}),

for some continuous convolution kernel J:J\mathrel{\mathop{\mathchar 58\relax}}\mathbb{R}\to\mathbb{R}, whose specific properties will be presented in Assumption 1.1 below, while Cb()C_{b}(\mathbb{R}) denotes the space of bounded and continuous functions. Here the function K(a,a)K(a,a^{\prime}) denotes the rate of the disease transmission from infective individuals of age aa^{\prime} to susceptible individuals of age aa. In addition, the function μ\mu denotes the age-specific death rate for both the infective and susceptible individuals, β1\beta_{1} and β2\beta_{2} denote the age-specific birth rate for the susceptible and infective ones, respectively. Moreover, we assume that these birth rates are identical, that is β1β2=β\beta_{1}\equiv\beta_{2}=\beta. The constant p[0,1]p\in[0,1] denotes the proportion of the vertical transmission, that is the proportion of the infective newborns inherited from their infective parents. Further, we assume that SS and II have the same diffusion coefficient, that is d=1d=1. This assumption combined with no additional death rate due to the disease allow us to reduce the system to a single scalar equation. We would like to mention that for the existence of traveling wave of age-structured SISI system with random diffusion have been obtained by Ducrot and Magal [11, 12] with/without external supply and Ducrot et al. [13] in a multigroup framework, respectively.

In this work, we consider that the total population has a demographic equilibrium, meaning that birth and death equilibrate the population. This prevents the population from going to extinction and from exploding as time increases. The total population stabilizes to a steady state as tt\to\infty. Mathematically speaking, this assumption can be written as the following condition on the demographic functions β\beta and μ\mu:

0a+β(a)exp(0aμ(a)𝑑a)𝑑a=1.\int_{0}^{a^{+}}\beta(a)\exp\left(-\int_{0}^{a}\mu(a^{\prime})da^{\prime}\right)da=1. (1.2)

This condition may imply that multiple steady states exist for the total population I+SI+S when d=1d=1 (see Webb [27] for some discussions on this condition). Next, we are looking for heteroclinic solutions of system (1.1) with the following behavior for x±x\to\pm\infty:

I(t,a,)=exp(0aμ(a)𝑑a),I(t,a,+)=0,S(t,a,)=0,S(t,a,+)=exp(0aμ(a)𝑑a),\begin{array}[]{ll}I(t,a,-\infty)=\exp\left(-\int_{0}^{a}\mu(a^{\prime})da^{\prime}\right),\;I(t,a,+\infty)=0,\\ S(t,a,-\infty)=0,\;S(t,a,+\infty)=\exp\left(-\int_{0}^{a}\mu(a^{\prime})da^{\prime}\right),\end{array} (1.3)

These conditions mean that at x=+x=+\infty, the population is only composed of the susceptible, whereas at x=x=-\infty the population is only composed of infected individuals.
Now when d=1d=1, adding-up the first two equations in (1.1), the nonlinear terms cancel and one obtains using (1.3), that

(I+S)(t,a,x)=exp(0aμ(a)da):=π(a), provided (I0+S0)(a,x)=π(a).(I+S)(t,a,x)=\exp\left(-\int_{0}^{a}\mu(a^{\prime})da^{\prime}\right)\mathrel{\mathop{\mathchar 58\relax}}=\pi(a),\text{ provided }(I_{0}+S_{0})(a,x)=\pi(a).

As a consequence, system (1.1)-(1.3) reduces to the scalar equation for the unknown function II

{(It+Ia)(t,a,x)=(JII)(t,a,x)+0a+K(a,a)I(t,a,x)𝑑a(π(a)I(t,a,x))μ(a)I(t,a,x),I(t,a,)=π(a),I(t,a,)=0,I(t,0,x)=p0a+β(a)I(t,a,x)𝑑a.\begin{cases}\left(\frac{\partial I}{\partial t}+\frac{\partial I}{\partial a}\right)(t,a,x)=(J\ast I-I)(t,a,x)+\int_{0}^{a^{+}}K(a,a^{\prime})I(t,a^{\prime},x)da^{\prime}\,(\pi(a)-I(t,a,x))-\mu(a)I(t,a,x),\\ I(t,a,-\infty)=\pi(a),\;I(t,a,\infty)=0,\\ I(t,0,x)=p\int_{0}^{a^{+}}\beta(a)I(t,a,x)da.\end{cases} (1.4)

We rewrite this problem with the new unknown function u(t,a,x)u(t,a,x) defined by

u(t,a,x)=I(t,a,x)π(a)u(t,a,x)=\frac{I(t,a,x)}{\pi(a)}

and we obtain the following equation for the function uu:

{(ut+ua)(t,a,x)=(Juu)(t,a,x)+0a+K(a,a)π(a)u(t,a,x)𝑑a(1u(t,a,x)),u(t,a,)=1,u(t,a,)=0,u(t,0,x)=p0a+β(a)π(a)u(t,a,x)𝑑a,\begin{cases}\left(\frac{\partial u}{\partial t}+\frac{\partial u}{\partial a}\right)(t,a,x)=(J\ast u-u)(t,a,x)+\int_{0}^{a^{+}}K(a,a^{\prime})\pi(a^{\prime})u(t,a^{\prime},x)da^{\prime}\,(1-u(t,a,x)),\\ u(t,a,-\infty)=1,\;u(t,a,\infty)=0,\\ u(t,0,x)=p\int_{0}^{a^{+}}\beta(a)\pi(a)u(t,a,x)da,\end{cases} (1.5)

Now we set

γ(a):=β(a)π(a),\gamma(a)\mathrel{\mathop{\mathchar 58\relax}}=\beta(a)\pi(a),

and recalling condition (1.2), we assume the following properties.

Assumption 1.1

We assume that the following properties hold true:

  • (i)

    the function γ\gamma is continuous and nonnegative in [0,a+][0,a^{+}] and satisfies 0a+γ(a)𝑑a=1\int_{0}^{a^{+}}\gamma(a)da=1;

  • (ii)

    the function π\pi is continuous, positive in [0,a+][0,a^{+}] and the function γπ\gamma\pi does not identically vanish on (0,a+)(0,a^{+});

  • (iii)

    the kernel JJ is continuous and nonnegative with J(x)𝑑x=1,J(0)>0\int_{\mathbb{R}}J(x)dx=1,J(0)>0 and J(x)=J(x),xJ(-x)=J(x),\forall x\in\mathbb{R}. Moreover, J(x)elx𝑑x<\int_{\mathbb{R}}J(x)e^{lx}dx<\infty for any l>0l>0;

  • (iv)

    the transmission rate K(,)K(\cdot,\,\cdot) is continuous and positive in [0,a+]×[0,a+][0,a^{+}]\times[0,a^{+}].

We now consider equation (1.5) and we are concerned with traveling wave solutions, that is particular solutions of the form U(a,ξ)=u(a,xct)U(a,\xi)=u(a,x-ct) with ξ=xct\xi=x-ct. Here cc is an unknown real number which should be found together with the unknown function UU. Before proceeding, we set the rate of vertical transmission to be one, i.e. p=1p=1. This setting is kind of technical mathematically, since p=1p=1 is used to guarantee that (1) the limiting equation (2.4) has only two solutions 0 and 11, see Section 2.2; (2) the hair trigger effect holds, see Lemma 4.1. Next, using the moving frame, that is the variable ξ=xct\xi=x-ct, the equation for the profile function UU becomes for ξ\xi\in\mathbb{R} and a[0,a+]a\in[0,a^{+}]:

{U(a,ξ)a=(JξUU)(a,ξ)+cU(a,ξ)ξ+0a+K(a,a)π(a)U(a,ξ)𝑑a(1U(a,ξ)),U(a,)=1,U(a,)=0, uniformly in [0,a+],U(0,ξ)=0a+γ(a)U(a,ξ)𝑑a.\begin{cases}\frac{\partial U(a,\xi)}{\partial a}=\left(J\ast_{\xi}U-U\right)(a,\xi)+c\frac{\partial U(a,\xi)}{\partial\xi}+\int_{0}^{a^{+}}K(a,a^{\prime})\pi(a^{\prime})U(a^{\prime},\xi)da^{\prime}\,(1-U(a,\xi)),\\ U(a,-\infty)=1,\;U(a,\infty)=0,\text{ uniformly in $[0,a^{+}]$},\\ U(0,\xi)=\int_{0}^{a^{+}}\gamma(a)U(a,\xi)da.\end{cases} (1.6)

When the age specific demographic and epidemic functions are ignored, this equation reduces to a monostable equation with nonlocal diffusion, which has been well studied by many researchers, see for example Coville et al. [4, 5, 2, 3], Fang and Zhao [14], Li et al. [17, 24] and Shen et al. [23, 22] and the references therein. In addition, we mention that the global dynamics of (1.4) with spatially dependent coefficients on the bounded domain can be studied using the sign of the spectral bound of a linearized operator at zero, the interested readers can refer to Ducrot et al. [9, 10] and Kang and Ruan [16] for more details.

Next, we solve (1.6) formally along characteristic lines. Define the characteristic line h(a,ξ)h(a,\xi) as the solution of the following equation,

{h(a,ξ)a=c,a[0,a+],ξ,h(0,ξ)=ξ,\begin{cases}\frac{\partial h(a,\xi)}{\partial a}=-c,&a\in[0,a^{+}],\;\xi\in\mathbb{R},\\ h(0,\xi)=\xi,\end{cases} (1.7)

that reads as h(a,ξ)=ξcah(a,\xi)=\xi-ca, which implies that ξh(a,ξ)=1\partial_{\xi}h(a,\xi)=1. Next, solving (1.6) along the characteristic line h(a,ξ)h(a,\xi), one obtains that the function

w(a,ξ)=U(a,h(a,ξ))=U(a,ξca)w(a,\xi)=U(a,h(a,\xi))=U(a,\xi-ca) (1.8)

satisfies the following equation,

{w(a,ξ)a=(Jξww)(a,ξ)+0a+K(a,a)π(a)w(a,ξ)𝑑a(1w(a,ξ)),w(a,)=1,w(a,)=0, uniformly in [0,a+],w(0,ξ)=0a+γ(a)w(a,ξ+ca)𝑑a.\begin{cases}\frac{\partial w(a,\xi)}{\partial a}=\left(J\ast_{\xi}w-w\right)(a,\xi)+\int_{0}^{a^{+}}K(a,a^{\prime})\pi(a^{\prime})w(a^{\prime},\xi)da^{\prime}\,(1-w(a,\xi)),\\ w(a,-\infty)=1,\;w(a,\infty)=0,\text{ uniformly in $[0,a^{+}]$},\\ w(0,\xi)=\int_{0}^{a^{+}}\gamma(a)w(a,\xi+ca)da.\end{cases} (1.9)

The above reformulation using the characteristic lines allows us to take care of the term Uξ\frac{\partial U}{\partial\xi} appearing in (1.6). We now give the definition of a traveling wave solution of (1.5) as follows.

Definition 1.2

A function w=w(a,ξ)w=w(a,\xi) is said to be a traveling wave solution of (1.5) if wC([0,a+]×)w\in C([0,a^{+}]\times\mathbb{R}) is a bounded solution of (1.9) such that for any a(0,a+)a\in(0,a^{+}) the map ξw(a,ξ)\xi\mapsto w(a,\xi) is globally Lipschitz continuous and for any ξ\xi\in\mathbb{R}, aw(a,ξ)W1,1(0,a+)a\mapsto w(a,\xi)\in W^{1,1}(0,a^{+}).

The first aim of this paper is to prove the existence of traveling wave solutions for (1.5). More precisely, we will prove the following result.

Theorem 1.3

Let Assumption 1.1 be satisfied. Then there exists c>0c^{*}>0 such that for any ccc\geq c^{*}, system (1.5) has a traveling wave solution. Moreover, any traveling wave solution w=w(a,ξ)w=w(a,\xi) is nonincreasing with respect to ξ\xi\in\mathbb{R}.

This theorem provides an existence result for equation (1.4). Indeed we have the following theorem

Theorem 1.4

Let Assumption 1.1 be satisfied and assume furthermore that μ:[0,a+)+\mu\mathrel{\mathop{\mathchar 58\relax}}[0,a^{+})\to\mathbb{R}_{+} and β:[0,a+)+\beta\mathrel{\mathop{\mathchar 58\relax}}[0,a^{+})\to\mathbb{R}_{+} are continuous functions satisfying

0a+μ(a)da=(π(a+)=0),0a+β(a)exp(0aμ(a)da)da=1.\int_{0}^{a^{+}}\mu(a)da=\infty\;\big{(}\Leftrightarrow\pi(a^{+})=0\big{)},\quad\int_{0}^{a^{+}}\beta(a)\exp\left(-\int_{0}^{a}\mu(a^{\prime})da^{\prime}\right)da=1.

Then there exists c>0c^{*}>0 such that equation (1.4) has a traveling wave solution for any wave speed ccc\geq c^{*} and there is no traveling wave solution of equation (1.4) if c<cc<c^{*}. Moreover, the obtained traveling wave for ccc\geq c^{*} vanishes at the maximum age, that is at a=a+a=a^{+}.

The proofs of these results are based on the comparison principle to construct suitable super- and sub-solutions for equation (1.9), motivated by Ducrot [6]. An important difference between this work and [6] (dealing with with random diffusion) is that the solution map for (1.9) has no compactness property, thus the Schauder fixed point theorem may not directly be applied. However, (1.9) preserves the monotonicity and thus a monotone iteration method still works and allows us to overcome the lack of smoothing effect for nonlocal diffusion equation. In addition, we establish the required regularity of traveling wave solutions with respect to the variable ξ\xi in Definition 1.2 using suitable comparison arguments. Thus this regularity will make the term Uξ\frac{\partial U}{\partial\xi}, appearing in (1.6), to be well-defined at least almost everywhere.

Our second aim in this work is to study the spreading speeds of the solutions of the Cauchy problem, that is of the following initial value problem associated to (1.5)

{(ut+ua)(t,a,x)=(Juu)(t,a,x)+0a+K(a,a)π(a)u(t,a,x)𝑑a(1u(t,a,x)),u(t,0,x)=0a+γ(a)u(t,a,x)𝑑a,u(0,a,x)=u0(a,x).\begin{cases}\left(\frac{\partial u}{\partial t}+\frac{\partial u}{\partial a}\right)(t,a,x)=\left(J\ast u-u\right)(t,a,x)+\int_{0}^{a^{+}}K(a,a^{\prime})\pi(a^{\prime})u(t,a^{\prime},x)da^{\prime}\,(1-u(t,a,x)),\\ u(t,0,x)=\int_{0}^{a^{+}}\gamma(a)u(t,a,x)da,\\ u(0,a,x)=u_{0}(a,x).\end{cases} (1.10)

Then our result reads as follows.

Theorem 1.5

Let Assumption 1.1 be satisfied and assume that the initial data u0C([0,a+]×)u_{0}\in C([0,a^{+}]\times\mathbb{R}) with u01u_{0}\leq 1 is compactly supported, that is, supp(u0)[0,a+]×[R,R]{\rm supp}(u_{0})\subset[0,a^{+}]\times[-R,R] for some R>0R>0. Recalling that cc^{*} is defined in Theorem 1.4, the solution u=u(t,a,x)u=u(t,a,x) of problem (1.10) enjoys the following properties

  • (i)

    For all c>cc>c^{*}, it holds that

    limtsup|x|ct,0<a<a+u(t,a,x)=0.\lim\limits_{t\to\infty}\sup_{|x|\geq ct,0<a<a^{+}}u(t,a,x)=0.
  • (ii)

    For all 0c<c0\leq c<c^{*}, it holds that

    limtinf|x|ct,0<a<a+u(t,a,x)=1.\lim\limits_{t\to\infty}\inf_{|x|\leq ct,0<a<a^{+}}u(t,a,x)=1.
Remark 1.6

The nonexistence of traveling wave solution of (1.1) for c<cc<c^{*} is a direct consequence of the above spreading property (Theorem 1.5).

Since the functions π\pi and γ\gamma do not depend on the spatial variable xx, the above spreading speed results are established by comparing the solutions of (1.10) with whose of the nonlocal diffusion problem with constant demographic functions. More precisely, for outer spreading, namely c>cc>c^{*}, we can use the classical super-solution based on the traveling wave profile, while for inner spreading, namely c<cc<c^{*}, we construct sub-solutions with suitable wave speed using a Fisher-KPP equation with nonlocal diffusion.

