1 Introduction

It is well known that malaria is a parasitic infections in humans, caused by the genus Plasmodium. In general, humans acquire malaria through effective biting by several species of infectious female anopheles mosquitoes [23, 25]. Susceptible mosquitoes acquire malaria through effective bites of infectious human host. It has been reported in [15] that about more than one hundred countries are under prevalence of malaria. Every year, two billion people are at risk affected by Plasmodium falciparum.

Reaction–diffusion model frameworks have been proved to be a powerful tool to generalize the classical Ross–Macdonald malaria models [4, 6, 11, 20,21,22, 32, 34]. For a spatial transmission dynamics of malaria, it is usually assumed that human and mosquitoes are confined in a bounded domain \(\Omega \). Laplacian operator \(\partial /\partial x^2\), \(x\in \Omega \subset {\mathbb {R}}^n (n\ge 1)\) are introduced to reflect the spatial random movement of humans and mosquitoes. Recent publications [13, 14, 21] demonstrated that the spatial heterogeneity is a more meaningful and important factor in disease transmission. In reality, as environmental conditions vary spatially, for example, temperature and humidity, etc., it comes natural to demonstrate spatial heterogeneity for disease transmission parameters. It is also argued in [26] that “the non-random distribution of humans and mosquitoes across the landscape can generate spatially heterogeneous biting patterns”.

In typical and pioneering work of malaria transmission model [1, 18,19,20, 23, 25], humans are categorized into susceptible and infected class. The adult female mosquitoes (termed as vector) are categorized into susceptible and infected mosquito classes. The mosquitoes have to breed in water, and the adult female mosquitoes emerge from aquatic mosquitoes. The adult female mosquitoes transmit malaria to human hosts. The infected mosquitoes may spend a incubation period of 10 to 14 days (termed as extrinsic incubation period (EIP)) to survive (during which the infected mosquitoes cannot transmit malaria to human hosts) and transmit malaria to human hosts [18]. Taking EIP and mobility of human and mosquitoes in a spatial domain into account, Lou and Zhao [21] generalized the model in [25] and proposed a nonlocal and time-delayed diffusive malaria model, for \(x\in \Omega ,\ t>0\),

$$\begin{aligned} \left\{ \begin{array}{ll} \displaystyle \frac{\partial S_m(t,x)}{\partial t} -D_m\Delta S_m(t,x)=\mu (x)-\frac{b\beta (x)}{H(x)}S_m(t,x)I_h(t,x)-d_mS_m(t,x), \\ \displaystyle \frac{\partial I_m(t,x)}{\partial t} -D_m\Delta I_m(t,x)=e^{-d_m\tau }\int \limits _{\Omega }\Gamma (\tau ,x,y)\frac{b\beta (y)}{H(y)}S_m(t-\tau ,y)I_h(t-\tau ,y)\mathrm{d}y-d_mI_m(t,x), \\ \displaystyle \frac{\partial I_h(t,x)}{\partial t} - D_h\Delta I_h(t,x) =\frac{c\beta (x)}{H(x)}(H(x)-I_h)I_m(t,x)-(d_h+\rho ) I_h(t,x), \end{array}\right. \end{aligned}$$
(1.1)

with

$$\begin{aligned} \frac{\partial S_m(t,x)}{\partial n}=\frac{\partial I_m(t,x)}{\partial n}=\frac{\partial I_h(t,x)}{\partial n}=0,\ x\in \partial \Omega , t>0. \end{aligned}$$

Here, \(S_m(t,x)\), \(I_m(t,x)\) and \(I_h(t,x)\) denote the density of susceptible, infected adult female mosquitoes and infected humans at space x and time t, respectively, which are equipped with diffusion rates \(D_m\), \(D_m\) and \(D_h\), respectively. The total population stabilizes at H(x). \(\beta (x)\) and \(\mu (x)\) stand for the space-dependent biting rate and recruitment rate of adult female mosquitoes, respectively. \(d_m\) represents the natural death rate of mosquitoes. b and c stand for the transmission probabilities per bite from \(I_h\) to \(S_m\) and from \(I_m\) to \(H(x)-I_h\), respectively. \(d_h\) and \(\rho \) represent the death rate and recovery rate of humans. \(\Gamma \) is the Green function corresponding to the operator \(D_m\Delta \) subject to the Neumann boundary condition. \(\tau \) represents the fixed incubation period constant.

Since the latent mosquitoes in one location can fly around during EIP, and arrive at any location in the domain when they can transmit malaria to human hosts, the spatial movement of mosquitoes in EIP will result in non-local infection [3, 13, 14]. In model (1.1), the nonlocal time delay term is obtained by introducing an infection age variable a. At time t and space x, denote by \(i_m(t,a,x)\) the density of the mosquitoes with infection age a. Suppose that \(\tau \) is a fixed latent period. It then follows that \(E_m(t,x)=\int \limits _0^{\tau }i_m(t,a,x)\mathrm{d}a\) and \(I_m(t,x)=\int \limits _{\tau }^{\infty }i_m(t,a,x)\mathrm{d}a\) are the density of latent and infected mosquitoes, respectively. Then, the evolution of \(i_m(t,a,x)\) satisfies

$$\begin{aligned} \left\{ \begin{array}{ll} \displaystyle \frac{\partial i_m(t,a,x)}{\partial t} +\frac{\partial i_m(t,a,x)}{\partial a} - D_m\Delta i_m(t, a, x)=-d_m i_m(t, a, x),&{}\quad x\in \Omega ,\ t>0, a\ge 0, \\ \displaystyle i_m(t,0,x)=\frac{b\beta (x)}{H(x)}S_mI_h,&{}\quad x\in \Omega ,\ t>0,\\ \displaystyle \frac{\partial i_m(t,a,x)}{\partial n}=0,\ &{}\quad x\in \partial \Omega ,\ a\ge 0, \end{array}\right. \end{aligned}$$
(1.2)

where \(i_m(t,0,x)\) is the newly infected mosquitoes. The nonlocal time delay term in (1.1) comes from the key point that \(i_m(t,\tau ,x)\) can be determined by the integration along characteristics line \(t-a=const.\),

$$\begin{aligned} i_m(t,a,x)=\left\{ \begin{array}{ll} \displaystyle \int \limits _\Omega \Gamma _1(a, x, y)i_m(t-a,0,y)\mathrm{d}y\Pi (a),&{}\quad t\ge a, \\ \displaystyle \int \limits _\Omega \Gamma _1(t, x, y)i_{m0}(a-t,y)\mathrm{d}y\frac{\Pi (a)}{\Pi (a-t)},&{}\quad t< a, \end{array}\right. \end{aligned}$$
(1.3)

where \(i_{m0}\) is initial condition, and \(\Pi (a)=e^{-d_ma}\). Hence,

$$\begin{aligned} i_m(t,\tau ,x)=\Pi (\tau )\int \limits _{\Omega }\Gamma _1(\tau ,x,y)i_m(t-\tau ,0,y)\mathrm{d}y,\ \forall t\ge \tau . \end{aligned}$$