The paper is organized as follows. In Section 2, we discuss the well-posedness of (1.10) and construct suitable super-/sub-solutions for equation (1.9). In Section 3, we prove the existence of traveling wave solutions via monotone iteration method and derive regularity estimates for traveling wave solutions. In Section 4, we study the spreading speed for some solutions of (1.10) and complete the proof of Theorem 1.5.

2 Preliminaries

2.1 Well-posedness of (1.10)

In this section, we first study the well-posedness of (1.10). To this aim, we introduce some functional framework and we define

X=Cb(),𝒳=X×L1((0,a+),X) and 𝒳0={0X}×L1((0,a+),X),X=C_{b}(\mathbb{R}),\;\;\mathcal{X}=X\times L^{1}((0,a^{+}),X)\text{ and }\mathcal{X}_{0}=\{0_{X}\}\times L^{1}((0,a^{+}),X),

wherein 0X0_{X} denotes the zero element in XX. These spaces become Bamach spaces when they are endowed with the usual product norms and we also introduce their positive cones as follows

𝒳+\displaystyle\mathcal{X}^{+} =\displaystyle= X+×L+1((0,a+),X)\displaystyle X_{+}\times L^{1}_{+}((0,a^{+}),X)
=\displaystyle= X+×{uL1((0,a+),X):u(a,)X+,a.e. in (0,a+)},\displaystyle X_{+}\times\{u\in L^{1}((0,a^{+}),X)\mathrel{\mathop{\mathchar 58\relax}}u(a,\cdot)\in X_{+},\;\text{a.e. in }(0,a^{+})\},
𝒳0+\displaystyle\mathcal{X}^{+}_{0} =\displaystyle= 𝒳+𝒳0 with X+:={uCb():u0}.\displaystyle\mathcal{X}^{+}\cap\mathcal{X}_{0}\text{ with }X_{+}\mathrel{\mathop{\mathchar 58\relax}}=\{u\in C_{b}(\mathbb{R})\mathrel{\mathop{\mathchar 58\relax}}u\geq 0\}.

We also define the linear positive and bounded convolution operator T(X)T\in\mathcal{L}(X) by

[Tφ]()=J(y)φ(y)dy,φX.[T\varphi](\cdot)=\int_{\mathbb{R}}J(\cdot-y)\varphi(y)dy,\;\forall\varphi\in X. (2.1)

Note that due to Assumption 1.1-(iii) one has

T(X)J(y)dy:=1.\|T\|_{\mathcal{L}(X)}\leq\int_{\mathbb{R}}J(y)dy\mathrel{\mathop{\mathchar 58\relax}}=1. (2.2)

Next define the operator :dom()𝒳𝒳\mathcal{B}\mathrel{\mathop{\mathchar 58\relax}}dom(\mathcal{B})\subset\mathcal{X}\to\mathcal{X} by

dom()={0X}×W1,1((0,a+),X) and (0ψ)=(ψ(0)ψa+[TI]ψ),(0ψ)dom().dom(\mathcal{B})=\{0_{X}\}\times W^{1,1}((0,a^{+}),X)\text{ and }\mathcal{B}\begin{pmatrix}0\\ \psi\end{pmatrix}=\begin{pmatrix}-\psi(0)\\ -\frac{\partial\psi}{\partial a}+[T-I]\psi\end{pmatrix},\;\;\forall\begin{pmatrix}0\\ \psi\end{pmatrix}\in dom(\mathcal{B}).

Note that \mathcal{B} is a closed Hille-Yosida operator. Moreover, define the nonlinear operator 𝒞\mathcal{C} from 𝒳0\mathcal{X}_{0} to 𝒳\mathcal{X} as follows:

𝒞(0ψ)=(0a+γ(a)ψ(a)𝑑a0a+K(,a)π(a)ψ(a)𝑑a(1ψ)),(0ψ)𝒳0.\mathcal{C}\begin{pmatrix}0\\ \psi\end{pmatrix}=\begin{pmatrix}\int_{0}^{a^{+}}\gamma(a)\psi(a)da\\ \int_{0}^{a^{+}}K(\cdot,a^{\prime})\pi(a^{\prime})\psi(a^{\prime})da^{\prime}\,(1-\psi)\end{pmatrix},\;\forall\begin{pmatrix}0\\ \psi\end{pmatrix}\in\mathcal{X}_{0}.

Then by identifying U(t)=(0u(t))U(t)=\begin{pmatrix}0\\ u(t)\end{pmatrix}, problem (1.10) rewrites as the following abstract Cauchy problem:

{dUdt=U+𝒞U,U(0)=U0, with U0=(0u0)𝒳0.\begin{cases}\frac{dU}{dt}=\mathcal{B}U+\mathcal{C}U,\\ U(0)=U_{0},\end{cases}\text{ with }U_{0}=\begin{pmatrix}0\\ u_{0}\end{pmatrix}\in\mathcal{X}_{0}. (2.3)

Based on the Lipshcitz property of 𝒞\mathcal{C}, by Thieme [25, 26] or Magal and Ruan [18], the existence and uniqueness of a mild solution for (2.3) is guaranteed. Next, recalling the definition of \mathcal{B}, one may observe that \mathcal{B} is resolvent positive. Moreover, there exists some constant L>0L>0 such that the operator 𝒞+LIdX\mathcal{C}+LId_{X} is monotone on the subset 𝒮0𝒳0\mathcal{S}_{0}\subset\mathcal{X}_{0} given by

𝒮0:={(0φ)𝒳0,φ1},\mathcal{S}_{0}\mathrel{\mathop{\mathchar 58\relax}}=\left\{\begin{pmatrix}0\\ \varphi\end{pmatrix}\in\mathcal{X}_{0},\;\;\varphi\leq 1\right\},

which means that for all (U,V)𝒮0(U,V)\in\mathcal{S}_{0} one has

UV𝒞U+LU𝒞V+LV,U\leq V\Rightarrow\mathcal{C}U+LU\leq\mathcal{C}V+LV,

where the partial order \leq in 𝒳0\mathcal{X}_{0} and 𝒳\mathcal{X} is induced by the positive cones 𝒳0+\mathcal{X}_{0}^{+} and 𝒳+\mathcal{X}^{+}, respectively. To see this let us observe that for any k>0k>0, if L>kL>k then the function (v,u)v(1u)+Lu(v,u)\mapsto v(1-u)+Lu is increasing with respect to both variables (v,u)(,k]×(,1](v,u)\in(-\infty,k]\times(-\infty,1].
Due to the above monotonicity property, one also obtains that 𝒞U+LU𝒳+\mathcal{C}U+LU\in\mathcal{X}^{+} for all U𝒳0+𝒮0U\in\mathcal{X}_{0}^{+}\cap\mathcal{S}_{0}. Hence the result by Magal et al. [19, Theorem 4.5] applies and ensures that Problem (2.3) is well posed in 𝒳0+𝒮0\mathcal{X}_{0}^{+}\cap\mathcal{S}_{0}, which is forward invariant and the comparison principle holds for (2.3) in 𝒮0\mathcal{S}_{0}. The latter result can be written as follows.

Lemma 2.1 (Comparison Principle)

Assume that U0,V0𝒮0U_{0},V_{0}\in\mathcal{S}_{0} and U0V0U_{0}\leq V_{0}, then the mild solutions U(t)U(t) and V(t)V(t) to (2.3) with initial data U0U_{0} and V0V_{0} respectively, satisfy U(t)V(t)U(t)\leq V(t) for any t0t\geq 0.

It follows that the comparison principle also holds for (1.10).

2.2 Limiting behavior

In this sub-section, we consider the steady state equation (2.4) below. The study of the problem will be used in the sequel to study the limit behavior of traveling wave solutions.

{du(a)da=0a+K(a,a)π(a)u(a)𝑑a(1u(a)),a[0,a+],u(0)=0a+γ(a)u(a)𝑑a.\begin{cases}\frac{du(a)}{da}=\int_{0}^{a^{+}}K(a,a^{\prime})\pi(a^{\prime})u(a^{\prime})da^{\prime}\,(1-u(a)),\;a\in[0,a^{+}],\\ u(0)=\int_{0}^{a^{+}}\gamma(a)u(a)da.\end{cases} (2.4)

Due to Assumption 1.1 (i)(i), let us observe that (2.4) has constant solutions: a0a\mapsto 0 and a1a\mapsto 1. We will show that (2.4) only has these two solutions with values into [0,1][0,1]. More specifically we have the following result.

Lemma 2.2

Problem (2.4) only has the two solutions uW1,1(0,a+)u\in W^{1,1}(0,a^{+}) such that 0u10\leq u\leq 1: u0u\equiv 0 and u1u\equiv 1.

Proof. To prove the above lemma we argue by contradiction by assuming that there exists a function uW1,1(0,a+)u^{*}\in W^{1,1}(0,a^{+}) with u0,1u^{*}\not\equiv 0,1, which is another solution of (2.4) satisfying 0u(a)10\leq u^{*}(a)\leq 1 for all a[0,a+]a\in[0,a^{+}]. Observe that since KK is continuous and uu^{*} is continuous by Sobolev embedding, then au(a)a\mapsto u^{*}(a) is of class C1C^{1} on [0,a+][0,a^{+}].

First we claim that 0<u(a)<10<u^{*}(a)<1 holds for all a[0,a+]a\in[0,a^{+}]. To prove this claim first note that if uu^{*} touches 11 at some a0[0,a+]a_{0}\in[0,a^{+}], then by the Cauchy Lipschitz theorem for ODE, one has u1u^{*}\equiv 1, which is a contradiction. Next if uu^{*} touches 0 at some point a0[0,a+]a_{0}\in[0,a^{+}], namely u(a0)=0u^{*}(a_{0})=0, then two cases may appear: either a0=0a_{0}=0 or a0(0,a+]a_{0}\in(0,a^{+}]. If a0=0a_{0}=0 then the boundary condition at a=0a=0 yields

0=u(a0)=0a+γ(a)u(a)𝑑a.0=u(a_{0})=\int_{0}^{a^{+}}\gamma(a)u^{*}(a)da.

But since γ0\gamma\not\equiv 0 there exists a1(0,a+)a_{1}\in(0,a^{+}) such that u(a1)=0u^{*}(a_{1})=0. Therefore it is sufficient to consider the case where there exists a0(0,a+]a_{0}\in(0,a^{+}] such that u(a0)=0u^{*}(a_{0})=0. Now to reach a contradiction, recalling u0u^{*}\geq 0, there holds u(a0)=0u(a_{0})=0 with a0(0,a+]a_{0}\in(0,a^{+}] which ensures that

du(a0)da0.\frac{du(a_{0})}{da}\leq 0.

Plugging this information together with u(a0)=0u^{*}(a_{0})=0 into (2.4) yields

0a+K(a0,a)π(a)u(a)𝑑a=0,\int_{0}^{a^{+}}K(a_{0},a^{\prime})\pi(a^{\prime})u^{*}(a^{\prime})da^{\prime}=0,

which by the positivity of KK and π\pi implies that u(a)0u^{*}(a)\equiv 0, which is a contradiction again. Thus we have obtained that 0<u(a)<10<u^{*}(a)<1 for all a[0,a+]a\in[0,a^{+}].

Now to complete the proof of the lemma let us obtain a contradiction with the existence of uu^{*}. To that aim let us show that one has u(a)1u^{*}(a)\geq 1 for all a[0,a+]a\in[0,a^{+}] so that u1u^{*}\equiv 1, a contradiction with u1u^{*}\not\equiv 1. To reach this issue, let us define κ\kappa^{*} by

κ:=inf{κ>0:κu(a)1,a[0,a+]}.\kappa^{*}\mathrel{\mathop{\mathchar 58\relax}}=\inf\{\kappa>0\mathrel{\mathop{\mathchar 58\relax}}\kappa u^{*}(a)\geq 1,\,\,\forall a\in[0,a^{+}]\}.

Note that κ\kappa^{*} is well-defined. Let us show that κ1\kappa^{*}\leq 1 by a contradiction argument and assuming that κ>1\kappa^{*}>1. Since κu(a)1\kappa^{*}u^{*}(a)\geq 1 for all a[0,a+]a\in[0,a^{+}], we consider two cases: (i) κu1\kappa^{*}u^{*}\geq 1 and κu1\kappa^{*}u^{*}\not\equiv 1 and (ii) κu1\kappa^{*}u^{*}\equiv 1.

Case (i) κu1,κu1\kappa^{*}u^{*}\geq 1,\kappa^{*}u^{*}\not\equiv 1. Set v(a)=κu(a)v^{*}(a)=\kappa^{*}u^{*}(a) that satisfies the problem

{dv(a)da=0a+K(a,a)π(a)v(a)𝑑a(11κv(a)),a[0,a+],v(0)=0a+γ(a)v(a)𝑑a.\begin{cases}\frac{dv^{*}(a)}{da}=\int_{0}^{a^{+}}K(a,a^{\prime})\pi(a^{\prime})v^{*}(a^{\prime})da^{\prime}\,\left(1-\frac{1}{\kappa^{*}}v^{*}(a)\right),\;a\in[0,a^{+}],\\ v^{*}(0)=\int_{0}^{a^{+}}\gamma(a)v^{*}(a)da.\end{cases} (2.5)

Note that from the definition of κ\kappa^{*}, there exists a0[0,a+]a_{0}\in[0,a^{+}] such that v(a0)=1v^{*}(a_{0})=1. Due to the condition

0a+γ(a)𝑑a=1,\int_{0}^{a^{+}}\gamma(a)da=1,

as above, the boundary condition ensures that one can choose a0(0,a+]a_{0}\in(0,a^{+}]. Hence we have

dv(a0)da0,\frac{dv^{*}(a_{0})}{da}\leq 0,

so that

0a+K(a0,a)π(a)v(a)𝑑a(11κ)0.\int_{0}^{a^{+}}K(a_{0},a^{\prime})\pi(a^{\prime})v^{*}(a^{\prime})da^{\prime}\,\left(1-\frac{1}{\kappa^{*}}\right)\leq 0.

Since κ>1\kappa^{*}>1 and K>0K>0, this yields v0v^{*}\equiv 0, a contradiction with 0<u<10<u^{*}<1, which proves κ1\kappa^{*}\leq 1 in the first case (i).

Case (ii) κu=v1\kappa^{*}u^{*}=v^{*}\equiv 1. In that case (2.5) yields for all a(0,a+)a\in(0,a^{+}) one has

0=0a+K(a,a)π(a)𝑑a(11κ),0=\int_{0}^{a^{+}}K(a,a^{\prime})\pi(a^{\prime})da^{\prime}\,\left(1-\frac{1}{\kappa^{*}}\right),

which is a contradiction again, since κ>1\kappa^{*}>1 and thus the right hand side is positive. Thus we still have κ1\kappa^{*}\leq 1.
To sum-up we have proved that u(a)1u^{*}(a)\geq 1 for all a[0,a+]a\in[0,a^{+}]. We also have 1u1\geq u^{*}. This proves that 11 is the unique positive solution of (2.4) and complete the proof of the lemma.  

2.3 Construction of sub-and super-solutions

In this section, we will construct exponentially bounded sub and super-solutions for problem (1.9). Exponential bound is used to define the convolution with the kernel JJ. As explained above, theses functions will be of particular importance to derive our existence result.

Super-solution

We start with the construction of a super-solution. We are looking for an exponentially bounded function U¯=U¯(a,ξ)>0\overline{U}=\overline{U}(a,\xi)>0 satisfying

{U(a,ξ)a(JξUU)(a,ξ)+cU(a,ξ)ξ+(1U(a,ξ))0a+K(a,a)π(a)U(a,ξ)𝑑a,U(0,ξ)0a+γ(a)U(a,ξ)𝑑a,\begin{cases}\frac{\partial U(a,\xi)}{\partial a}\geq\left(J\ast_{\xi}U-U\right)(a,\xi)+c\frac{\partial U(a,\xi)}{\partial\xi}+(1-U(a,\xi))\int_{0}^{a^{+}}K(a,a^{\prime})\pi(a^{\prime})U(a^{\prime},\xi)da^{\prime},\\ U(0,\xi)\geq\int_{0}^{a^{+}}\gamma(a)U(a,\xi)da,\end{cases} (2.6)

for some speed c>0c>0 to be specified latter.

Since U¯>0\overline{U}>0 it is sufficient to construct U¯\overline{U} as a solution of the following linear problem

{U(a,ξ)a(JξUU)(a,ξ)+cU(a,ξ)ξ+0a+K(a,a)π(a)U(a,ξ)𝑑a,U(0,ξ)0a+γ(a)U(a,ξ)𝑑a.\begin{cases}\frac{\partial U(a,\xi)}{\partial a}\geq\left(J\ast_{\xi}U-U\right)(a,\xi)+c\frac{\partial U(a,\xi)}{\partial\xi}+\int_{0}^{a^{+}}K(a,a^{\prime})\pi(a^{\prime})U(a^{\prime},\xi)da^{\prime},\\ U(0,\xi)\geq\int_{0}^{a^{+}}\gamma(a)U(a,\xi)da.\end{cases} (2.7)

We now construct such a super-solution by looking for it with the following form

U¯(a,ξ)=eλξϕ(a),\overline{U}(a,\xi)=e^{-\lambda\xi}\phi(a),

with λ>0\lambda>0 and ϕ(a)>0\phi(a)>0. Due to the differential inequality (2.7), we are looking for λ>0\lambda>0 and ϕ>0\phi>0 such that

{U¯(a,ξ)a=(JξU¯U¯)(a,ξ)+cU¯(a,ξ)ξ+0a+K(a,a)π(a)U¯(a,ξ)𝑑a,U¯(0,ξ)=0a+γ(a)U¯(a,ξ)𝑑a.\begin{cases}\frac{\partial\overline{U}(a,\xi)}{\partial a}=\left(J\ast_{\xi}\overline{U}-\overline{U}\right)(a,\xi)+c\frac{\partial\overline{U}(a,\xi)}{\partial\xi}+\int_{0}^{a^{+}}K(a,a^{\prime})\pi(a^{\prime})\overline{U}(a^{\prime},\xi)da^{\prime},\\ \overline{U}(0,\xi)=\int_{0}^{a^{+}}\gamma(a)\overline{U}(a,\xi)da.\end{cases} (2.8)

To solve this problem, we end-up with the following equation for λ>0,c\lambda>0,c\in\mathbb{R} and ϕ=ϕ(a)>0\phi=\phi(a)>0.

{ϕ(a)=Λ(λ,c)ϕ(a)+0a+K(a,a)π(a)ϕ(a)𝑑a,ϕ(0)=0a+γ(a)ϕ(a)𝑑a,\begin{cases}\phi^{\prime}(a)=\Lambda(\lambda,c)\phi(a)+\int_{0}^{a^{+}}K(a,a^{\prime})\pi(a^{\prime})\phi(a^{\prime})da^{\prime},\\ \phi(0)=\int_{0}^{a^{+}}\gamma(a)\phi(a)da,\end{cases} (2.9)

wherein we have set

Λ(λ,c)=J(y)eλy𝑑y1cλ.\Lambda(\lambda,c)=\int_{\mathbb{R}}J(y)e^{\lambda y}dy-1-c\lambda.

Setting s=Λ(λ,c)s=\Lambda(\lambda,c), the above problem is equivalent to find ss\in\mathbb{R} and ϕ=ϕ(a)>0\phi=\phi(a)>0 such that

ϕ(a)=esa0a+γ(a)ϕ(a)𝑑a+0aes(al)0a+K(l,a)π(a)ϕ(a)𝑑a𝑑l,a[0,a+].\phi(a)=e^{sa}\int_{0}^{a^{+}}\gamma(a)\phi(a)da+\int_{0}^{a}e^{s(a-l)}\int_{0}^{a^{+}}K(l,a^{\prime})\pi(a^{\prime})\phi(a^{\prime})da^{\prime}dl,\;\forall a\in[0,a^{+}]. (2.10)

Define the linear operator Ls(L1(0,a+))L_{s}\in\mathcal{L}(L^{1}(0,a^{+})) for any parameter ss\in\mathbb{R} and ϕL1(0,a+)\phi\in L^{1}(0,a^{+}) as follows,

[Lsϕ](a):=esa0a+γ(a)ϕ(a)da+0aes(al)0a+K(l,a)π(a)ϕ(a)dadl.[L_{s}\phi](a)\mathrel{\mathop{\mathchar 58\relax}}=e^{sa}\int_{0}^{a^{+}}\gamma(a)\phi(a)da+\int_{0}^{a}e^{s(a-l)}\int_{0}^{a^{+}}K(l,a^{\prime})\pi(a^{\prime})\phi(a^{\prime})da^{\prime}dl. (2.11)

We claim now that LsL_{s} enjoys the following properties:

  • (1)

    LsL_{s} is a positive operator, that is Ls(L+1(0,a+))L+1(0,a+)L_{s}\left(L^{1}_{+}(0,a^{+})\right)\subset L^{1}_{+}(0,a^{+});

  • (2)

    LsL_{s} is a compact operator;

  • (3)

    LsL_{s} is a non-supporting operator.

Note that (1) is true due to Assumption 1.1-(i), (ii) and (iv). Next, applying Fréchet-Kolmogorov-Riesz compactness theorem to LsL_{s} yields (2). To prove (3), by taking the dual product between LsϕL_{s}\phi with ϕL+1(0,a+){0}\phi\in L^{1}_{+}(0,a^{+})\setminus\{0\} and any ψL+(0,a+){0}\psi\in L^{\infty}_{+}(0,a^{+})\setminus\{0\}, we obtain

ψ,Lsϕ=0a+γ(a)ϕ(a)𝑑a0a+ψ(a)esa𝑑a+0a+ψ(a)0aes(al)0a+K(l,a)π(a)ϕ(a)𝑑a𝑑l𝑑a.\langle\psi,L_{s}\phi\rangle=\int_{0}^{a^{+}}\gamma(a)\phi(a)da\int_{0}^{a^{+}}\psi(a)e^{sa}da+\int_{0}^{a^{+}}\psi(a)\int_{0}^{a}e^{s(a-l)}\int_{0}^{a^{+}}K(l,a^{\prime})\pi(a^{\prime})\phi(a^{\prime})da^{\prime}dlda.

Recalling Assumption 1.1-(ii) and (iv), one can see that the second term of the above equality is positive. Thus (3) is proved.

Hence by Krein-Rutman theorem, the spectral radius ρ(Ls)\rho(L_{s}) is the principal eigenvalue of LsL_{s} (see Sawashima [21]) and moreover the map sρ(Ls)s\to\rho(L_{s}) is continuous and strictly increasing (see Marek [20]). In addition, notice that ρ(Ls)0\rho(L_{s})\to 0 as ss\to-\infty.

On the other hand, for s=0s=0, applying the constant function a1a\mapsto 1 to L0L_{0} yields, due to Assumption 1.1-(i),

[L01](a)=1+0a0a+K(l,a)π(a)𝑑a𝑑l>1,a[0,a+],[L_{0}1](a)=1+\int_{0}^{a}\int_{0}^{a^{+}}K(l,a^{\prime})\pi(a^{\prime})da^{\prime}dl>1,\;\forall a\in[0,a^{+}],

so that ρ(L0)>1\rho(L_{0})>1. It follows that there exists a unique s0<0s_{0}<0 such that ρ(Ls0)=1\rho(L_{s_{0}})=1. We now consider the equation Λ(λ,c)=s0\Lambda(\lambda,c)=s_{0}, that reads as

J(y)eλy𝑑y1cλ=s0.\int_{\mathbb{R}}J(y)e^{\lambda y}dy-1-c\lambda=s_{0}. (2.12)

Thus we have the following lemma.

Lemma 2.3

There exists c>0c^{*}>0 such that equation (2.12) with respect to the unknown λ\lambda has no positive solution when c<cc<c^{*}, only one solution for c=cc=c^{*}, and exactly two solutions for c>cc>c^{*}. In addition for c>cc>c^{*}, these two solutions λ1\lambda_{1} and λ2\lambda_{2} are such that

0<λ1<λ(c)<λ2,0<\lambda_{1}<\lambda(c)<\lambda_{2},

where λ(c)\lambda(c) is the unique root of the following equation

J(y)eλ(c)yy𝑑y=c.\int_{\mathbb{R}}J(y)e^{\lambda(c)y}ydy=c.

We will see in the proof of the lemma that the critical value cc^{*} is implicitly given by the equation

J(y)eλ(c)y𝑑y1cλ(c)=s0,\int_{\mathbb{R}}J(y)e^{\lambda(c^{*})y}dy-1-c^{*}\lambda(c^{*})=s_{0}, (2.13)

which uniquely provides the value of c>0c^{*}>0.

Proof. First, observe that

Λ(λ,c)λ=J(y)eλyy𝑑yc,2Λ(λ,c)λ2=J(y)eλyy2𝑑y>0,\frac{\partial\Lambda(\lambda,c)}{\partial\lambda}=\int_{\mathbb{R}}J(y)e^{\lambda y}ydy-c,\;\;\frac{\partial^{2}\Lambda(\lambda,c)}{\partial\lambda^{2}}=\int_{\mathbb{R}}J(y)e^{\lambda y}y^{2}dy>0,

and

J(y)eλyy𝑑y=0yJ(y)[eλyeλy]𝑑y± as λ±.\int_{\mathbb{R}}J(y)e^{\lambda y}ydy=\int_{0}^{\infty}yJ(y)\left[e^{\lambda y}-e^{-\lambda y}\right]dy\to\pm\infty\text{ as }\lambda\to\pm\infty.

As a consequence, for any c>0c>0, the function λΛ(λ,c)λ\lambda\mapsto\frac{\partial\Lambda(\lambda,c)}{\partial\lambda} is non-decreasing from -\infty to \infty and, using again the symmetry of JJ, satisfies

Λ(0,c)λ=c<0.\frac{\partial\Lambda(0,c)}{\partial\lambda}=-c<0.

Hence for any c>0c>0 there exists a unique λ(c)>0\lambda(c)>0 such that J(y)eλyy𝑑yc<0\int_{\mathbb{R}}J(y)e^{\lambda y}ydy-c<0 (resp. >0>0) for all λ<λ(c)\lambda<\lambda(c) (resp. for all λ>λ(c)\lambda>\lambda(c)). We also deduce that for any c>0c>0 the function λΛ(λ,c)\lambda\to\Lambda(\lambda,c) is decreasing on (0,λ(c))(0,\lambda(c)) and increasing on (λ(c),)(\lambda(c),\infty).

As a consequence of the above analysis, the solution of (2.12) relies on the position of its minimum over [0,)[0,\infty), i.e. Λ(λ(c),c)\Lambda(\lambda(c),c), with respect to s0s_{0}. This quantity is given by the formula

Λ(λ(c),c)=J(y)eλ(c)y𝑑y1cλ(c).\Lambda(\lambda(c),c)=\int_{\mathbb{R}}J(y)e^{\lambda(c)y}dy-1-c\lambda(c).

To study the equation Λ(λ(c),c)=s0\Lambda(\lambda(c),c)=s_{0}, let us note that the implicit function theorem ensures that the map cλ(c)c\mapsto\lambda(c) is of class C1C^{1} with respect to c(0,)c\in(0,\infty), so is the function cΛ(λ(c),c)c\mapsto\Lambda(\lambda(c),c) and we have:

dΛ(λ(c),c)dc=Λ(λ(c),c)λλ(c)c+Λ(λ(c),c)c=λ(c)<0.\displaystyle\frac{d\Lambda(\lambda(c),c)}{dc}=\frac{\partial\Lambda(\lambda(c),c)}{\partial\lambda}\frac{\partial\lambda(c)}{\partial c}+\frac{\partial\Lambda(\lambda(c),c)}{\partial c}=-\lambda(c)<0.

The above expression implies that cΛ(λ(c),c)c\mapsto\Lambda(\lambda(c),c) is monotonically decreasing with respect to cc. Furthermore, this function takes the value 0 for c=0c=0 and tends to -\infty as cc\to\infty. Indeed since JJ is symmetric and has a unit mass, we have

J(y)eλ(c)y𝑑ycλ(c)\displaystyle\int_{\mathbb{R}}J(y)e^{\lambda(c)y}dy-c\lambda(c)
=\displaystyle= 0J(y)eλ(c)y[1λ(c)y]𝑑y+0J(y)eλ(c)y[1+λ(c)y]𝑑y, as c.\displaystyle\int_{0}^{\infty}J(y)e^{\lambda(c)y}\left[1-\lambda(c)y\right]dy+\int_{0}^{\infty}J(y)e^{-\lambda(c)y}\left[1+\lambda(c)y\right]dy\to-\infty,\text{ as }c\to\infty.

Recalling that s0<0s_{0}<0, this proves there exists a unique c>0c^{*}>0 such that

Λ(λ(c),c){<s0,c(c,),>s0,c(0,c),\Lambda(\lambda(c),c)\begin{cases}<s_{0},\;\forall c\in(c^{*},\infty),\\ >s_{0},\;\forall c\in(0,c^{*}),\end{cases}

that concludes the proof of the lemma.  

To sum-up for all c>cc>c^{*}, the function U¯(a,ξ)=eλξϕ(a)\overline{U}(a,\xi)=e^{-\lambda\xi}\phi(a), where λ\lambda is a solution of (2.12), is a super-solution, that satisfies (2.6). Note also that the constant function (a,ξ)1(a,\xi)\mapsto 1 is also a super-solution.

Sub-solution

In this subsection, we construct a suitable sub-solution. Let c>cc>c^{*} be fixed and let λ>0\lambda>0 be the smallest solution of (2.12). Moreover, let ϕ=ϕ(a)\phi=\phi(a) be defined as in (2.10) with s=s0s=s_{0}. Observe that the function aϕ(a)a\mapsto\phi(a) is continuous and ϕ(a)>0\phi(a)>0 for any a[0,a+]a\in[0,a^{+}].

Now our aim is to construct an exponentially bounded sub-solution with the speed cc, that is a solution U¯=U¯(a,ξ)\underline{U}=\underline{U}(a,\xi) such that

U¯min{1,U¯},\underline{U}\leq\min\left\{1,\overline{U}\right\},

and satisfying the following differential inequality

{U¯(a,ξ)a(JξU¯U¯)(a,ξ)+cU¯(a,ξ)ξ+(1U¯(a,ξ))0a+K(a,a)π(a)U¯(a,ξ)𝑑a,U¯(0,ξ)0a+γ(a)U¯(a,ξ)𝑑a.\begin{cases}\frac{\partial\underline{U}(a,\xi)}{\partial a}\leq\left(J\ast_{\xi}\underline{U}-\underline{U}\right)(a,\xi)+c\frac{\partial\underline{U}(a,\xi)}{\partial\xi}+(1-\underline{U}(a,\xi))\int_{0}^{a^{+}}K(a,a^{\prime})\pi(a^{\prime})\underline{U}(a^{\prime},\xi)da^{\prime},\\ \underline{U}(0,\xi)\leq\int_{0}^{a^{+}}\gamma(a)\underline{U}(a,\xi)da.\end{cases} (2.14)

Note that we have

1U¯max{0,1U¯}(1αeλξ)+,1-\underline{U}\geq\max\{0,1-\overline{U}\}\geq\left(1-\alpha e^{-\lambda\xi}\right)^{+},

for some sufficiently large constant α>0\alpha>0 and where the superscript ++ denotes the positive part.
Hence it is sufficient to look for an exponentially bounded function U¯\underline{U} satisfying

{U¯(a,ξ)aJξU¯(a,ξ)U¯(a,ξ)+cU¯(a,ξ)ξ+0a+K(a,a)π(a)U¯(a,ξ)𝑑a(1αeλξ)+,U¯(0,ξ)0a+γ(a)U¯(a,ξ)𝑑a.\begin{cases}\frac{\partial\underline{U}(a,\xi)}{\partial a}\leq J\ast_{\xi}\underline{U}(a,\xi)-\underline{U}(a,\xi)+c\frac{\partial\underline{U}(a,\xi)}{\partial\xi}+\int_{0}^{a^{+}}K(a,a^{\prime})\pi(a^{\prime})\underline{U}(a^{\prime},\xi)da^{\prime}\,(1-\alpha e^{-\lambda\xi})^{+},\\ \underline{U}(0,\xi)\leq\int_{0}^{a^{+}}\gamma(a)\underline{U}(a,\xi)da.\end{cases} (2.15)

We look for such a function U¯\underline{U} of the form

U¯(a,ξ)=(eλξke(λ+η)ξ)ϕ(a),\underline{U}(a,\xi)=(e^{-\lambda\xi}-ke^{-(\lambda+\eta)\xi})\phi(a), (2.16)

where the function ϕ\phi is defined in (2.10) with s=s0s=s_{0}, while k>0k>0 and η>0\eta>0 have to be determined. We split our analysis into two regions: 1αeλξ>01-\alpha e^{-\lambda\xi}>0 and (1αeλξ)+=0\left(1-\alpha e^{-\lambda\xi}\right)^{+}=0. We define ξM\xi_{M} by

ξM:=1λlnα,\xi_{M}\mathrel{\mathop{\mathchar 58\relax}}=\frac{1}{\lambda}\ln\alpha,

so that the two above regions becomes {ξ>ξM}\{\xi>\xi_{M}\} and {ξξM}\{\xi\leq\xi_{M}\}.
Case 1 : ξ>ξM\xi>\xi_{M}.
First simple computations show that U¯\underline{U} satisfies the differential inequality (2.15) for ξ>ξM\xi>\xi_{M} if and only if we have for all ξ>1λlogα\xi>\frac{1}{\lambda}\log\alpha and all a(0,a+)a\in(0,a^{+}),

kϕ(a)\displaystyle-k\phi(a) (J(y)eλy𝑑ycλJ(y)e(λ+η)y𝑑y+c(λ+η))\displaystyle\left(\int_{\mathbb{R}}J(y)e^{\lambda y}dy-c\lambda-\int_{\mathbb{R}}J(y)e^{(\lambda+\eta)y}dy+c(\lambda+\eta)\right) (2.17)
α0a+K(a,a)π(a)ϕ(a)𝑑aeλξ(keηξ).\displaystyle\leq\alpha\int_{0}^{a^{+}}K(a,a^{\prime})\pi(a^{\prime})\phi(a^{\prime})da^{\prime}\,e^{-\lambda\xi}(k-e^{\eta\xi}).