From the standpoint of mathematical analysis, the authors in [21] confirmed that the threshold dynamics of (1.1) is governed by the sign of the principle eigenvalue of the associated linear and nonlocal eigenvalue problem. Subsequently, a typical feature of malaria termed as vector-bias, was introduced to describe the difference between the probability (denote by p and l) of a mosquito picking the human when he/she is infected and susceptible (see, e.g., [8, 29, 33, 34]). To explore the seasonal patterns of malaria epidemics (may be caused by annual temperature and rainfall variation (see, e.g., [12])), Bai et al. [4] further extended the models in [21, 33] by incorporating the seasonality, vector-bias, EIP and the spatial heterogeneity, and formulated a time-delayed periodic reaction-diffusion model. Their results suggest that the sharp threshold results can be achieved by defining the basic reproduction number. They also confirmed the existence of a periodic solution for the proposed models, which built an interesting biological implication: Seasonal patterns of malaria epidemics will occur. It is important to mention that recent studies on diffusive Zika epidemic models (see, e.g., [11, 22, 32]) can also be regarded as a generalization of the classical model in [25]. Like other vector-borne disease models, Zika epidemic models share the same cross-infection mechanism between humans and mosquitoes.

Our goal of this paper is to perform an original analysis of (1.1). A more complete understanding of (1.1) can help to get better understanding of the malaria transmission in regions. As in studies of Zika outbreak in Rio De Janerio, Bastos et al. [28] argued that compared to the total population, infected human density takes a fairly small number [28]. Subsequently, Fitzgibbon et al. [11], Magal et al. [22] and Wang and Chen [32] proposed the diffusive vector-borne disease model by assuming the density of susceptible humans to be H(x), which depends only on space x, and will not be altered within a short time. On the other hand, in (1.1), the force of infection for humans and mosquitoes is described by adopting standard infection mechanism, \(\frac{b\beta (x)}{H(x)}S_mI_h\) and \(\frac{c\beta (x)}{H(x)}(H(x)-I_h)I_m\), respectively, where \(H(x)-I_h\) represents the density of susceptible humans. It comes naturally to wonder what happens if standard incidence is replaced by the mass action. In this work, we continue to borrow the idea in [11, 22, 32] that the density of susceptible humans is H(x) and adopt the mass action incidence, which result in the force of infection for humans and mosquitoes at space x and time t given by \(b\beta (x)S_mI_h\) and \(c\beta (x)H(x)I_m\), respectively.

This work is also inspired by some recent works on diffusive disease models with age structure (see, for example, [5, 9, 10, 35, 36], aiming to understand the effects of the spatial heterogeneity and infection age on disease transmission. For the standard age-space structured susceptible–infective–recovered (SIR) model, Chekroun and Kuniya [5] reformulated the model by a hybrid system of one diffusive equation and one Volterra integral equation, and studied the threshold dynamics for the disease extinction and persistence in one-dimensional domain. Further, the global stability problem of a constant equilibrium was achieved by constructing Lyapunov function. In another works, the existence of travelling wave solutions of age-space structured SIR model with or without birth and death processes was established in a spatially unbounded domain [9, 10]. For a age-space structured SIR model with seasonality, Zhang and Wang [36] established the threshold dynamics that basic reproduction number more than one or less than one determines whether or not disease extinction. Yang et al. [35] made an attempt to extend the methods and ideas in [5] to propose a model for the spatial spreading of brucellosis in a continuous bounded domain. Some basic mathematical arguments, including the existence and uniqueness of the solution and threshold dynamics, were successfully addressed.

Unlike in [13, 21] where disease transmission is modeled with a fixed incubation period in a spatial bounded domain, here we intend to incorporate an infection age to be a continuous variable and the fixed incubation period is ignored. This work can be considered as a continuation of the work [21]. For convenience, we shall adopt some notations used in [21], as proceeded below. With these considerations, we propose the following age-space-structured malaria model:

$$\begin{aligned} \left\{ \begin{array}{ll} \displaystyle \frac{\partial S_m(t,x)}{\partial t} -D_m\Delta S_m(t,x) = \mu (x)-\beta (x)S_m(t,x)I_h(t,x)-d_m S_m(t,x),&{}\quad x\in \Omega ,t>0, \\ \displaystyle \frac{\partial i_m(t,a,x)}{\partial t}+ \frac{\partial i_m(t,a,x)}{\partial a} - D_m\Delta i_m(t,a,x) =-d_m i_m(t,a,x),&{}\quad x\in \Omega ,t>0,a\ge 0, \\ \displaystyle \frac{\partial I_h(t,x)}{\partial t} - D_h\Delta I_h(t,x) =H(x)\int \limits _0^\infty \beta _1(a)i_m(t,a,x)\mathrm{d}a-(d_h+\rho )I_h(t,x),&{}\quad x\in \Omega ,t>0, \\ \displaystyle i_m(t,0,x)=\beta (x)S_mI_h,&{}\quad x\in \Omega ,t>0, \end{array}\right. \end{aligned}$$
(1.4)

with the following initial condition

$$\begin{aligned} S_m(0,x)=\phi _1(x)\ge 0,\ i_m(0,a, x)=\phi _2(a, x)\in L_+^1({\mathbb {R}}_+; C(\Omega )),\ I_h(0,x)=\phi _3(x)\ge 0,\ x\in {\overline{\Omega }}, \end{aligned}$$
(1.5)

and boundary condition

$$\begin{aligned} \frac{\partial S_m(t,x)}{\partial n}=\frac{\partial i_m(t,a,x)}{\partial n}=\frac{\partial I_h(t,x)}{\partial n}=0,\ x\in \partial \Omega ,\ t>0. \end{aligned}$$
(1.6)

Here, \(\beta (x)\) and \(\beta _1(a)\in L_+^\infty ({\mathbb {R}}_+)\) are the age-dependent disease transmission rate. All the location-dependent parameters are continuous, strictly positive and uniformly bounded functions on \({\overline{\Omega }}\).

We point here that the main difficulty lies in that (1.4) is formulated in a non-uniform Banach space. The existence and positivity of the solution should be carefully verified, to which adopt the methods used in [5, 35]. Comparing to [5, 35], we need to turn the existence of global nonnegative classical solution into a fixed point problem by inserting the first and third equation to the second equation due to the cross-infection mechanism. On the other hand, the cross-infection mechanism brings us difficulties in obtaining the explicit formula for the basic reproduction number through defining the next generation operator. Finally, in a homogeneous case, the local stability of equilibria achieved by studying the distribution of characteristic roots of characteristic equation, the strong persistence result and the global stability of equilibria achieved by Lyapunov functions in different cases are indeed not trivial, as proceed below.

We proceed the paper as follows: In section 2, we present the preliminary results on (1.4) with (1.5) and (1.6), and the well-posedness of the model is shown by verifying that the solution exists globally and the model admits a global attractor. Section 3 is spent on giving the basic reproduction number for (1.4) by appealing to the theories of the next generation operator and renewal equations. In section 4, the local and global stability problem for space-independent equilibria of (1.4) is studied in a homogeneous case. A brief conclusion section ends the paper.

2 Preliminaries

Denote by \({\mathbb {X}} = C({\overline{\Omega }},{\mathbb {R}})\) the Banach space endowed with the norm \(\Vert \phi \Vert _{\mathbb {X}} = \sup \limits _{x\in {\overline{\Omega }}}\{|\phi (x)|\}\), and let \({\mathbb {X}}_+\) be its positive cone. Denote by \({\mathbb {Y}} = L_1({\mathbb {R}}_+, {\mathbb {X}})\) the space of Lebesgue integrable functions endowed with the norm \(\Vert \varphi \Vert _{{\mathbb {Y}}} = \int \limits _0^\infty |\varphi (a)|_{{\mathbb {X}}}\mathrm{d}a\). The positive cone of \({\mathbb {Y}}\) is \({\mathbb {Y}}_+\).

Suppose that \(T_1(t), T_2(t): {\mathbb {X}}\rightarrow {\mathbb {X}},\ t\ge 0\) are the strongly continuous semigroups corresponding to the operators, \(D_m \Delta \) and \(D_h\Delta \), subject to (1.6), respectively. Clearly,

$$\begin{aligned} (T_i(t)\phi )(x) =\int \limits _{\Omega }\Gamma _i(t, x, y)\phi (y)\mathrm{d}y,\ t \ge 0,\ \phi \in {\mathbb {X}}, \end{aligned}$$
(2.1)

where \(\Gamma _i(t, x, y), i=1,2\) is the Green function associated with \(D_m \triangle \) and \(D_h\triangle \) subject to (1.6). It then follows from [27, Corollary 7.2.3] (see also in [24, Theorem 1.5]) that \(T_i(t) : {\mathbb {X}}\rightarrow {\mathbb {X}},\ t\ge 0\), \(i = 1, 2\), is strongly positive and compact. Further, \(T(t) = (T_1(t), T_1(t), T_2(t)): {\mathbb {X}}^3\rightarrow {\mathbb {X}}^3,\ t\ge 0\) forms a strongly continuous semigroup.

We place our problem on the state space \({\mathbb {X}}\times {\mathbb {Y}}\times {\mathbb {X}}\). For convenience, in what follows, we denote

$$\begin{aligned} {\overline{\beta }}_1=ess.\sup \limits _{a\in {\mathbb {R}}_+}\beta _1(a),\ \chi ^* = \max _{x\in {\overline{\Omega }}}\chi (x)\ \text{ and }\ \chi _* = \min _{x\in {\overline{\Omega }}}\chi (x), \end{aligned}$$

where \(\chi =\mu , \beta , H\), respectively.

We next aim to prove that the solution of (1.4) exists globally and (1.4) admits a global attractor. For simplicity, we denote by \(\mathbb {{\tilde{X}}}={\mathbb {X}}\times {\mathbb {Y}}\times {\mathbb {X}}\) and \(\mathbb {{\tilde{X}}}_+={\mathbb {X}}_+\times {\mathbb {Y}}_+\times {\mathbb {X}}_+\). We first present the main result.

Theorem 2.1

For any \(\phi =(\phi _1, \phi _2, \phi _3)\in \mathbb {{\tilde{X}}}\), (1.4) admits a unique global nonnegative classical solution. Further, the solution semiflow, i.e.,

$$\begin{aligned} \Phi (t)\phi := u(t, x; \phi ) = (S_m(t, x; \phi _1), i_m(t, a, x; \phi _2), I_h(t, x; \phi _3)),\ a\ge 0, \end{aligned}$$

possesses a global attractor in \(\mathbb {{\tilde{X}}}_+\).

We shall first introduce the following lemmas and then combine with them to prove Theorem 2.1. For convenience, we always denote

$$\begin{aligned} {\mathcal {B}}(t,x):=i_m(t,0,x)=\beta (x)S_mI_h. \end{aligned}$$
(2.2)

For positive T, we define the space

$$\begin{aligned} {\mathbb {Y}}_T=C([0,T],{\mathbb {X}})\ \text{ with }\ \Vert \psi \Vert _{{\mathbb {Y}}_T} = \sup \limits _{0\le t\le T}\Vert \psi (t,\cdot )\Vert _{\mathbb {X}},\ \psi \in {\mathbb {Y}}_T. \end{aligned}$$

Lemma 2.2

For any \(\phi \in \mathbb {{\tilde{X}}}_+\), (1.4) admits a unique nonnegative solution defined on \([0,T]\times {\overline{\Omega }}\) with \(T>0\).