Then if we set

p(η):=J(y)eλydycλJ(y)e(λ+η)ydy+c(λ+η),p(\eta)\mathrel{\mathop{\mathchar 58\relax}}=\int_{\mathbb{R}}J(y)e^{\lambda y}dy-c\lambda-\int_{\mathbb{R}}J(y)e^{(\lambda+\eta)y}dy+c(\lambda+\eta),

the above inequality becomes for all ξ>1λlogα\xi>\frac{1}{\lambda}\log\alpha and all a(0,a+)a\in(0,a^{+}),

kϕ(a)p(η)α0a+K(a,a)π(a)ϕ(a)𝑑aeλξ(keηξ).-k\phi(a)p(\eta)\leq\alpha\int_{0}^{a^{+}}K(a,a^{\prime})\pi(a^{\prime})\phi(a^{\prime})da^{\prime}\,e^{-\lambda\xi}(k-e^{\eta\xi}).

Recalling that λ<λ(c)\lambda<\lambda(c), function satisfies p(0)=0p(0)=0 and p(0)>0p^{\prime}(0)>0. Therefore for any sufficiently small and nonnegative η\eta, we have p(η)>0p(\eta)>0. Next, for kk sufficiently large and η\eta small enough, due to infa[0,a+]ϕ(a)>0\inf_{a\in[0,a^{+}]}\phi(a)>0 it is sufficient to have for any ξ>ξM\xi>\xi_{M}:

kp(η)ϕ(a)+α0a+K(a,a)π(a)ϕ(a)𝑑ae(λ+η)ξ<0,a[0,a+].-kp(\eta)\phi(a)+\alpha\int_{0}^{a^{+}}K(a,a^{\prime})\pi(a^{\prime})\phi(a^{\prime})da^{\prime}\,e^{(-\lambda+\eta)\xi}<0,\;\forall a\in[0,a^{+}].

Therefore, with such a choice of parameters kk and η\eta, inequality (2.17) is satisfied for any ξ>ξM\xi>\xi_{M}, which completes Case 1.
Case 2 : ξξM\xi\leq\xi_{M}.
For ξξM\xi\leq\xi_{M}, U¯\underline{U} satisfies (2.15) if and only if the following inequality holds:

0a+K(a,a)π(a)ϕ(a)𝑑a(1keηξ)kϕ(a)eηξp(η), for all a[0,a+] and ξξM.\int_{0}^{a^{+}}K(a,a^{\prime})\pi(a^{\prime})\phi(a^{\prime})da^{\prime}\,(1-ke^{-\eta\xi})\leq k\phi(a)e^{-\eta\xi}p(\eta),\;\text{ for all $a\in[0,a^{+}]$ and $\xi\leq\xi_{M}$}.

In order to reach this inequality, it is sufficient to have

(supa[0,a+]0a+K(a,a)π(a)ϕ(a)𝑑a)eηξMkp(η)infa[0,a+]ϕ(a),\left(\sup_{a\in[0,a^{+}]}\int_{0}^{a^{+}}K(a,a^{\prime})\pi(a^{\prime})\phi(a^{\prime})da^{\prime}\right)e^{\eta\xi_{M}}\leq kp(\eta)\inf_{a\in[0,a^{+}]}\phi(a),

which is satisfied for sufficiently large values of kk, since p(η)>0p(\eta)>0. This completes the Case 2.
In summary, we have proved that U¯\underline{U} satisfies (2.15) for all ξ\xi\in\mathbb{R} as soon as η>0\eta>0 is small enough and K>0K>0 is chosen large enough.

Next, define the functions

w¯(a,ξ)=U¯(a,ξca),w¯(a,ξ)=U¯(a,ξca),\underline{w}(a,\xi)=\underline{U}(a,\xi-ca),\;\overline{w}(a,\xi)=\overline{U}(a,\xi-ca),

so that w¯w¯\underline{w}\leq\overline{w}. These functions will be used in the next section for the construction of wave solutions and to derive our existence result.

3 Traveling wave solution

In this section, we prove Theorems 1.3 and 1.4.

3.1 Existence

In this subsection, we will employ a monotone iteration method to derive an existence result for traveling wave solutions. To this aim, we study the following nonlinear problem with cc\in\mathbb{R}.

{ua(a,ξ)=(Jξuu)(a,ξ)+0a+K(a,a)π(a)u(a,ξ)𝑑a(1u(a,ξ)),(a,ξ)(0,a+)×,u(0,ξ)=0a+γ(a)u(a,ξ+ca)𝑑a,ξ.\begin{cases}\frac{\partial u}{\partial a}(a,\xi)=(J\ast_{\xi}u-u)(a,\xi)+\int_{0}^{a^{+}}K(a,a^{\prime})\pi(a^{\prime})u(a^{\prime},\xi)da^{\prime}\,(1-u(a,\xi)),&(a,\xi)\in(0,a^{+})\times\mathbb{R},\\ u(0,\xi)=\int_{0}^{a^{+}}\gamma(a)u(a,\xi+ca)da,&\xi\in\mathbb{R}.\end{cases} (3.1)
Definition 3.1

We say that a function u¯W1,1((0,a+),C())\overline{u}\in W^{1,1}((0,a^{+}),C(\mathbb{R})) satisfying

supa(0,a+)|u¯(a,ξ)|=𝒪(eτ|ξ|), as ξ±, for some τ>0\sup_{a\in(0,a^{+})}|\overline{u}(a,\xi)|=\mathcal{O}(e^{\tau|\xi|}),\text{ as }\xi\to\pm\infty,\text{ for some }\tau>0 (3.2)

is a super-solution of (3.1), if it satisfies that

{u¯a(a,ξ)(Jξu¯u¯)(a,ξ)+0a+K(a,a)π(a)u¯(a,ξ)𝑑a(1u¯(a,ξ)),(a,ξ)[0,a+]×,u¯(0,ξ)0a+γ(a)u¯(a,ξ+ca)𝑑a,ξ.\begin{cases}\frac{\partial\overline{u}}{\partial a}(a,\xi)\geq(J\ast_{\xi}\overline{u}-\overline{u})(a,\xi)+\int_{0}^{a^{+}}K(a,a^{\prime})\pi(a^{\prime})\overline{u}(a^{\prime},\xi)da^{\prime}\,(1-\overline{u}(a,\xi)),&(a,\xi)\in[0,a^{+}]\times\mathbb{R},\\ \overline{u}(0,\xi)\geq\int_{0}^{a^{+}}\gamma(a)\overline{u}(a,\xi+ca)da,&\xi\in\mathbb{R}.\end{cases} (3.3)

Similarly, we define a sub-solution by interchanging the inequalities. Here, (3.2) is used to guarantee the convolution Ju¯J\ast\overline{u} to be well-defined.

Proposition 3.2

Assume that there exists a pair of sub/super-solution of (3.1) such that 0u¯u¯10\leq\underline{u}\leq\overline{u}\leq 1, then there exist a minimal solution uu_{*} and a maximal solution uu^{*} of (3.1), in the sense that for any other solution u{vW1,1((0,a+),Lloc1()):u¯vu¯}u\in\{v\in W^{1,1}((0,a^{+}),L^{1}_{loc}(\mathbb{R}))\mathrel{\mathop{\mathchar 58\relax}}\underline{u}\leq v\leq\overline{u}\}, it holds that

u¯uuuu¯.\underline{u}\leq u_{*}\leq u\leq u^{*}\leq\overline{u}.

Proof. Define M>0M>0 as follows,

M=maxa[0,a+]0a+K(a,a)π(a)𝑑a,M=\max_{a\in[0,a^{+}]}\int_{0}^{a^{+}}K(a,a^{\prime})\pi(a^{\prime})da^{\prime}, (3.4)

and define the sequence of the functions (un)n0(u_{n})_{n\geq 0} as u0=u¯u_{0}=\underline{u} and for n1n\geq 1, for (a,ξ)[0,a+]×(a,\xi)\in[0,a^{+}]\times\mathbb{R} by

{un(a,ξ)a(JξununMun)(a,ξ)=0a+K(a,a)π(a)un1(a,ξ)𝑑a(1un1(a,ξ))+Mun1(a,ξ),un(0,ξ)=0a+γ(a)un1(a,ξ+ca)𝑑a.\begin{cases}\frac{\partial u_{n}(a,\xi)}{\partial a}-(J\ast_{\xi}u_{n}-u_{n}-Mu_{n})(a,\xi)=\int_{0}^{a^{+}}K(a,a^{\prime})\pi(a^{\prime})u_{n-1}(a^{\prime},\xi)da^{\prime}\,(1-u_{n-1}(a,\xi))+Mu_{n-1}(a,\xi),\\ u_{n}(0,\xi)=\int_{0}^{a^{+}}\gamma(a)u_{n-1}(a,\xi+ca)da.\end{cases} (3.5)

Next, set u0=u¯u^{0}=\overline{u} and define unu^{n} with n1n\geq 1, for (a,ξ)[0,a+]×(a,\xi)\in[0,a^{+}]\times\mathbb{R}, by the resolution of the problem

{un(a,ξ)a(JξununMun)(a,ξ)=0a+K(a,a)π(a)un1(a,ξ)𝑑a(1un1(a,ξ))+Mun1(a,ξ),un(0,ξ)=0a+γ(a)un1(a,ξ+ca)𝑑a.\begin{cases}\frac{\partial u^{n}(a,\xi)}{\partial a}-(J\ast_{\xi}u^{n}-u^{n}-Mu^{n})(a,\xi)=\int_{0}^{a^{+}}K(a,a^{\prime})\pi(a^{\prime})u^{n-1}(a^{\prime},\xi)da^{\prime}\,(1-u^{n-1}(a,\xi))+Mu^{n-1}(a,\xi),\\ u^{n}(0,\xi)=\int_{0}^{a^{+}}\gamma(a)u^{n-1}(a,\xi+ca)da.\end{cases} (3.6)

First we show that unW1,1((0,a+),C())u_{n}\in W^{1,1}((0,a^{+}),C(\mathbb{R})) is well defined. Note that

0a+K(a,a)π(a)u0(a,ξ)𝑑a(1u0(a,ξ))=0a+K(a,a)π(a)u¯(a,ξ)𝑑a(1u¯(a,ξ)),\int_{0}^{a^{+}}K(a,a^{\prime})\pi(a^{\prime})u_{0}(a^{\prime},\xi)da^{\prime}\,(1-u_{0}(a,\xi))=\int_{0}^{a^{+}}K(a,a^{\prime})\pi(a^{\prime})\underline{u}(a^{\prime},\xi)da^{\prime}\,(1-\underline{u}(a,\xi)),

so that u1u_{1} is well defined by the linear nonlocal diffusion equations. Analogously by induction we get that unu_{n} is well defined. Similarly, the existence of unu^{n} follows the same induction arguments.

We will show that the sequence unu_{n} (respectively unu^{n}) is increasing (respectively decreasing) in the sense that

u¯unun+1un+1unu¯.\underline{u}\leq\cdots\leq u_{n}\leq u_{n+1}\leq u^{n+1}\leq u^{n}\leq\cdots\leq\overline{u}. (3.7)

Indeed, taking w:=u1u0w\mathrel{\mathop{\mathchar 58\relax}}=u_{1}-u_{0}, it satisfies

{w(a,ξ)a(Jξw+w+Mw)(a,ξ)0,w(0,ξ)0.\begin{cases}\frac{\partial w(a,\xi)}{\partial a}-(J\ast_{\xi}w+w+Mw)(a,\xi)\geq 0,\\ w(0,\xi)\geq 0.\end{cases}

Using the comparison principle of nonlocal diffusion problems, we conclude that w0w\geq 0, i.e. u¯=u0u1\underline{u}=u_{0}\leq u_{1}. Now assume that un1unu_{n-1}\leq u_{n}. Observe that

0a+K(a,a)π(a)un(a,ξ)𝑑a(1un(a,ξ))+Mun(a,ξ)\displaystyle\int_{0}^{a^{+}}K(a,a^{\prime})\pi(a^{\prime})u_{n}(a^{\prime},\xi)da^{\prime}\,(1-u_{n}(a,\xi))+Mu_{n}(a,\xi)
0a+K(a,a)π(a)un1(a,ξ)𝑑a(1un1(a,ξ))Mun1(a,ξ)\displaystyle-\int_{0}^{a^{+}}K(a,a^{\prime})\pi(a^{\prime})u_{n-1}(a^{\prime},\xi)da^{\prime}\,(1-u_{n-1}(a,\xi))-Mu_{n-1}(a,\xi)
\displaystyle\geq 0a+K(a,a)π(a)un1(a,ξ)𝑑a(un1un)(a,ξ)+M(unun1)(a,ξ)\displaystyle\int_{0}^{a^{+}}K(a,a^{\prime})\pi(a^{\prime})u_{n-1}(a^{\prime},\xi)da^{\prime}(u_{n-1}-u_{n})(a,\xi)\,+M(u_{n}-u_{n-1})(a,\xi)
\displaystyle\geq 0.\displaystyle 0.

Then w:=un+1unw\mathrel{\mathop{\mathchar 58\relax}}=u_{n+1}-u_{n} satisfies

{w(a,ξ)a(Jξw+w+Mw)(a,ξ)0,w(0,ξ)0a+γ(a)(unun1)(a,ξ+ca)𝑑a0.\begin{cases}\frac{\partial w(a,\xi)}{\partial a}-(J\ast_{\xi}w+w+Mw)(a,\xi)\geq 0,\\ w(0,\xi)\geq\int_{0}^{a^{+}}\gamma(a)(u_{n}-u_{n-1})(a,\xi+ca)da\geq 0.\end{cases}

Again the comparison principle shows that unun+1u_{n}\leq u_{n+1}. The rest of the inequalities in (3.7) can be proved similarly.