Proof

In view of the \(S_m\) equation of (1.4), direct calculation yields

$$\begin{aligned} S_m ={\mathbb {F}}_1(t,x) + \int \limits _0^t e^{-d_m (t-s)}\int \limits _\Omega \Gamma _1(t-s,x,y)[\mu (y)-{\mathcal {B}}(s,y)]\mathrm{d}y\mathrm{d}s,\ (t,x)\in [0,T]\times \Omega , \end{aligned}$$
(2.3)

where \({\mathbb {F}}_1(t,x) = e^{-d_m t}\int \limits _\Omega \Gamma _1(t,x,y)\phi _1(y)\mathrm{d}y\). Similarly, in view of the \(I_h\) equation of (1.4), we get

$$\begin{aligned} I_h ={\mathbb {F}}_2(t,x) + \int \limits _0^t e^{-(d_h+\rho ) (t-s)}\int \limits _\Omega \Gamma _2(t-s,x,y)H(y)\int \limits _0^\infty \beta _1(a)i_m(s,a,y)\mathrm{d}a\mathrm{d}y\mathrm{d}s,\ (t,x)\in [0,T]\times \Omega ,\nonumber \\ \end{aligned}$$
(2.4)

where \({\mathbb {F}}_2(t,x) = e^{-(d_h+\rho )t}\int \limits _\Omega \Gamma _2(t,x,y)\phi _3(y)\mathrm{d}y\). It follows from (1.3) that

$$\begin{aligned} \int \limits _0^\infty \beta _1(a)i_m(t,a,x)\mathrm{d}a = \int \limits _0^t\beta _1(a)\Pi (a)\int \limits _\Omega \Gamma _1(a,x,y){\mathcal {B}}(t-a,y)\mathrm{d}y\mathrm{d}a + {\mathbb {F}}_3(t,x), \end{aligned}$$
(2.5)

where \({\mathbb {F}}_3(t,x) = \int \limits _0^\infty \beta _1(a+t)\frac{\Pi (a+t)}{\Pi (a)}\int \limits _\Omega \Gamma _1(t,x,y)\phi _2(a,y)\mathrm{d}y\mathrm{d}a\).

Plugging (2.5) into (2.4) and then substituting (2.3) and (2.4) in (2.2) yield

$$\begin{aligned} {\mathcal {B}}(t,x)= & {} \beta (x)\left( {\mathbb {F}}_1 + \int \limits _0^t e^{-d_m (t-s)} \int \limits _\Omega \Gamma _1(t-s,x,y)[\mu (y)-{\mathcal {B}}(s,y)]\mathrm{d}y\mathrm{d}s\right) \times \left( {\mathbb {F}}_2 + {\mathbb {F}}_4\right. \nonumber \\&\left. + \int \limits _0^t e^{-(d_h+\rho ) (t-s)}\int \limits _\Omega \Gamma _2(t-s,x,y)H(y)\int \limits _0^s\beta _1(a)\Pi (a) \int \limits _\Omega \Gamma _1(a,y,z){\mathcal {B}}(s-a,z)\mathrm{d}z\mathrm{d}a\mathrm{d}y\mathrm{d}s \right) \nonumber \\:= & {} \ {\mathbb {F}}({\mathcal {B}})(t,x), \end{aligned}$$
(2.6)

where \({\mathbb {F}}_4(t,x)=\int \limits _0^t e^{-(d_h+\rho ) (t-s)}\int \limits _\Omega \Gamma _2(t-s,x,y)H(y)\int \limits _0^\infty \beta _1(a+s)\frac{\Pi (a+s)}{\Pi (a)}\int \limits _\Omega \Gamma _1(s,y,z)\phi _2(a,z)\mathrm{d}z\mathrm{d}a\mathrm{d}y\mathrm{d}s\).

Next, we claim that the nonlinear operator \({\mathbb {F}}\) has a fixed point, which in turn ensures that (1.4) admits a unique solution on [0, T]. For ease of notations, we denote

$$\begin{aligned} \left\{ \begin{array}{ll} \displaystyle \overline{{\mathbb {F}}}_2(t,x):=\ {\mathbb {F}}_2(t,x)+{\mathbb {F}}_4(t,x), \\ \displaystyle {\mathbb {G}}_1({\mathcal {B}}) := \int \limits _0^t e^{-d_m (t-s)}\int \limits _\Omega \Gamma _1(t-s,x,y)[\mu (y)-{\mathcal {B}}(s,y)]\mathrm{d}y\mathrm{d}s,\\ \displaystyle {\mathbb {G}}_2({\mathcal {B}}) := \int \limits _0^t e^{-(d_h+\rho ) (t-s)}\int \limits _\Omega \Gamma _2(t-s,x,y)H(y)\int \limits _0^s\beta _1(a)\Pi (a) \int \limits _\Omega \Gamma _1(a,y,z){\mathcal {B}}(s-a,z)\mathrm{d}z\mathrm{d}a\mathrm{d}y\mathrm{d}s. \end{array}\right. \end{aligned}$$
(2.7)

It then follows that the nonlinear operator \({\mathbb {F}}\) can be rewritten as:

$$\begin{aligned} {\mathbb {F}}({\mathcal {B}}) = \beta (x)({\mathbb {F}}_1 + {\mathbb {G}}_1({\mathcal {B}}))(\overline{{\mathbb {F}}}_2 + {\mathbb {G}}_2({\mathcal {B}})). \end{aligned}$$