Next, for all (a,ξ)(0,a+)×(a,\xi)\in(0,a^{+})\times\mathbb{R}, un(a,ξ)u_{n}(a,\xi) has a limit as nn\to\infty by monotone convergence theorem, denoted by u(a,ξ)u_{*}(a,\xi); that is un(a,ξ)u(a,ξ)u_{n}(a,\xi)\to u_{*}(a,\xi) for all (0,a+)×(0,a^{+})\times\mathbb{R}. Thus, one has that for all ξ\xi\in\mathbb{R},

0a+γ(a)un(a,ξ+ca)𝑑an0a+γ(a)u(a,ξ+ca)𝑑a.\displaystyle\int_{0}^{a^{+}}\gamma(a)u_{n}(a,\xi+ca)da\xrightarrow{n\to\infty}\int_{0}^{a^{+}}\gamma(a)u_{*}(a,\xi+ca)da. (3.8)

In addition, one has

φn(a,ξ)\displaystyle\varphi_{n}(a,\xi) :=\displaystyle\mathrel{\mathop{\mathchar 58\relax}}= Jξun(a,ξ)un(a,ξ)+0a+K(a,a)π(a)un(a,ξ)𝑑a(1un(a,ξ))\displaystyle J\ast_{\xi}u_{n}(a,\xi)-u_{n}(a,\xi)+\int_{0}^{a^{+}}K(a,a^{\prime})\pi(a^{\prime})u_{n}(a^{\prime},\xi)da^{\prime}\,(1-u_{n}(a,\xi))
n\displaystyle\xrightarrow{n\to\infty} Jξu(a,ξ)u(a,ξ)+0a+K(a,a)π(a)u(a,ξ)𝑑a(1u(a,ξ))\displaystyle J\ast_{\xi}u_{*}(a,\xi)-u_{*}(a,\xi)+\int_{0}^{a^{+}}K(a,a^{\prime})\pi(a^{\prime})u_{*}(a^{\prime},\xi)da^{\prime}\,(1-u_{*}(a,\xi))
:=\displaystyle\mathrel{\mathop{\mathchar 58\relax}}= φ(a,ξ),\displaystyle\varphi(a,\xi),

in (0,a+)×(0,a^{+})\times\mathbb{R} and in L1((0,a+),Lloc1())L^{1}((0,a^{+}),L^{1}_{loc}(\mathbb{R})). Hence, for all ξ\xi\in\mathbb{R} and (η,θ)(0,a+)(\eta,\theta)\subset(0,a^{+}), one has

un(θ,ξ)un(η,ξ)=ηθφn(a,ξ)𝑑a,n0,u_{n}(\theta,\xi)-u_{n}(\eta,\xi)=\int_{\eta}^{\theta}\varphi_{n}(a,\xi)da,\;\forall n\geq 0,

which implies by letting nn\to\infty that

u(θ,ξ)u(η,ξ)=ηθφ(a,ξ)𝑑a.u_{*}(\theta,\xi)-u_{*}(\eta,\xi)=\int_{\eta}^{\theta}\varphi(a,\xi)da.

Since φL1((0,a+),Lloc1())\varphi\in L^{1}((0,a^{+}),L^{1}_{loc}(\mathbb{R})), it follows that uW1,1((0,a+),Lloc1())u_{*}\in W^{1,1}((0,a^{+}),L^{1}_{loc}(\mathbb{R})) satisfies the first equation of (3.1) with au=φ\partial_{a}u_{*}=\varphi a.e. in (0,a+)×(0,a^{+})\times\mathbb{R}. Further, u(,ξ)u_{*}(\cdot,\xi) is continuous in [0,a+][0,a^{+}] for all ξ\xi\in\mathbb{R}. Thus, one has un(0,ξ)u(0,ξ)u_{n}(0,\xi)\to u_{*}(0,\xi) as nn\to\infty in \mathbb{R}, which by (3.8) implies that

u(0,ξ)=0a+γ(a)u(a,ξ+ca)𝑑a, a.e. ξ.u_{*}(0,\xi)=\int_{0}^{a^{+}}\gamma(a)u_{*}(a,\xi+ca)da,\text{ a.e. }\xi\in\mathbb{R}.

It follows that uW1,1((0,a+),Lloc1())u_{*}\in W^{1,1}((0,a^{+}),L^{1}_{loc}(\mathbb{R})) satisfies the equation (3.1).

We now claim that uu_{*} is the minimal solution of (3.1). Indeed, if uu is a solution of (3.1) such that u[u¯,u¯]u\in[\underline{u},\overline{u}], it can be shown that the sequence unu_{n} built in (3.5) satisfies that u¯unu\underline{u}\leq u_{n}\leq u. Hence, unuuu_{n}\uparrow u_{*}\leq u. We can argue similarly with the sequence unu^{n} and conclude uu^{*} is maximal. This completes the proof.  

Now define

u¯(a,ξ)=max{w¯(a,ξ), 0} and u¯(a,ξ)=min{1,w¯(a,ξ)}.\underline{u}(a,\xi)=\max\{\underline{w}(a,\xi),\,0\}\text{ and }\overline{u}(a,\xi)=\min\{1,\,\overline{w}(a,\xi)\}. (3.9)

By the constructions proposed in Section 2.3, u¯\underline{u} and u¯\overline{u} are indeed the sub- and super-solutions of (3.1) respectively. Next, due to u0=u¯1u^{0}=\overline{u}\leq 1 and recalling MM defined in (3.4), Proposition 3.2 applies to the system (1.9) and one concludes to the existence of a maximal solution wW1,1((0,a+),Lloc1())w\in W^{1,1}((0,a^{+}),L^{1}_{loc}(\mathbb{R})) satisfying

max{w¯(a,ξ), 0}w(a,ξ)min{1,w¯(a,ξ)}.\max\{\underline{w}(a,\xi),\,0\}\leq w(a,\xi)\leq\min\{1,\,\overline{w}(a,\xi)\}.

It follows that w(a,ξ)0w(a,\xi)\to 0 as ξ\xi\to\infty uniformly in a[0,a+]a\in[0,a^{+}]. Hence to complete the proof of Theorem 1.3, it remains to study the behavior of ww as ξ\xi\to-\infty. Note that the function ξw¯(a,ξ)\xi\to\overline{w}(a,\xi) is nonincreasing on \mathbb{R}, then the limit function ξw(a,ξ)\xi\to w(a,\xi) is also nonincreasing on \mathbb{R} for a.e. a(0,a+)a\in(0,a^{+}).

On the other hand, recalling the definition of w¯\underline{w} in (2.16), there exists ξ0\xi_{0} large enough such that

infa(0,a+)w¯(a,ξ0)>0.\inf_{a\in(0,a^{+})}\underline{w}(a,\xi_{0})>0.

Hence due to the monotonicity of ww with respect to ξ\xi, it follows that

limξinfa(0,a+)w(a,ξ)>0.\lim_{\xi\to-\infty}\inf_{a\in(0,a^{+})}w(a,\xi)>0.

Moreover the monotonicity of ξw(a,ξ)\xi\mapsto w(a,\xi), for any a[0,a+]a\in[0,a^{+}], ensures that the limit w(a):=limξw(a,ξ)w^{-}(a)\mathrel{\mathop{\mathchar 58\relax}}=\lim\limits_{\xi\to-\infty}w(a,\xi) does exist for any a[0,a+]a\in[0,a^{+}]. Hence, since 0w(a,ξ)10\leq w(a,\xi)\leq 1 for all (a,ξ)[0,a+]×(a,\xi)\in[0,a^{+}]\times\mathbb{R}, the Lebesgue convergence theorem yields

limξJ(y)[w(a,ξy)w(a,ξ)]𝑑y=J(y)[w(a)w(a)]𝑑y=0,a[0,a+].\lim\limits_{\xi\to-\infty}\int_{\mathbb{R}}J(y)[w(a,\xi-y)-w(a,\xi)]dy=\int_{\mathbb{R}}J(y)[w^{-}(a)-w^{-}(a)]dy=0,\;\forall a\in[0,a^{+}].

As a consequence ww^{-} satisfies (2.4). Now recall that (2.4) has a unique positive solution, the constant function a1a\mapsto 1, which is shown in Section 2.2, thus we have

limξw(a,ξ)=1, for all a[0,a+],\lim\limits_{\xi\to-\infty}w(a,\xi)=1,\text{ for all $a\in[0,a^{+}]$},

the convergence being in fact uniform for a[0,a+]a\in[0,a^{+}] since aw(a,ξ)\partial_{a}w(a,\xi) is globally bounded.

3.2 Regularity

At last, we show the regularity of ww with respect to ξ\xi. To this aim, let us fix λ>0\lambda>0 as the smallest solution of (2.12). We now summarize some Lipschitz regularity estimates in the following lemma, motivated by Shen and Shen [22] and Ducrot and Jin [7].

Lemma 3.3

There exists some constant mλm\geq\lambda large enough such that for all n1n\geq 1, one has

|un(a,ξ+h)un(a,ξ)|em|h|1,(a,ξ)[0,a+]×.|u^{n}(a,\xi+h)-u^{n}(a,\xi)|\leq e^{m|h|}-1,\;\forall(a,\xi)\in[0,a^{+}]\times\mathbb{R}.

Proof. In the following we prove it for h>0h>0. The case for h<0h<0 can be proved similarly.

Recall that u0(a,ξ)=u¯(a,ξ)u^{0}(a,\xi)=\overline{u}(a,\xi) defined in (3.9) is given as follows:

u0(a,ξ)={eλ(ξca)ϕ(a), if ξca+1λlnϕ(a),1, if ξ<ca+1λlnϕ(a),u^{0}(a,\xi)=\begin{cases}e^{-\lambda(\xi-ca)}\phi(a),&\text{ if }\xi\geq ca+\frac{1}{\lambda}\ln\phi(a),\\ 1,&\text{ if }\xi<ca+\frac{1}{\lambda}\ln\phi(a),\end{cases}

while for h>0h>0, one has

u0(a,ξ+h)={eλ(ξ+hca)ϕ(a), if ξ+hca+1λlnϕ(a),1, if ξ+h<ca+1λlnϕ(a).u^{0}(a,\xi+h)=\begin{cases}e^{-\lambda(\xi+h-ca)}\phi(a),&\text{ if }\xi+h\geq ca+\frac{1}{\lambda}\ln\phi(a),\\ 1,&\text{ if }\xi+h<ca+\frac{1}{\lambda}\ln\phi(a).\end{cases}

Then we infer from these formulas,

u0(a,ξ+h)u0(a,ξ)\displaystyle\frac{u^{0}(a,\xi+h)}{u^{0}(a,\xi)}
=\displaystyle\!\!=\!\!\!\! {eλh, if ξca+1λlnϕ(a),eλ(ξ+hca)ϕ(a), if ca+1λlnϕ(a)hξ<ca+1λlnϕ(a),1, if ξ+h<ca+1λlnϕ(a).\displaystyle\!\!\begin{cases}e^{-\lambda h},\!\!\!\!&\text{ if }\xi\geq ca+\frac{1}{\lambda}\ln\phi(a),\\ e^{-\lambda(\xi+h-ca)}\phi(a),\!\!\!\!&\text{ if }ca+\frac{1}{\lambda}\ln\phi(a)-h\leq\xi<ca+\frac{1}{\lambda}\ln\phi(a),\\ 1,\!\!\!\!&\text{ if }\xi+h<ca+\frac{1}{\lambda}\ln\phi(a).\end{cases}

Hence one can choose m>λm>\lambda large enough such that

emhu0(a,ξ+h)u0(a,ξ)1,(a,ξ)[0,a+]×.e^{-mh}\leq\frac{u^{0}(a,\xi+h)}{u^{0}(a,\xi)}\leq 1,\;\forall(a,\xi)\in[0,a^{+}]\times\mathbb{R}.

Now consider the functions vn,h=vn,h(a,ξ),n0v_{n,h}=v_{n,h}(a,\xi),n\geq 0 given by

vn,h(a,ξ):=un(a,ξ+h)emh.v_{n,h}(a,\xi)\mathrel{\mathop{\mathchar 58\relax}}=\frac{u^{n}(a,\xi+h)}{e^{-mh}}. (3.10)

Next, we choose M>0M>0 a little bit different from (3.4), since now vn,hv_{n,h} could be larger than 11. Observe that vn,hemhv_{n,h}\leq e^{mh} for all n0n\geq 0, so set

M:=emhmaxa[0,a+]0a+K(a,a)π(a)da.M\mathrel{\mathop{\mathchar 58\relax}}=e^{mh}\,\max_{a\in[0,a^{+}]}\int_{0}^{a^{+}}K(a,a^{\prime})\pi(a^{\prime})da^{\prime}.

Direct computations for any 0v10\leq v\leq 1 and vuemhv\leq u\leq e^{mh} yields

0a+K(a,a)π(a)u(a,ξ)𝑑a(1u(a,ξ))+Mu(a,ξ)\displaystyle\int_{0}^{a^{+}}K(a,a^{\prime})\pi(a^{\prime})u(a^{\prime},\xi)da^{\prime}\,(1-u(a,\xi))+Mu(a,\xi) (3.11)
0a+K(a,a)π(a)v(a,ξ)𝑑a(1v(a,ξ))Mv(a,ξ)\displaystyle-\int_{0}^{a^{+}}K(a,a^{\prime})\pi(a^{\prime})v(a^{\prime},\xi)da^{\prime}\,(1-v(a,\xi))-Mv(a,\xi)
=\displaystyle= 0a+K(a,a)π(a)(u(a,ξ)v(a,ξ))𝑑a(1v(a,ξ))\displaystyle\int_{0}^{a^{+}}K(a,a^{\prime})\pi(a^{\prime})(u(a^{\prime},\xi)-v(a^{\prime},\xi))da^{\prime}(1-v(a,\xi))
0a+K(a,a)π(a)(u(a,ξ)v(a,ξ))𝑑a(u(a,ξ)v(a,ξ))\displaystyle-\int_{0}^{a^{+}}K(a,a^{\prime})\pi(a^{\prime})(u(a^{\prime},\xi)-v(a^{\prime},\xi))da^{\prime}(u(a,\xi)-v(a,\xi))
0a+K(a,a)π(a)v(a,ξ)𝑑a(u(a,ξ)v(a,ξ))+M(uv)(a,ξ)\displaystyle-\int_{0}^{a^{+}}K(a,a^{\prime})\pi(a^{\prime})v(a^{\prime},\xi)da^{\prime}(u(a,\xi)-v(a,\xi))+M(u-v)(a,\xi)
\displaystyle\geq (M0a+K(a,a)π(a)u(a,ξ)𝑑a)(uv)(a,ξ)\displaystyle\left(M-\int_{0}^{a^{+}}K(a,a^{\prime})\pi(a^{\prime})u(a^{\prime},\xi)da^{\prime}\right)(u-v)(a,\xi)
\displaystyle\geq 0.\displaystyle 0.

First let us consider n=1n=1. Observe that v1,hv_{1,h} satisfies the following equation,

{v1,ha=Jξv1,hv1,hMv1,h+Mv0,h+0a+K(a,a)π(a)v0,h(a,ξ)𝑑a(1u0(a,ξ+h)),v1,h(0,ξ)=0a+γ(a)v0,h(a,ξ+ca)𝑑a.\begin{cases}\frac{\partial v_{1,h}}{\partial a}=J\ast_{\xi}v_{1,h}-v_{1,h}-Mv_{1,h}+Mv_{0,h}+\int_{0}^{a^{+}}K(a,a^{\prime})\pi(a^{\prime})v_{0,h}(a^{\prime},\xi)da^{\prime}\,(1-u^{0}(a,\xi+h)),\\ v_{1,h}(0,\xi)=\int_{0}^{a^{+}}\gamma(a)v_{0,h}(a,\xi+ca)da.\end{cases}

It follows from v0,h(a,ξ)u0(a,ξ+h)v_{0,h}(a,\xi)\geq u^{0}(a,\xi+h) that

{v1,haJξv1,hv1,hMv1,h+Mv0,h+0a+K(a,a)π(a)v0,h(a,ξ)𝑑a(1v0,h),v1,h(0,ξ)=0a+γ(a)v0,h(a,ξ+ca)𝑑a.\begin{cases}\frac{\partial v_{1,h}}{\partial a}\geq J\ast_{\xi}v_{1,h}-v_{1,h}-Mv_{1,h}+Mv_{0,h}+\int_{0}^{a^{+}}K(a,a^{\prime})\pi(a^{\prime})v_{0,h}(a^{\prime},\xi)da^{\prime}\,(1-v_{0,h}),\\ v_{1,h}(0,\xi)=\int_{0}^{a^{+}}\gamma(a)v_{0,h}(a,\xi+ca)da.\end{cases}

On the other hand, recall that u1u^{1} satisfies (3.6), thus due to the monotonicity (3.11) and v0,hu0v_{0,h}\geq u^{0}, we have

Mv0,h(a,ξ)+0a+K(a,a)π(a)v0,h(a,ξ)𝑑a(1v0,h(a,ξ))\displaystyle Mv_{0,h}(a,\xi)+\int_{0}^{a^{+}}K(a,a^{\prime})\pi(a^{\prime})v_{0,h}(a^{\prime},\xi)da^{\prime}\,(1-v_{0,h}(a,\xi))
\displaystyle\geq Mu0(a,ξ)+0a+K(a,a)π(a)u0(a,ξ)𝑑a(1u0(a,ξ)),\displaystyle Mu^{0}(a,\xi)+\int_{0}^{a^{+}}K(a,a^{\prime})\pi(a^{\prime})u^{0}(a^{\prime},\xi)da^{\prime}\,(1-u^{0}(a,\xi)),

and the initial data satisfies,

v1,h(0,ξ)\displaystyle v_{1,h}(0,\xi) =\displaystyle= 0a+γ(a)v0,h(a,ξ+ca)𝑑a=emh0a+γ(a)u0(a,ξ+ca+h)𝑑a\displaystyle\int_{0}^{a^{+}}\gamma(a)v_{0,h}(a,\xi+ca)da=e^{mh}\int_{0}^{a^{+}}\gamma(a)u^{0}(a,\xi+ca+h)da
\displaystyle\geq 0a+γ(a)u0(a,ξ+ca)𝑑a=u1(0,ξ),ξ.\displaystyle\int_{0}^{a^{+}}\gamma(a)u^{0}(a,\xi+ca)da=u^{1}(0,\xi),\;\forall\xi\in\mathbb{R}.