By appealing the Banach–Picard fixed point theorem, we only need to verify that the nonlinear operator \({\mathbb {F}}\) is a strict contraction in \({\mathbb {Y}}_T\). To this end, we choose any \({\mathcal {B}}_1, {\mathcal {B}}_2\in {\mathbb {Y}}_T\) and let \(\widetilde{{\mathcal {B}}} = {\mathcal {B}}_1 - {\mathcal {B}}_2\) such that

$$\begin{aligned} {\mathbb {F}}({\mathcal {B}}_1) - {\mathbb {F}}({\mathcal {B}}_2)&= \beta (x)({\mathbb {F}}_{1}{\mathbb {G}}_2(\widetilde{{\mathcal {B}}}) - \overline{{\mathbb {F}}}_2\widehat{{\mathbb {G}}}_1(\widetilde{{\mathcal {B}}}) + {\mathbb {G}}_1({\mathcal {B}}_1){\mathbb {G}}_2(\widetilde{{\mathcal {B}}}) - {\mathbb {G}}_2({\mathcal {B}}_2)\widehat{{\mathbb {G}}}_1(\widetilde{{\mathcal {B}}}))\\&\le \beta ^*\left| {\mathbb {F}}_{1}\overline{{\mathbb {G}}}_2 - \overline{{\mathbb {F}}}_2\overline{{\mathbb {G}}}_1 + {\mathbb {G}}_1 \overline{{\mathbb {G}}}_2 - \overline{{\mathbb {G}}}_1{\mathbb {G}}_2 \right| \widetilde{{\mathcal {B}}}, \end{aligned}$$

where

$$\begin{aligned} \widehat{{\mathbb {G}}}_1({\mathcal {B}})&:= \int \limits _0^t e^{-d_m (t-s)}\int \limits _\Omega \Gamma _1(t-s,x,y){\mathcal {B}}(s,y)\mathrm{d}y\mathrm{d}s,\\ \overline{{\mathbb {G}}}_1&= \int \limits _0^t e^{-d_m (t-s)}\int \limits _\Omega \Gamma _1(t-s,x,y)\mathrm{d}y\mathrm{d}s, \end{aligned}$$

and

$$\begin{aligned} \overline{{\mathbb {G}}}_2 = \int \limits _0^t e^{-(d_h+\rho ) (t-s)}\int \limits _\Omega \Gamma _2(t-s,x,y)H(y)\int \limits _0^s\beta _1(a)\Pi (a) \int \limits _\Omega \Gamma _1(a,y,z)\mathrm{d}z\mathrm{d}a\mathrm{d}y\mathrm{d}s. \end{aligned}$$

Putting

$$\begin{aligned} {\tilde{h}}(T) = \beta ^*\sup \limits _{0\le s\le T}\left| {\mathbb {F}}_{1}(s,x)\overline{{\mathbb {G}}}_2(s,x) - \overline{{\mathbb {F}}}_2(s,x)\overline{{\mathbb {G}}}_1(s,x) + {\mathbb {G}}_1(s,x)\overline{{\mathbb {G}}}_2(s,x) + \overline{{\mathbb {G}}}_1(s,x){\mathbb {G}}_2(s,x)\right| _{\mathbb {X}}, \end{aligned}$$

which leads to

$$\begin{aligned} \left| {\mathbb {F}}({\mathcal {B}}_1) - {\mathbb {F}}({\mathcal {B}}_2)\right| \le {\tilde{h}}(T)\left| {\mathcal {B}}_1 - {\mathcal {B}}_2\right| . \end{aligned}$$

Choosing sufficiently small \(T >0\) such that \({\tilde{h}}(T)<1\). We arrive at the conclusion that \({\mathbb {F}}\) is a strict contraction in \({\mathbb {Y}}_T\), that is, the nonlinear operator \({\mathbb {F}}\) has a unique fixed point. This proves Lemma 2.2.

\(\square \)

We next to prove that the local solution of (1.4) remains positive for any \(\phi \in \mathbb {{\tilde{X}}}_+\).

Lemma 2.3

For any \(\phi \in \mathbb {{\tilde{X}}}_+\), we have that \(\forall t\in [0,T]\) and \(x\in \Omega \),

$$\begin{aligned} S_m(t,x)>0,\ {\mathcal {B}}(t,x)>0\ \text{ and }\ I_h(t,x)>0. \end{aligned}$$

Proof

It is easily seen that due to the positivity of \(\mu (x)\) and \(\phi _1\),

$$\begin{aligned} S_m(t,x)=\widehat{{\mathbb {F}}}_1 + \int \limits _0^t e^{-\int \limits _s^t\left( d_m + \frac{{\mathcal {B}}(\tau ,y)}{S_m(\tau ,y)}\right) d\tau } \mu (x)\int \limits _\Omega \Gamma _1(t-s,x,y)\mathrm{d}y\mathrm{d}s>0, \end{aligned}$$

where \(\widehat{{\mathbb {F}}}_1 = e^{-\int \limits _{s-t}^s\left( d_m + \frac{{\mathcal {B}}(\tau ,y)}{S_m(\tau ,y)}\right) d\tau } \int \limits _\Omega \Gamma _1(t,x,y)\phi _1(y)\mathrm{d}y\).

We next to prove the positivity of \({\mathcal {B}}(t,x)\) by the method of Picard sequences. We first set

$$\begin{aligned} {\mathcal {B}}_0(t,x) = \beta (x)S_m(t,x)\overline{{\mathbb {F}}}_2(t,x)>0, \end{aligned}$$

where \(\overline{{\mathbb {F}}}_2(t,x)\) is defined in (2.7). Assume that \({\mathcal {B}}_n(t, x)>0\ (n\in {\mathbb {N}})\). We then directly get

$$\begin{aligned} {\mathcal {B}}_{n+1}(t,x)&= {\mathcal {B}}_0(t,x)+ \beta (x)S_m(t,x)\int \limits _0^t e^{-(d_h+\rho ) (t-s)}\int \limits _\Omega \Gamma _2(t-s,x,y)H(y)\\&\quad \times \int \limits _0^s\beta _1(s-a)\Pi (s-a) \int \limits _\Omega \Gamma _1(s-a,y,z){\mathcal {B}}_n(a,z)\mathrm{d}z\mathrm{d}a\mathrm{d}y\mathrm{d}s\\&> 0. \end{aligned}$$

It remains to show that the sequence \(\{{\mathcal {B}}_n(t,x)\}_{0}^{\infty }\) converges to \({\mathcal {B}}(t,x)\) in the sense that \(\lim \limits _{n\rightarrow \infty }{\mathcal {B}}_n(t,x)={\mathcal {B}}(t,x)\). To achieve this, we introduce

$$\begin{aligned} \widehat{{\mathcal {B}}}_{n}(t,x) =&\ e^{-\lambda t}{\mathcal {B}}_{n}(t,x),~ \text{ for }\ \lambda \in {\mathbb {R}}_+. \end{aligned}$$

Hence,

$$\begin{aligned} \widehat{{\mathcal {B}}}_{n+1}(t,x)&= e^{-\lambda t}{\mathcal {B}}_0(t,x)+ \beta (x)S_m(t,x)\int \limits _0^t e^{-(d_h+\rho ) (t-s)}\int \limits _\Omega \Gamma _2(t-s,x,y)H(y)\\&\quad \times \int \limits _0^s\beta _1(a)\Pi (a) \int \limits _\Omega \Gamma _1(a,y,z)e^{-\lambda t}{\mathcal {B}}_n(s-a,z)\mathrm{d}z\mathrm{d}a\mathrm{d}y\mathrm{d}s\\&= e^{-\lambda t}{\mathcal {B}}_0(t,x)+ \beta (x)S_m(t,x)\int \limits _0^t e^{-(d_h+\rho ) s}\int \limits _\Omega \Gamma _2(s,x,y)H(y)\\&\quad \times \int \limits _0^{t-s}\beta _1(a)\Pi (a) \int \limits _\Omega \Gamma _1(a,y,z)e^{-\lambda t}{\mathcal {B}}_n(t-s-a,z)\mathrm{d}z\mathrm{d}a\mathrm{d}y\mathrm{d}s\\&=e^{-\lambda t}{\mathcal {B}}_0(t,x)+ \beta (x)S_m(t,x)\int \limits _0^t e^{-(d_h+\rho ) s}\int \limits _\Omega \Gamma _2(s,x,y)H(y)\\&\quad \times \int \limits _0^{t-s}\beta _1(a)\Pi (a) \int \limits _\Omega \Gamma _1(a,y,z)e^{-\lambda (a+s)}e^{-\lambda (t-a-s)}{\mathcal {B}}_n(t-s-a,z)\mathrm{d}z\mathrm{d}a\mathrm{d}y\mathrm{d}s\\&= e^{-\lambda t}{\mathcal {B}}_{0}(t,x) + \beta (x)S_m(t,x)\int \limits _0^t e^{-(d_h +\rho )s}\int \limits _\Omega \Gamma _2(s,x,y)H(y)\\&\quad \times \int \limits _0^{t-s}\beta _1(a)\Pi (a)e^{-\lambda a}\int \limits _\Omega \Gamma _1(a,y,z)e^{-\lambda s}\widehat{{\mathcal {B}}}_n(t-s-a,z)\mathrm{d}z\mathrm{d}a\mathrm{d}y\mathrm{d}s. \end{aligned}$$

Note that both sides of above equality always hold. It allows us to choose \({\tilde{x}}\in {\overline{\Omega }}\) such that \({\mathcal {B}}^*(t,{\widetilde{x}}) = \max \limits _{t\in [0,T], x\in {\overline{\Omega }}}{\widehat{{\mathcal {B}}}(t,x)}\). Hence, for any \(n\in {\mathbb {N}}\),

$$\begin{aligned} \left\| {\mathcal {B}}^*_{n+1}-{\mathcal {B}}^*_{n}\right\| _\infty&\le \beta ^*S^*_m\int \limits _0^\infty e^{-(d_h +\rho )s}\int \limits _\Omega \Gamma _2(s,x,y)H(y)\\&\quad \times \int \limits _0^{\infty }\beta _1(a)\Pi (a)e^{-\lambda a}\int \limits _\Omega \Gamma _1(a,y,z)e^{-\lambda s}\mathrm{d}z\mathrm{d}a\mathrm{d}y\mathrm{d}s\left\| {\mathcal {B}}^*_{n}-{\mathcal {B}}^*_{n-1}\right\| _\infty \\&\le \frac{\beta ^*S^*_m{\overline{\beta }}_1H^*}{\lambda ^2}\left\| {\mathcal {B}}^*_{n}-{\mathcal {B}}^*_{n-1}\right\| _\infty , \end{aligned}$$

where \(H^*=\max \limits _{x\in \Omega }\{H(x)\}\) and \(S^*_m = \max \limits _{t\in [0,T]}\left\| S_m(t,\cdot )\right\| _{\mathbb {X}}\). By repeating the iteration, it is easily seen that

$$\begin{aligned} \left\| {\mathcal {B}}^*_{n+1}-{\mathcal {B}}^*_{n}\right\| _\infty \le L_\lambda \left\| {\mathcal {B}}^*_{n}-{\mathcal {B}}^*_{n-1}\right\| _\infty \le L_\lambda ^n\left\| {\mathcal {B}}^*_{1}-{\mathcal {B}}^*_{0}\right\| _\infty , \end{aligned}$$

where \(L_\lambda = \frac{\beta ^*S^*_m{\overline{\beta }}_1H^*}{\lambda ^2}\). Hence, for any \(m, n\in {\mathbb {N}}\),

$$\begin{aligned} \left\| {\mathcal {B}}^*_{m}-{\mathcal {B}}^*_{n}\right\| _\infty \le \frac{L_\lambda ^n}{1-L_\lambda }\left\| {\mathcal {B}}^*_{1}-{\mathcal {B}}^*_{0}\right\| _\infty . \end{aligned}$$

Choose \(\lambda \) large enough that \(L_\lambda < 1\). Consequently, as \(n\rightarrow \infty \), \(\left\| {\mathcal {B}}^*_{m}-{\mathcal {B}}^*_{n}\right\| _\infty \rightarrow 0\), implying that \(\lim \limits _{n\rightarrow \infty }{\mathcal {B}}_n(t,x)={\mathcal {B}}(t,x)\), \(\forall t\in [0,T]\) and \(x\in \Omega \). The positivity of \({\mathcal {B}}_n(t,x)\) directly implies that \({\mathcal {B}}(t,x)\) is positive.