The comparison principle applies and provides

u1(a,ξ)v1,h(a,ξ),(a,ξ)[0,a+]×.u^{1}(a,\xi)\leq v_{1,h}(a,\xi),\;\forall(a,\xi)\in[0,a^{+}]\times\mathbb{R}. (3.12)

Now since for all a[0,a+]a\in[0,a^{+}] the function ξu1(a,ξ)\xi\to u^{1}(a,\xi) is nonincreasing, for all a[0,a+]a\in[0,a^{+}] and ξ\xi\in\mathbb{R}, we get

|u1(a,ξ+h)u1(a,ξ)|u1(a,ξ)u1(a,ξ+h)\displaystyle|u^{1}(a,\xi+h)-u^{1}(a,\xi)|\leq u^{1}(a,\xi)-u^{1}(a,\xi+h) \displaystyle\leq (1emh)v1,h(a,ξ)\displaystyle(1-e^{-mh})v_{1,h}(a,\xi) (3.13)
\displaystyle\leq emh1.\displaystyle e^{mh}-1.

Next, let us prove the result (3.12) for any n>1n>1 by induction. Assume that vn,h(a,ξ)un(a,ξ)v_{n,h}(a,\xi)\geq u^{n}(a,\xi) for any (a,ξ)[0,a+]×(a,\xi)\in[0,a^{+}]\times\mathbb{R}. Observe that vn+1,hv_{n+1,h} satisfies the following equation,

{vn+1,ha=Jξvn+1,hvn+1,hMvn+1,h+Mvn,h+0a+K(a,a)π(a)vn,h(a,ξ)𝑑a(1un(a,ξ+h)),vn+1,h(0,ξ)=0a+γ(a)vn,h(a,ξ+ca)𝑑a.\begin{cases}\frac{\partial v_{n+1,h}}{\partial a}=J\ast_{\xi}v_{n+1,h}-v_{n+1,h}-Mv_{n+1,h}+Mv_{n,h}+\int_{0}^{a^{+}}K(a,a^{\prime})\pi(a^{\prime})v_{n,h}(a^{\prime},\xi)da^{\prime}\,(1-u^{n}(a,\xi+h)),\\ v_{n+1,h}(0,\xi)=\int_{0}^{a^{+}}\gamma(a)v_{n,h}(a,\xi+ca)da.\end{cases}

It follows from vn,h(a,ξ)un(a,ξ+h)v_{n,h}(a,\xi)\geq u^{n}(a,\xi+h) by (3.10) that

{vn+1,haJξvn+1,hvn+1,hMvn+1,h+Mvn,h+0a+K(,a)π(a)vn,h(a,ξ)𝑑a(1vn,h),vn+1,h(0,ξ)=0a+γ(a)vn,h(a,ξ+ca)𝑑a.\begin{cases}\frac{\partial v_{n+1,h}}{\partial a}\geq J\ast_{\xi}v_{n+1,h}-v_{n+1,h}-Mv_{n+1,h}+Mv_{n,h}+\int_{0}^{a^{+}}K(\cdot,a^{\prime})\pi(a^{\prime})v_{n,h}(a^{\prime},\xi)da^{\prime}\,(1-v_{n,h}),\\ v_{n+1,h}(0,\xi)=\int_{0}^{a^{+}}\gamma(a)v_{n,h}(a,\xi+ca)da.\end{cases}

On the other hand, recalling that un+1u^{n+1} satisfies (3.6) and due to the monotonicity (3.11) again along with vn,hunv_{n,h}\geq u^{n}, we have

Mvn,h(a,ξ)+0a+K(a,a)π(a)vn,h(a,ξ)𝑑a(1vn,h(a,ξ))\displaystyle Mv_{n,h}(a,\xi)+\int_{0}^{a^{+}}K(a,a^{\prime})\pi(a^{\prime})v_{n,h}(a^{\prime},\xi)da^{\prime}\,(1-v_{n,h}(a,\xi))
\displaystyle\geq Mun(a,ξ)+0a+K(a,a)π(a)un(a,ξ)𝑑a(1un(a,ξ)).\displaystyle Mu^{n}(a,\xi)+\int_{0}^{a^{+}}K(a,a^{\prime})\pi(a^{\prime})u^{n}(a^{\prime},\xi)da^{\prime}\,(1-u^{n}(a,\xi)).

As the initial data satisfies,

vn+1,h(0,ξ)\displaystyle v_{n+1,h}(0,\xi) =\displaystyle= 0a+γ(a)vn,h(a,ξ+ca)𝑑a\displaystyle\int_{0}^{a^{+}}\gamma(a)v_{n,h}(a,\xi+ca)da
\displaystyle\geq 0a+γ(a)un(a,ξ+ca)𝑑a=un+1(0,ξ),ξ,\displaystyle\int_{0}^{a^{+}}\gamma(a)u^{n}(a,\xi+ca)da=u^{n+1}(0,\xi),\;\forall\xi\in\mathbb{R},

the comparison principle applies and provides

un+1(a,ξ)vn+1,h(a,ξ),(a,ξ)[0,a+]×.u^{n+1}(a,\xi)\leq v_{n+1,h}(a,\xi),\;\forall(a,\xi)\in[0,a^{+}]\times\mathbb{R}.

Now since for all a[0,a+]a\in[0,a^{+}] the function ξun+1(a,ξ)\xi\to u^{n+1}(a,\xi) is nonincreasing, for all a[0,a+]a\in[0,a^{+}] and ξ\xi\in\mathbb{R}, we get

|un+1(a,ξ+h)un+1(a,ξ)|un+1(a,ξ)un+1(a,ξ+h)\displaystyle|u^{n+1}(a,\xi+h)-u^{n+1}(a,\xi)|\leq u^{n+1}(a,\xi)-u^{n+1}(a,\xi+h) \displaystyle\leq (1emh)vn+1,h(a,ξ)\displaystyle(1-e^{-mh})v_{n+1,h}(a,\xi) (3.14)
\displaystyle\leq emh1.\displaystyle e^{mh}-1.

Hence, we have obtained that, for all n1n\geq 1 and for all h>0h>0,

|un(a,ξ+h)un(a,ξ)|min{1,em|h|1},(a,ξ)[0,a+]×.|u^{n}(a,\xi+h)-u^{n}(a,\xi)|\leq\min\{1,e^{m|h|}-1\},\,\forall(a,\xi)\in[0,a^{+}]\times\mathbb{R}.

As mentioned before, the case h<0h<0 can be handled similarly and thus the result is desired.  

Now the estimates provided in Lemma 3.3 allows us to conclude that the limit ww obtained in Section 3.1 is globally Lipschitz continuous. This complete the desired regularity of ww with respect to ξ\xi as needed in Definition 1.2. Thus we have obtained a continuous and bounded solution ww satisfying (1.9), which is a traveling wave solution of (1.5). This completes the proof of Theorem 1.3 for c>cc>c^{*}.

Next, let us prove the existence of traveling waves for the critical wave speed c=cc=c^{*}. To this aim, assume that {cl}l(c,c+1)\{c_{l}\}_{l\in\mathbb{N}}\in(c^{*},c^{*}+1) is a decreasing sequence satisfying limlcl=c\lim\limits_{l\to\infty}c_{l}=c^{*}. Following the argument after Proposition 3.2, for each clc_{l} there exists a traveling wave solution satisfying (1.9), denoted by wlw_{l}. Since wl(a,+η)w_{l}(a,\cdot+\eta) is also a solution of (1.9) for any η\eta\in\mathbb{R}, we can assume that wl(0,0)=12w_{l}(0,0)=\frac{1}{2}. Recalling Lemma 2.3, for any clc_{l}, there exists λl\lambda_{l} satisfying (2.12). Observe that clλl=λl(cl)c_{l}\to\lambda_{l}=\lambda_{l}(c_{l}) is continuous due to (2.12). It follows that {λl}l\{\lambda_{l}\}_{l\in\mathbb{N}} is bounded due to {cl}l(c,c+1)\{c_{l}\}_{l\in\mathbb{N}}\in(c^{*},c^{*}+1) and thus we can choose m>0m>0 large enough such that m>{λl}lm>\{\lambda_{l}\}_{l\in\mathbb{N}}. Applying the argument in Lemma 3.3 to {wl}l\{w_{l}\}_{l\in\mathbb{N}}, we obtain that {wl}l\{w_{l}\}_{l\in\mathbb{N}} is equi-continuous with respect to ξ\xi. Due to {wl}l[0,1]\{w_{l}\}_{l\in\mathbb{N}}\in[0,1], one see from the equation that {wl}l\{w_{l}\}_{l\in\mathbb{N}} is equi-continuous with respect to aa. Now by Arzelà-Ascoli Theorem, there exists a subsequence of {wl}l\{w_{l}\}_{l\in\mathbb{N}}, again denoted by {wl}l\{w_{l}\}_{l\in\mathbb{N}} satisfying wlww_{l}\to w^{*} as ll\to\infty and that ξw(a,ξ)\xi\to w^{*}(a,\xi) is globally Lipschitz continuous for any a[0,a+]a\in[0,a^{+}] and that aw(a,ξ)W1,1(0,a+)a\to w^{*}(a,\xi)\in W^{1,1}(0,a^{+}) for any ξ\xi\in\mathbb{R}. It is clear that w(0,0)=12w^{*}(0,0)=\frac{1}{2}. In addition, by the same argument in Proposition 3.2, ww^{*} satisfies the following equation

{w(a,ξ)a=(Jξww)(a,ξ)+0a+K(a,a)π(a)w(a,ξ)𝑑a(1w(a,ξ)),w(0,ξ)=0a+γ(a)w(a,ξ+ca)𝑑a.\begin{cases}\frac{\partial w^{*}(a,\xi)}{\partial a}=\left(J\ast_{\xi}w^{*}-w^{*}\right)(a,\xi)+\int_{0}^{a^{+}}K(a,a^{\prime})\pi(a^{\prime})w^{*}(a^{\prime},\xi)da^{\prime}\,(1-w^{*}(a,\xi)),\\ w^{*}(0,\xi)=\int_{0}^{a^{+}}\gamma(a)w^{*}(a,\xi+ca)da.\end{cases} (3.15)

On the other hand, 0w10\leq w^{*}\leq 1 and ξw(a,ξ)\xi\to w^{*}(a,\xi) being nonincreasing for any a[0,a+]a\in[0,a^{+}] imply that the limits w(a,±)=w±(a)w^{*}(a,\pm\infty)=w^{\pm}(a) exist for any a[0,a+]a\in[0,a^{+}]. Moreover, the limits satisfy the equation (2.4), together with the condition

w(0)12w+(0).w^{-}(0)\geq\frac{1}{2}\geq w^{+}(0).

Hence Lemma 2.2 applies and ensures that w(a,)=1w^{*}(a,-\infty)=1 and w(a,)=0w^{*}(a,\infty)=0 uniformly in [0,a+][0,a^{+}]. Hence the existence for c=cc=c^{*} is complete.

Now we set I=π(a)uI=\pi(a)u and then obtain that it is a traveling wave solution of (1.4) with speed cc. Finally, let us notice that π(a+)=0\pi(a^{+})=0 and uu being bounded implies that function II vanishes at a=a+a=a^{+}. Hence Theorem 1.4 is complete.

4 Spreading speeds

In this section, we shall prove Theorem 1.5.

4.1 Outer spreading

For Theorem 1.5-(i), let c>cc>c^{*} be given and fixed. We now look for a super-solution of (1.10) with the form v(t,a,x)=v0eλ(xct)ϕ(a)v(t,a,x)=v_{0}e^{-\lambda(x-ct)}\phi(a) for some positive constant v0v_{0} to be chosen later and λ>0\lambda>0 satisfying (2.12). Here ϕ\phi is defined in (2.10) with s=s0s=s_{0}, where s0s_{0} is defined in Section 2.3. Note that, for all t,a(0,a+)t\in\mathbb{R},a\in(0,a^{+}) and xx\in\mathbb{R} one has

vt+vav0ϕ(a)[J(xy)eλ(yct)𝑑yeλ(xct)]0a+K(a,a)π(a)v(a)𝑑a\displaystyle\frac{\partial v}{\partial t}+\frac{\partial v}{\partial a}-v_{0}\phi(a)\left[\int_{\mathbb{R}}J(x-y)e^{-\lambda(y-ct)}dy-e^{-\lambda(x-ct)}\right]-\int_{0}^{a^{+}}K(a,a^{\prime})\pi(a^{\prime})v(a^{\prime})da^{\prime}
=\displaystyle= cλv+ϕ(a)vϕv[J(xy)eλ(xy)𝑑y1]0a+K(a,a)π(a)v(a)𝑑a\displaystyle c\lambda v+\phi^{\prime}(a)\frac{v}{\phi}-v\left[\int_{\mathbb{R}}J(x-y)e^{\lambda(x-y)}dy-1\right]-\int_{0}^{a^{+}}K(a,a^{\prime})\pi(a^{\prime})v(a^{\prime})da^{\prime}
=\displaystyle= 0.\displaystyle 0.

Besides, again by (2.10) one has

v(t,0,x)=v0eλ(xct)ϕ(0)=v0eλ(xct)0a+γ(a)ϕ(a)𝑑a=0a+γ(a)v(t,a,x)𝑑a.v(t,0,x)=v_{0}e^{-\lambda(x-ct)}\phi(0)=v_{0}e^{-\lambda(x-ct)}\int_{0}^{a^{+}}\gamma(a)\phi(a)da=\int_{0}^{a^{+}}\gamma(a)v(t,a,x)da.

On the other hand, we know from (1.10) that

ut+uaJuu+0a+K(a,a)π(a)u(a)𝑑a.\frac{\partial u}{\partial t}+\frac{\partial u}{\partial a}\leq J\ast u-u+\int_{0}^{a^{+}}K(a,a^{\prime})\pi(a^{\prime})u(a^{\prime})da^{\prime}.

Finally, since u0u_{0} is compactly supported and infa[0,a+]ϕ(a)>0\inf_{a\in[0,a^{+}]}\phi(a)>0, we can choose v0v_{0} large enough such that

v(0,a,x)=v0eλxϕ(a)u0(a,x), for all x and a(0,a+)v(0,a,x)=v_{0}e^{-\lambda x}\phi(a)\geq u_{0}(a,x),\;\text{ for all }x\in\mathbb{R}\text{ and }a\in(0,a^{+})

which implies that vv is a super-solution of (1.10). Hence we end-up with the following estimate due to Lemma 2.1

u(t,a,x)v(t,a,x)=v0eλ(xct)ϕ(a), for all x,t0 and a(0,a+).u(t,a,x)\leq v(t,a,x)=v_{0}e^{-\lambda(x-ct)}\phi(a),\;\text{ for all }x\in\mathbb{R},\,t\geq 0\text{ and }a\in(0,a^{+}).

Now let c1c_{1} be any real number such that c1>c>cc_{1}>c>c^{*}. Then we have

sup|x|c1t, 0<a<a+v(t,a,x)=v0eλ(c1c)tϕ(a)0, as t,\sup_{|x|\geq c_{1}t,\,0<a<a^{+}}v(t,a,x)=v_{0}e^{-\lambda(c_{1}-c)t}\phi(a)\to 0,\text{ as }t\to\infty,

therefore as c>cc>c^{*} can be chosen arbitrary close to cc^{*}, Theorem 1.5-(i) holds true in the case xctx\geq ct and consequently for |x|ct|x|\geq ct as well. This concludes the proof of Theorem 1.5-(i).