We next to show that \(I_h(t,x)>0\) by contradiction. Assume to the contrary that there exist \(x_*\in \Omega \) and \(t_1\in (0, T)\) such that

$$\begin{aligned} \left\{ \begin{array}{ll} I_h(t,x)\ge 0, &{} \quad t \in [0, t_1]\hbox { and }x\in \Omega ; \\ I_h(t, x_*) = 0, &{} \quad t=t_1\hbox { and }x_*\in \Omega ; \\ I_h(t+\epsilon , x_*)< 0, &{}\quad t=t_1, x_*\in \Omega \hbox { and }0 < \epsilon \ll 1. \end{array} \right. \end{aligned}$$

By (2.4), together with the positivity of \({\mathcal {B}}(t,x)\), we can easily obtain

$$\begin{aligned} I_h(t_1+\epsilon , x_*)&= {\mathbb {F}}_2+ \int \limits _0^{t_1+\epsilon } e^{-(d_h+\rho ) (t_1+\epsilon -s)}\int \limits _\Omega \Gamma _2(t_1+\epsilon -s,x,y)H(y)\\&\quad \times \int \limits _0^s\beta _1(a)\Pi (a) \int \limits _\Omega \Gamma _1(a,y,z){\mathcal {B}}(s-a,z)\mathrm{d}z\mathrm{d}a\mathrm{d}y\mathrm{d}s + {\mathbb {F}}_4 > 0, \end{aligned}$$

which results in a contradiction with \(I_h(t_1+\epsilon , x_*)< 0\). Hence, \(I_h(t,x)>0\) directly follows. This completes the proof. \(\square \)

Next, we confirm that the solution of (1.4) exists globally by extending solution existence interval from \([0, T]\times {\overline{\Omega }}\) to \([0,+\infty )\times {\overline{\Omega }}\).

Lemma 2.4

For any \(\phi \in \tilde{{\mathbb {X}}}_+\), (1.4) admits a unique nonnegative solution defined on \([0,\infty )\times {\overline{\Omega }}\).

Proof

In view of \(S_m\) equation of system (1.4), we know that \(S_m\) is governed by

$$\begin{aligned} \left\{ \begin{array}{ll} \displaystyle \frac{\partial {\hat{S}}_m}{\partial t} - D_m\Delta {\hat{S}}_m = \mu (x) - d_m{\hat{S}}_m,&{}\quad x\in \Omega ,\ t>0, \\ \displaystyle \frac{\partial {\hat{S}}_m}{\partial n} = 0 ,&{}\quad x\in \partial \Omega ,\ t> 0. \end{array}\right. \end{aligned}$$
(2.8)

According to [21, Lemma 1] and comparison principle, \(S_m\) is bounded above by the upper solution \(\frac{\mu ^*}{d_m}\), i.e., \(S_m \le \frac{\mu ^*}{d_m}=M_S\) for all \(t>0\) and \(x\in {\overline{\Omega }}\).

We now claim that for all \(t>0\) and \(x\in \Omega \), \({\mathcal {B}}(t,x)<\infty \) by contradiction. Assume to the contrary that there exist \(t^*>0\) and \(x^*\in \Omega \) such that \(\lim _{t\rightarrow t^*}{\mathcal {B}}(t,x^*) = +\infty \). In view of \(S_m\) equation of system (1.4), we get \(\lim _{t\rightarrow t^*}\partial _{t}S_m(t,x^*) = -\infty \), i.e., \(S_m(t,x^*)<0\) in near of \(t^*\). This leads to a contradiction to the positivity of \(S_m\) (see Lemma 2.3). Hence, \({\mathcal {B}}(t,x)<\infty \). Based on this fact, we assume that there exists \(M_B>0\) such that \({\mathcal {B}}(t,x)<M_B\).

Finally, we determine the boundedness of \(I_h(t,x)\). Let \({\widetilde{I}}= \int \limits _0^\infty i_m(t,a,x)\mathrm{d}a\), then it satisfies

$$\begin{aligned} {\widetilde{I}}&= \int \limits _0^t\Pi (a)\int \limits _\Omega \Gamma _1(a,x,y) {\mathcal {B}}(t-a,y)\mathrm{d}y\mathrm{d}a+\int \limits _0^\infty \frac{\Pi (a+t)}{\Pi (a)} \int \limits _\Omega \Gamma _1(t,x,y)\phi _2(a,y)\mathrm{d}y\mathrm{d}a\\&\le \int \limits _0^\infty \Pi (a)\int \limits _\Omega \Gamma _1(a,x,y)\mathrm{d}y\mathrm{d}aM_B +\int \limits _0^\infty \int \limits _\Omega \Gamma _1(t,x,y)\phi _2(a,y)\mathrm{d}y\mathrm{d}a := M_I. \end{aligned}$$

It then follows that

$$\begin{aligned} I_h(t,x)&\le H^*{\overline{\beta }}_1\int \limits _0^\infty e^{-(d_h+\rho ) (t-s)}\int \limits _\Omega \Gamma _2(t-s,x,y){\widetilde{I}}(t,y)\mathrm{d}y\mathrm{d}s + \int \limits _\Omega \Gamma _2(t,x,y)\phi _3(y)\mathrm{d}y\nonumber \\&\le \frac{H^*{\overline{\beta }}_1M_I}{d_h+\rho } + \Vert \phi _3\Vert _X := M_{I_h}. \end{aligned}$$
(2.9)

Therefore, the solution of (1.4) exists globally. This proves Lemma 2.4. \(\square \)

By utilizing the previous lemmas, we shall give the proof of Theorem 2.1.

Proof of Theorem 2.1

By the assertions in Lemmas 2.2, 2.3 and 2.4, we know that the solution of (1.4) with initial condition \(\phi =(\phi _1, \phi _2, \phi _3)\in \mathbb {{\tilde{X}}}_+\) exists globally. Let \(\Phi (t) : \mathbb {{\tilde{X}}}_+\rightarrow \mathbb {{\tilde{X}}}_+, t \ge 0\), be the semiflow generated by the solution of (1.4), i.e.,

$$\begin{aligned} \Phi (t)\phi := u(t, x; \phi ) = (S_m(t, x; \phi _1), i_m(t, a, x; \phi _2), I_h(t, x; \phi _3)). \end{aligned}$$

From Lemma 2.4, solution semiflow \(\Phi (t)\) is ultimately bounded. Hence, we can apply [16, Theorem 2.4.6] to confirm that (1.4) possesses a global attractor. This proves Theorem 2.1.

Throughout of the paper, we define the following positively invariant set, at which the dynamics of (1.4) are confined.