4.2 Hair-trigger effects

In order to prove Theorem 1.5-(ii), we first investigate the so-called hair trigger effect of (1.10), that roughly speaking indicates the ability of the solutions of the system to become uniformly positive whatever the smallness of the non-zero and non-negative initial data.

Lemma 4.1

Let u=u(t,a,x)u=u(t,a,x) be the solution of (1.10) with the initial data u0=u0(a,x)u_{0}=u_{0}(a,x). If there exist two constants x0x_{0}\in\mathbb{R} and ρ0(0,1)\rho_{0}\in(0,1) such that

u0(a,x)ρ0 for all x[x01,x0+1] and a(0,a+),u_{0}(a,x)\geq\rho_{0}\text{ for all $x\in[x_{0}-1,x_{0}+1]$ and }a\in(0,a^{+}),

then for any ρ(0,1)\rho\in(0,1), there exists Tρ0ρ0T_{\rho_{0}}^{\rho}\geq 0 which is independent of x0x_{0} such that

u(t,a,x)ρ for all x[x01,x0+1],tTρ0ρ and a(0,a+).u(t,a,x)\geq\rho\text{ for all $x\in[x_{0}-1,x_{0}+1]$},\,t\geq T_{\rho_{0}}^{\rho}\text{ and }a\in(0,a^{+}). (4.1)

Proof. We consider the following auxiliary problem

{wt=Jww+Φminw(1w),w(0,x)=ρ0𝟏[x01,x0+1](x),\begin{cases}\frac{\partial w}{\partial t}=J\ast w-w+\Phi_{\min}w(1-w),\\ w(0,x)=\rho_{0}\mathbf{1}_{[x_{0}-1,x_{0}+1]}(x),\end{cases} (4.2)

where

Φmin:=mina[0,a+]0a+K(a,a)π(a)da>0, due to Assumption 1.1-(ii) and (iv) \Phi_{\min}\mathrel{\mathop{\mathchar 58\relax}}=\min_{a\in[0,a^{+}]}\int_{0}^{a^{+}}K(a,a^{\prime})\pi(a^{\prime})da^{\prime}>0,\text{ due to Assumption \ref{Assump}-(ii) and (iv) } (4.3)

and 𝟏S\mathbf{1}_{S} denotes the indicator function of the set SS. Then it is easy to see that the solution ww of (4.2) is a sub-solution of (1.10) with u0(a,x)ρ0𝟏B1(x0)(x)u_{0}(a,x)\geq\rho_{0}\mathbf{1}_{B_{1}(x_{0})}(x).

However, recall Xu et al. [28, Lemma 4.1] (see also Alfaro [1, Theorem 2.6] or Finkelschtein and Tkachov [15, Theorems 2.5 and 2.7]), the equation (4.2) exhibits the hair trigger effect, that is there exists Tρ0ρ0T_{\rho_{0}}^{\rho}\geq 0 which is independent of x0x_{0} such that

w(t,x)ρ for any x[x01,x0+1],tTρ0ρ.w(t,x)\geq\rho\text{ for any }x\in[x_{0}-1,x_{0}+1],\,t\geq T_{\rho_{0}}^{\rho}.

Now the comparison principles applies to conclude the desired result (4.1).  

Corollary 4.2

Let u=u(t,a,x)u=u(t,a,x) be the solution of (1.10) with non-negative initial data u0=u0(a,x)C([0,a+]×){0}u_{0}=u_{0}(a,x)\in C([0,a^{+}]\times\mathbb{R})\setminus\{0\}. Then one has for each ta+t\geq a^{+} and for any xx\in\mathbb{R},

infa(0,a+)u(t,a,x)>0.\inf_{a\in(0,a^{+})}u(t,a,x)>0.

Due to the above property, Lemma 4.1 implies that, if in addition u01u_{0}\leq 1, then

limtsupa(0,a+)|u(t,a,x)1|=0, locally uniformly for x.\lim_{t\to\infty}\sup_{a\in(0,a^{+})}|u(t,a,x)-1|=0,\text{ locally uniformly for $x\in\mathbb{R}$.}

Proof. First we define an operator 𝒦:L1((0,a+),X)L1((0,a+),X)\mathcal{K}\mathrel{\mathop{\mathchar 58\relax}}L^{1}((0,a^{+}),X)\to L^{1}((0,a^{+}),X) as follows,

𝒦(u):=0a+K(,a)π(a)u(a)da(1u),uL1((0,a+),X).\mathcal{K}(u)\mathrel{\mathop{\mathchar 58\relax}}=\int_{0}^{a^{+}}K(\cdot,a^{\prime})\pi(a^{\prime})u(a^{\prime})da^{\prime}\;(1-u),\quad u\in L^{1}((0,a^{+}),X).

Next solving the problem (1.10) along the characteristic line at=ca-t=c, where cc\in\mathbb{R}, we now derive the formula for a solution to (1.1). For fixed cc\in\mathbb{R}, we set w(t)=u(t,t+c)w(t)=u(t,t+c) for t[max(c,0),)t\in[\max(-c,0),\infty). With a=t+ca=t+c one obtains for t[max(c,0),)t\in[\max(-c,0),\infty) the equation

tw(t)=Tww+[𝒦(w(t))](t+c),\partial_{t}w(t)=Tw-w+[\mathcal{K}(w(t))](t+c), (4.4)

where TT is defined in (2.1). We first study the case c0c\geq 0. Clearly, w(0)=u(0,c)=u(0,at)=u0(at)w(0)=u(0,c)=u(0,a-t)=u_{0}(a-t). Considering the equation (4.4) with initial data w(0)0w(0)\geq 0 and w(0)0w(0)\not\equiv 0, we have w(t)>0w(t)>0 for t>0t>0 by the strong comparison principle of the nonlinear nonlocal diffusion problem, due to J(0)>0J(0)>0 in Assumption 1.1-(iii). It follows that u(t,a)>0u(t,a)>0 for ata\geq t. On the other hand, integrating (4.4) from 0 to tt, one obtains

w(t)=e(TI)tw(0)+0te(TI)(ts)[𝒦(w(s))](s+c)𝑑s,w(t)=e^{(T-I)t}w(0)+\int_{0}^{t}e^{(T-I)(t-s)}[\mathcal{K}(w(s))](s+c)ds,

and thus

u(t,a)=e(TI)tu0(at)+0te(TI)(ts)[𝒦(u(s))](s+at)𝑑s.u(t,a)=e^{(T-I)t}u_{0}(a-t)+\int_{0}^{t}e^{(T-I)(t-s)}[\mathcal{K}(u(s))](s+a-t)ds.

Next we consider the case c<0c<0. Integrating (4.4) from c-c to tt, one gets

w(t)=e(TI)(t+c)w(c)+cte(TI)(ts)[𝒦(u(s))](s+c)𝑑s,w(t)=e^{(T-I)(t+c)}w(-c)+\int_{-c}^{t}e^{(T-I)(t-s)}[\mathcal{K}(u(s))](s+c)ds,

and thus

u(t,a)=e(TI)au(ta,0)+tate(TI)(ts)[𝒦(u(s))](s+at)𝑑s.u(t,a)=e^{(T-I)a}u(t-a,0)+\int_{t-a}^{t}e^{(T-I)(t-s)}[\mathcal{K}(u(s))](s+a-t)ds.

Thus now the solution to (1.1) reads as follows:

u(t,a)={e(TI)tu0(at)+0te(TI)(ts)[𝒦(u(s))](s+at)𝑑s,at,e(TI)au(ta,0)+tate(TI)(ts)[𝒦(u(s))](s+at)𝑑s,a<t.u(t,a)=\begin{cases}e^{(T-I)t}u_{0}(a-t)+\int_{0}^{t}e^{(T-I)(t-s)}[\mathcal{K}(u(s))](s+a-t)ds,&a\geq t,\\ e^{(T-I)a}u(t-a,0)+\int_{t-a}^{t}e^{(T-I)(t-s)}[\mathcal{K}(u(s))](s+a-t)ds,&a<t.\end{cases} (4.5)

Next we plug the explicit formula (4.5) into u(t,0)u(t,0) to obtain

u(t,0)\displaystyle u(t,0) =\displaystyle= 0tχ(a)γ(a)[e(TI)au(ta,0)+tate(TI)(ts)[𝒦(u(s))](s+at)𝑑s]𝑑a\displaystyle\int_{0}^{t}\chi(a)\gamma(a)\left[e^{(T-I)a}u(t-a,0)+\int_{t-a}^{t}e^{(T-I)(t-s)}[\mathcal{K}(u(s))](s+a-t)ds\right]da (4.6)
+ta+χ(a)γ(a)[e(TI)tu0(at)+0te(TI)(ts)[𝒦(u(s))](s+at)𝑑s]𝑑a,\displaystyle+\int_{t}^{a^{+}}\chi(a)\gamma(a)\left[e^{(T-I)t}u_{0}(a-t)+\int_{0}^{t}e^{(T-I)(t-s)}[\mathcal{K}(u(s))](s+a-t)ds\right]da,

where χ(a)\chi(a) is the cutoff function satisfying χ(a)=1\chi(a)=1 when a(0,a+)a\in(0,a^{+}) otherwise χ(a)=0\chi(a)=0. Now we consider two cases.

Case 1. If t<a+t<a^{+}, (4.6) is written as follows:

u(t,0)\displaystyle u(t,0) =\displaystyle= 0tγ(a)[e(TI)au(ta,0)+tate(TI)(ts)[𝒦(u(s))](s+at)𝑑s]𝑑a\displaystyle\int_{0}^{t}\gamma(a)\left[e^{(T-I)a}u(t-a,0)+\int_{t-a}^{t}e^{(T-I)(t-s)}[\mathcal{K}(u(s))](s+a-t)ds\right]da (4.7)
+ta+γ(a)[e(TI)tu0(at)+0te(TI)(ts)[𝒦(u(s))](s+at)𝑑s]𝑑a.\displaystyle+\int_{t}^{a^{+}}\gamma(a)\left[e^{(T-I)t}u_{0}(a-t)+\int_{0}^{t}e^{(T-I)(t-s)}[\mathcal{K}(u(s))](s+a-t)ds\right]da.

Since e(TI)tu0(at)>0e^{(T-I)t}u_{0}(a-t)>0 for ata\geq t and γ0\gamma\geq 0 for any a[0,a+]a\in[0,a^{+}] by Assumption 1.1-(i), the second term on the right hand of (4.7) must be positive. Thus we have u(t,0)>0u(t,0)>0 which implies u(t,a)>0u(t,a)>0 for a<ta<t via (4.5).

Case 2. If ta+t\geq a^{+}, (4.6) is written as follows:

u(t,0)=0a+γ(a)[e(TI)au(ta,0)+tate(TI)(ts)[𝒦(u(s))](s+at)𝑑s]𝑑a.u(t,0)=\int_{0}^{a^{+}}\gamma(a)\left[e^{(T-I)a}u(t-a,0)+\int_{t-a}^{t}e^{(T-I)(t-s)}[\mathcal{K}(u(s))](s+a-t)ds\right]da. (4.8)

Let us claim that u(t,0,x):=[u(t,0)](x)>0u(t,0,x)\mathrel{\mathop{\mathchar 58\relax}}=[u(t,0)](x)>0 in [a+,)×[a^{+},\infty)\times\mathbb{R}. By contradiction, suppose that there exists (t0,x0)[a+,)×(t_{0},x_{0})\in[a^{+},\infty)\times\mathbb{R} such that u(t0,0,x0)=0u(t_{0},0,x_{0})=0. Thus one obtains

00a+γ(a)eaeTau(t0a,0,x0)𝑑a,0\geq\int_{0}^{a^{+}}\gamma(a)e^{-a}e^{Ta}u(t_{0}-a,0,x_{0})da,

where we used the fact that eae^{-a} and eTae^{Ta} are commuting. By Assumption 1.1-(i) on γ\gamma, we can find one point b0(0,a+]b_{0}\in(0,a^{+}] such that eTau(t0b0,0,x0)=0e^{Ta}u(t_{0}-b_{0},0,x_{0})=0. By definition, one has

eTau(t0b0,0,x0)=n=0(a)nn!Jnu(t0b0,0,x0),e^{Ta}u(t_{0}-b_{0},0,x_{0})=\sum_{n=0}^{\infty}\frac{(a)^{n}}{n!}J^{*n}\ast u(t_{0}-b_{0},0,x_{0}),

where JnJ^{*n} denotes the nn-fold convolution of KK; that is Jn=JJJ^{*n}=J\ast\cdots\ast J, nn times. It follows that for each nn\in\mathbb{N},

Jnu(t0b0,0,x0)=0.J^{*n}\ast u(t_{0}-b_{0},0,x_{0})=0.

However, by Assumption 1.1-(iii) on JJ, one has J>0J>0 in (r,r)(-r,r) for some r>0r>0, which implies that

u(t0b0,0,x)=0, for all x(x0nr,x0+nr).u(t_{0}-b_{0},0,x)=0,\text{ for all }x\in(x_{0}-nr,x_{0}+nr).

Since =n=0(x0nr,x0+nr)\mathbb{R}=\sum_{n=0}^{\infty}(x_{0}-nr,x_{0}+nr), we have u(t0b0,0,)0u(t_{0}-b_{0},0,\cdot)\equiv 0 in \mathbb{R}.

Next replace t0t_{0} by t0b0t_{0}-b_{0} in (4.6). If t0b0t_{0}-b_{0} falls in [0,a+][0,a^{+}], by the argument as Case 1, one has u(t0b0,0)>0u(t_{0}-b_{0},0)>0, which is a contradiction. Hence, t0b0t_{0}-b_{0} must fall in (a+,)(a^{+},\infty). Then by the same argument as Case 2, one can find b1(0,a+]b_{1}\in(0,a^{+}] such that u(t0b0b1,0)=0u(t_{0}-b_{0}-b_{1},0)=0. Now repeating the above process by induction, one can find a sequence {bi}i0\{b_{i}\}_{i\geq 0} such that u(t0i=0M^bi,0)=0u(t_{0}-\sum_{i=0}^{\widehat{M}}b_{i},0)=0 for any M^0\widehat{M}\geq 0. But we know every bib_{i} is in (0,a+](0,a^{+}], then there always exists a minimal M0>0M_{0}>0 such that t0i=0M0bi<a+t_{0}-\sum_{i=0}^{M_{0}}b_{i}<a^{+}. Then by Case 1, one has u(t0i=0M0bi,0)>0u(t_{0}-\sum_{i=0}^{M_{0}}b_{i},0)>0.

Now consider (4.8) at t=t0i=0M01bit=t_{0}-\sum_{i=0}^{M_{0}-1}b_{i}, which is larger than or equal to a+a^{+}, we get a contradiction, since now the left hand side of (4.8) equals to zero, while the right hand side of (4.8) is larger than zero.

In summary, we cannot have (t,x)(0,)×(t,x)\in(0,\infty)\times\mathbb{R} such that u(t,0,x)=0u(t,0,x)=0, which implies that u(t,0,x)>0u(t,0,x)>0 and thus u(t,a)>0u(t,a)>0 by (4.5). Hence, the proof is complete.  

4.3 Inner spreading

In this section, we prove the inner spreading. To this aim, we consider the following auxiliary equation

{vt=Jvv+λ0v(1Pv),v(0,x)=v0(x),\begin{cases}\frac{\partial v}{\partial t}=J\ast v-v+\lambda_{0}v(1-Pv),\\ v(0,x)=v_{0}(x),\end{cases} (4.9)

where v0v_{0} denotes a Lipschitz continuous in \mathbb{R} with compact support satisfying

v0(x)1P in ,v_{0}(x)\leq\frac{1}{P}\text{ in }\mathbb{R},

and, recalling the definition of the function ϕ\phi in (2.10) with s=s0<0s=s_{0}<0, PP and λ0\lambda_{0} are the positive constants given by

P=supa(0,a+)0a+K(a,a)π(a)ϕ(a)𝑑aλ0>0 and λ0:=s0>0 (see the proof of Lemma 2.3P=\frac{\sup_{a\in(0,a^{+})}\int_{0}^{a^{+}}K(a,a^{\prime})\pi(a^{\prime})\phi(a^{\prime})da^{\prime}}{\lambda_{0}}>0\text{ and }\lambda_{0}\mathrel{\mathop{\mathchar 58\relax}}=-s_{0}>0\text{ (see the proof of Lemma \ref{speed}) }

Recall that u0=u0(a,x)C0([0,a+]×){0}u_{0}=u_{0}(a,x)\in C^{0}([0,a^{+}]\times\mathbb{R})\setminus\{0\} is non-negative, u01u_{0}\leq 1 and supp(u0){\rm supp}(u_{0}) is compact in [0,a+]×[0,a^{+}]\times\mathbb{R}. Hence according to Corollary 4.2, one has there exists ρ>0\rho>0 such that

u~0(a,x):=u(a+,a,x)ρ,a[0,a+],x[R,R].\tilde{u}_{0}(a,x)\mathrel{\mathop{\mathchar 58\relax}}=u(a^{+},a,x)\geq\rho,\;\;\forall a\in[0,a^{+}],\;x\in[-R,R].

Next we fix v0:+v_{0}\mathrel{\mathop{\mathchar 58\relax}}\mathbb{R}\to\mathbb{R}^{+} Lipschitz continuous such that

supp(v0)[R,R],supxv0(x)1P and (supa[0,a+]ϕ(a))(supxv0(x))ρ.{\rm supp}\,(v_{0})\subset[-R,R],\;\;\sup_{x\in\mathbb{R}}v_{0}(x)\leq\frac{1}{P}\text{ and }\left(\sup_{a\in[0,a^{+}]}\phi(a)\right)\left(\sup_{x\in\mathbb{R}}v_{0}(x)\right)\leq\rho.

so that the function u¯0C([0,a+]×)\underline{\textbf{u}}_{0}\in C([0,a^{+}]\times\mathbb{R}) given by

u¯0(a,x):=ϕ(a)v0(x),(a,x)[0,a+]×,\underline{\textbf{u}}_{0}(a,x)\mathrel{\mathop{\mathchar 58\relax}}=\phi(a)v_{0}(x),\;\forall(a,x)\in[0,a^{+}]\times\mathbb{R},

satisfies

u¯0(a,x)u(a+,a,x),(a,x)[0,a+]×.\underline{\textbf{u}}_{0}(a,x)\leq u(a^{+},a,x),\;\;\forall(a,x)\in[0,a^{+}]\times\mathbb{R}.

Next set u¯(t,a,x)=ϕ(a)v(t,x)\underline{\textbf{u}}(t,a,x)=\phi(a)v(t,x), where v=v(t,x)v=v(t,x) is the solution of (4.9) with the initial data v0v_{0}. Next, we verify that u¯\underline{\textbf{u}} is a sub-solution of (1.10). Indeed, we have

u¯t+u¯a\displaystyle\frac{\partial\underline{\textbf{u}}}{\partial t}+\frac{\partial\underline{\textbf{u}}}{\partial a} =\displaystyle= Ju¯u¯+λ0u¯(1Pv)+s0u¯+v0a+K(a,a)π(a)ϕ(a)𝑑a\displaystyle J\ast\underline{\textbf{u}}-\underline{\textbf{u}}+\lambda_{0}\underline{\textbf{u}}(1-Pv)+s_{0}\underline{\textbf{u}}+v\int_{0}^{a^{+}}K(a,a^{\prime})\pi(a^{\prime})\phi(a^{\prime})da^{\prime}\, (4.10)
=\displaystyle= Ju¯u¯+v0a+K(a,a)π(a)ϕ(a)𝑑aPλ0u¯v\displaystyle J\ast\underline{\textbf{u}}-\underline{\textbf{u}}+v\int_{0}^{a^{+}}K(a,a^{\prime})\pi(a^{\prime})\phi(a^{\prime})da^{\prime}-P\lambda_{0}\underline{\textbf{u}}v
\displaystyle\leq Ju¯u¯+v0a+K(a,a)π(a)ϕ(a)𝑑a(1u¯),\displaystyle J\ast\underline{\textbf{u}}-\underline{\textbf{u}}+v\int_{0}^{a^{+}}K(a,a^{\prime})\pi(a^{\prime})\phi(a^{\prime})da^{\prime}\,(1-\underline{\textbf{u}}),

and by (2.8) and the choice of v0v_{0}, one has

u¯(t,0,x)=ϕ(0)v(t,x)=0a+γ(a)ϕ(a)𝑑av(t,x)=0a+γ(a)u¯(t,a,x)𝑑a,\displaystyle\underline{\textbf{u}}(t,0,x)=\phi(0)v(t,x)=\int_{0}^{a^{+}}\gamma(a)\phi(a)da\,v(t,x)=\int_{0}^{a^{+}}\gamma(a)\underline{\textbf{u}}(t,a,x)da,
u¯(0,a,x)=ϕ(a)v0(x)u0(a,x).\displaystyle\underline{\textbf{u}}(0,a,x)=\phi(a)v_{0}(x)\leq u_{0}(a,x).

Now define

c0:=infλ>0J(y)eλy𝑑y1+λ0λ.c_{0}\mathrel{\mathop{\mathchar 58\relax}}=\inf_{\lambda>0}\frac{\int_{\mathbb{R}}J(y)e^{\lambda y}dy-1+\lambda_{0}}{\lambda}.

Note that c0>0c_{0}>0 since JJ is symmetric. Then let us recall a spreading result for equation (4.9), from Ducrot and Jin [8, Lemma 3.6].

Lemma 4.3

let v=v(t,x)v=v(t,x) be the solution of (4.9) supplemented with a continuous initial data 0v0()1P0\leq v_{0}(\cdot)\leq\frac{1}{P} and v00v_{0}\not\equiv 0 with compact support. Let us further assume that vv is uniformly continuous for all t0,xt\geq 0,x\in\mathbb{R}. Then one has

limtsup|x|ct|v(t,x)1P|=0,0<c<c0.\lim_{t\to\infty}\sup_{|x|\leq ct}\left|v(t,x)-\frac{1}{P}\right|=0,\;\forall 0<c<c_{0}.

Next, let us show c0=cc_{0}=c^{*}. Define the function H(λ),λ>0H(\lambda),\lambda>0 as follows,

H(λ):=J(y)eλy𝑑y1+λ0λ.H(\lambda)\mathrel{\mathop{\mathchar 58\relax}}=\frac{\int_{\mathbb{R}}J(y)e^{\lambda y}dy-1+\lambda_{0}}{\lambda}.

Differentiating HH with respect to λ\lambda and setting λ=λ(c)\lambda^{*}=\lambda(c^{*}), one obtains by (2.13)

H(λ)=J(y)eλy(λy1)𝑑yJ(y)eλy(λy1)𝑑yλ2.H^{\prime}(\lambda)=\frac{\int_{\mathbb{R}}J(y)e^{\lambda y}(\lambda y-1)dy-\int_{\mathbb{R}}J(y)e^{\lambda^{*}y}(\lambda^{*}y-1)dy}{\lambda^{2}}.

Next, when H(λ)=0H^{\prime}(\lambda)=0, one obtains that λ\lambda satisfies

J(y)eλy(λy1)𝑑y=J(y)eλy(λy1)𝑑y,\int_{\mathbb{R}}J(y)e^{\lambda y}(\lambda y-1)dy=\int_{\mathbb{R}}J(y)e^{\lambda^{*}y}(\lambda^{*}y-1)dy,

which implies λ=λ\lambda=\lambda^{*} due to the monotonicity of J(y)eλy(λy1)𝑑y\int_{\mathbb{R}}J(y)e^{\lambda y}(\lambda y-1)dy with respect to λ\lambda and thus c0=cc_{0}=c^{*}. Hence now for any 0c<c0\leq c<c^{*} we have

limtinf|x|ct,0<a<a+u(t+a+,a,x)limtinf|x|ct,0<a<a+ϕ(a)v(t,x)mina[0,a+]ϕ(a)P=:ρ0>0.\lim_{t\to\infty}\inf_{|x|\leq ct,0<a<a^{+}}u(t+a^{+},a,x)\geq\lim_{t\to\infty}\inf_{|x|\leq ct,0<a<a^{+}}\phi(a)v(t,x)\geq\frac{\min_{a\in[0,a^{+}]}\phi(a)}{P}=\mathrel{\mathop{\mathchar 58\relax}}\rho_{0}>0. (4.11)

Moreover, one may assume that ρ0<1\rho_{0}<1. Thus (4.11) implies that for any c[0,c)c\in[0,c^{*}), there exists T>0T>0 such that

u(t,a,x)ρ02 for tT,|x|ct and a(0,a+).u(t,a,x)\geq\frac{\rho_{0}}{2}\text{ for }t\geq T,\,|x|\leq ct\,\text{ and }a\in(0,a^{+}).

Now for any ρ(0,1)\rho\in(0,1), applying Lemma 4.1 to (1.10) yields that there exists Tρ0ρ>0T_{\rho_{0}}^{\rho}>0 such that

u(t+Tρ0ρ,a,x)ρ, for tT,|x|ct and a(0,a+),u(t+T_{\rho_{0}}^{\rho},a,x)\geq\rho,\text{ for }t\geq T,\,|x|\leq ct\,\text{ and }a\in(0,a^{+}),

which implies that

inf|x|ctcTρ0ρ,0<a<a+u(t,a,x)ρ, for tT+Tρ0ρ.\inf_{|x|\leq ct-cT_{\rho_{0}}^{\rho},0<a<a^{+}}u(t,a,x)\geq\rho,\text{ for }t\geq T+T_{\rho_{0}}^{\rho}.

For any ϵ(0,c)\epsilon\in(0,c), there exists a constant TTρ0ρT^{\prime}\geq T_{\rho_{0}}^{\rho} such that ϵTcTρ0ρ\epsilon T^{\prime}\geq cT_{\rho_{0}}^{\rho}. Then we have that ctcTρ0ρ(cϵ)tct-cT_{\rho_{0}}^{\rho}\geq(c-\epsilon)t and

inf|x|(cϵ)t,0<a<a+u(t,a,x)ρ, for tT.\inf_{|x|\leq(c-\epsilon)t,0<a<a^{+}}u(t,a,x)\geq\rho,\text{ for }t\geq T^{\prime}.

Now since ρ(0,1)\rho\in(0,1) is arbitrary close to 11, we obtain

limtinf|x|(cϵ)t,0<a<a+u(t,a,x)=1.\lim_{t\to\infty}\inf_{|x|\leq(c-\epsilon)t,0<a<a^{+}}u(t,a,x)=1.

Due to the arbitrariness of ϵ\epsilon, hence we have the following result.

Theorem 4.4

let u=u(t,a,x)u=u(t,a,x) be the solution of (1.10) supplemented with a continuous initial data 0u010\leq u_{0}\leq 1 and u00u_{0}\not\equiv 0 with u0u_{0} being compactly supported in [0,a+]×[0,a^{+}]\times\mathbb{R}, then one has

limtsup|x|ct,0<a<a+|u(t,a,x)1|=0, for all c(0,c).\lim_{t\to\infty}\sup_{|x|\leq ct,0<a<a^{+}}\left|u(t,a,x)-1\right|=0,\;\text{ for all }c\in(0,c^{*}).

This proves Theorem 1.5-(ii).

References

  • [1] M. Alfaro. Fujita blow up phenomena and hair trigger effect: the role of dispersal tails. Ann. Inst. H. Poincaré Anal. Non Linéaire, 34(5):1309–1327, 2017.
  • [2] J. Coville, J. Dávila, and S. Martínez. Existence and uniqueness of solutions to a nonlocal equation with monostable nonlinearity. SIAM J. Math. Anal., 39(5):1693–1709, 2008.
  • [3] J. Coville, J. Dávila, and S. Martínez. Nonlocal anisotropic dispersal with monostable nonlinearity. J. Differential Equations, 244(12):3080–3118, 2008.
  • [4] J. Coville and L. Dupaigne. Propagation speed of travelling fronts in non local reaction–diffusion equations. Nonlinear Anal. Theory, Methods & Applications, 60(5):797–819, 2005.
  • [5] J. Coville and L. Dupaigne. On a non-local equation arising in population dynamics. Proc. Roy. Soc. Edinburgh Sect. A, 137(4):727–755, 2007.
  • [6] A. Ducrot. Travelling wave solutions for a scalar age-structured equation. Discrete Contin. Dyn. Syst. Ser. B, 7(2):251, 2007.
  • [7] A. Ducrot and Z. Jin. Generalized travelling fronts for non-autonomous Fisher-KPP equations with nonlocal diffusion. Ann. Mat. Pura Appl. (4), pages 1–32, 2021.
  • [8] A. Ducrot and Z. Jin. Spreading properties for non-autonomous Fisher–KPP equations with non-local diffusion. J. Nonlinear Sci., 33(6):100, 2023.
  • [9] A. Ducrot, H. Kang, and S. Ruan. Age-structured models with nonlocal diffusion, Part I: principal spectral theory, limiting properties. J. Anal. Math., to appear.
  • [10] A. Ducrot, H. Kang, and S. Ruan. Age-structured models with nonlocal diffusion, Part II: global dynamics. Israel J. Math., to appear.
  • [11] A. Ducrot and P. Magal. Travelling wave solutions for an infection-age structured model with diffusion. Proc. Roy. Soc. Edinburgh Sect. A, 139(3):459–482, 2009.
  • [12] A. Ducrot and P. Magal. Travelling wave solutions for an infection-age structured epidemic model with external supplies. Nonlinearity, 24(10):2891, 2011.
  • [13] A. Ducrot, P. Magal, and S. Ruan. Travelling wave solutions in multigroup age-structured epidemic models. Arch. Ration. Mech. Anal., 195(1):311–331, 2010.
  • [14] J. Fang and X.-Q. Zhao. Traveling waves for monotone semiflows with weak compactness. SIAM J. Math. Anal., 46(6):3678–3704, 2014.
  • [15] D. Finkelshtein and P. Tkachov. The hair-trigger effect for a class of nonlocal nonlinear equations. Nonlinearity, 31(6):2442, 2018.
  • [16] H. Kang and S. Ruan. Mathematical analysis on an age-structured SIS epidemic model with nonlocal diffusion. J. Math. Biol., 83(1):1–30, 2021.
  • [17] W.-T. Li, Y.-J. Sun, and Z.-C. Wang. Entire solutions in the Fisher-KPP equation with nonlocal dispersal. Nonlinear Anal. Real World Appl., 11(4):2302–2313, 2010.
  • [18] P. Magal and S. Ruan. Theory and Applications of Abstract Semilinear Cauchy Problems. Springer, New York, 2018.
  • [19] P. Magal, O. Seydi, and F.-B. Wang. Monotone abstract non-densely defined cauchy problems applied to age structured population dynamic models. J. Math. Anal. Appl., 479(1):450–481, 2019.
  • [20] I. Marek. Frobenius theory of positive operators: Comparison theorems and applications. SIAM J. Appl. Math., 19(3):607–628, 1970.
  • [21] I. Sawashima. On spectral properties of some positive operators. Nat. Sci. Rep., Ochanomizu University, 15(2):53–64, 1964.
  • [22] W. Shen and Z. Shen. Transition fronts in nonlocal Fisher-KPP equations in time heterogeneous media. Commun. Pure Appl. Anal., 15(4):1193–1213, 2016.
  • [23] W. Shen and A. Zhang. Spreading speeds for monostable equations with nonlocal dispersal in space periodic habitats. J. Differential Equations, 249(4):747–795, 2010.
  • [24] Y.-J. Sun, W.-T. Li, and Z.-C. Wang. Traveling waves for a nonlocal anisotropic dispersal equation with monostable nonlinearity. Nonlinear Anal. Theory, Methods & Applications, 74(3):814–826, 2011.
  • [25] H. R. Thieme. Positive perturbation of operator semigroups: growth bounds, essential compactness and asynchronous exponential growth. Discrete Contin. Dyn. Syst. A, 4(4):735, 1998.
  • [26] H. R. Thieme. Spectral bound and reproduction number for infinite-dimensional population structure and time heterogeneity. SIAM J. Appl. Math., 70(1):188–211, 2009.
  • [27] G. F. Webb. Theory of Nonlinear Age-Dependent Population Dynamics. Marcel Dekker, New York, 1984.
  • [28] W.-B. Xu, W.-T. Li, and S. Ruan. Spatial propagation in nonlocal dispersal Fisher-KPP equations. J. Funct. Anal., page 108957, 2021.