Letting \({\mathcal {M}}(t):=\int \limits _\Omega S_m \mathrm{d}x+\int \limits _\Omega {\widetilde{I}}\mathrm{d}x\). Integrate the \(i_m\) equation from 0 to \(\infty \), and integrating the \(S_m\) and \(i_m\) equation of system (1.4) over \(\Omega \) and then adding them up, yield

$$\begin{aligned} \frac{d{\mathcal {M}}(t)}{\mathrm{d}t} \le \mu ^*|\Omega | - d_m{\mathcal {M}}(t). \end{aligned}$$

Hence, we obtain the a priori estimate

$$\begin{aligned} {\mathcal {M}}(t)\le {\mathcal {M}}(0)e^{-d_mt}+\frac{\mu ^*|\Omega |}{d_m}(1-e^{-d_mt})\le \max \left\{ {\mathcal {M}}(0), \frac{\mu ^*|\Omega |}{d_m}\right\} . \end{aligned}$$

This implies

$$\begin{aligned} \limsup _{t\rightarrow \infty }{\mathcal {M}}(t)\le \frac{\mu ^*|\Omega |}{d_m}. \end{aligned}$$

Hence, there exists \(t_0>0\) such that for any \(t > t_0\), \({\mathcal {M}}(t) \le \frac{\mu ^*|\Omega |}{d_m}\).

Denote by \({\widehat{I}}_h(t) = \int \limits _\Omega I_h(t,x)\mathrm{d}x\). We directly have

$$\begin{aligned} \frac{d{\widehat{I}}_h(t)}{\mathrm{d}t} \le \frac{H^*{\overline{\beta }}_1\mu ^*|\Omega |}{d_m} - (d_h + \rho ){\widehat{I}}_h(t). \end{aligned}$$

We integrate this differential inequality to obtain that for \(t_1 > t_0\),

$$\begin{aligned} \widehat{I_h}(t) \le \frac{H^*{\overline{\beta }}_1\mu ^*|\Omega |}{d_m(d_h + \rho )},\ \text{ for }\ t > t_1. \end{aligned}$$

Consequently, we can define the following positively invariant set

$$\begin{aligned} \begin{array}{ll} {\mathcal {D}} = \left\{ \phi \in \mathbb {{\tilde{X}}}_+|\ 0< {\mathcal {M}}(t)< \frac{\mu ^*|\Omega |}{d_m}, 0< {\widehat{I}}_h(t) < \frac{H^*{\overline{\beta }}_1\mu ^*|\Omega |}{d_m(d_h + \rho )}\right\} . \end{array} \end{aligned}$$
(2.10)

\(\square \)

3 The basic reproduction number

Obviously, (1.4) has a disease-free steady state \(E_0 = (\frac{\mu (x)}{d_m},0,0)\). Assuming that both humans and mosquitos are near \(E_0\). Following the standard procedures as those in [30] and [7], this section is spent on defining basic reproduction number of model (1.4).

Linearizing system (1.4) at \(E_0\) for infectious components yields

$$\begin{aligned} \left\{ \begin{array}{ll} \displaystyle \frac{\partial i_m(t,a,x)}{\partial t}+ \frac{\partial i_m(t,a,x)}{\partial a} - D_m\Delta i_m(t,a,x) =-d_m i_m(t,a,x),&{}\quad x\in \Omega ,\ a\ge 0, \\ \displaystyle i_m(t,0,x) = \beta (x)\frac{\mu (x)}{d_m}I_h:=\overline{{\mathcal {B}}}(t,x),&{}\quad x\in \Omega , \\ \displaystyle \frac{\partial I_h}{\partial t} -D_h \triangle I_h= H(x)\int \limits _0^\infty \beta _1(a)i_m(t,a,x)\mathrm{d}a - (d_h+\rho )I_h,&{}\quad x\in \Omega ,\ a\ge 0,\\ \displaystyle \frac{\partial i_m(t,a,x)}{\partial n}=\frac{\partial I_h}{\partial n}=0,&{}\quad x\in \partial \Omega . \end{array}\right. \end{aligned}$$
(3.1)

We integrate the \(i_m\) equation of (3.1) along the characteristic line \(t-a=const.\)

$$\begin{aligned} i_m(t,a,x)= \left\{ \begin{array}{ll} \displaystyle \int \limits _{\Omega } \Gamma _1(a,x,y)\overline{{\mathcal {B}}}(t-a,y)\mathrm{d}y\Pi (a), &{}\quad t>a,\ x\in \Omega ,\\ \displaystyle \int \limits _{\Omega } \Gamma _1(t,x,y)\phi _2(a-t,y)\mathrm{d}y\frac{\Pi (a)}{\Pi (a-t)},&{}\quad a\ge t,\ x\in \Omega . \end{array}\right. \end{aligned}$$
(3.2)

We substitute (3.2) in \(I_h\) (defined in (2.4)) resulting in

$$\begin{aligned} \overline{{\mathcal {B}}}(t,x)&= \beta (x)\frac{\mu (x)}{d_m}\bigg ({\mathbb {F}}_2 + {\mathbb {F}}_4+\int \limits _0^t e^{-(d_h+\rho ) (t-s)}\int \limits _\Omega \Gamma _2(t-s,x,y)H(y)\nonumber \\&\quad \times \int \limits _0^s\beta _1(a)\Pi (a)\int \limits _{\Omega } \Gamma _1(a,y,z)\overline{{\mathcal {B}}}(s-a,z)\mathrm{d}z\mathrm{d}a\mathrm{d}y\mathrm{d}s\bigg )\nonumber \\&= \beta (x)\frac{\mu (x)}{d_m}\bigg ({\mathbb {F}}_2 + {\mathbb {F}}_4+\int \limits _0^t e^{-(d_h+\rho )s}\int \limits _\Omega \Gamma _2(s,x,y)H(y)\nonumber \\&\quad \times \int \limits _0^{t-s}\beta _1(a)\Pi (a)\int \limits _{\Omega } \Gamma _1(a,y,z)\overline{{\mathcal {B}}}(t-s-a,z)\mathrm{d}z\mathrm{d}a\mathrm{d}y\mathrm{d}s\bigg ), \end{aligned}$$
(3.3)

where \({\mathbb {F}}_2\) and \({\mathbb {F}}_4\) are defined in (2.4) and (2.6), respectively. Since (3.3) is a renewal equation, we can make Laplace transformation to (3.3)

$$\begin{aligned} \int \limits _0^{\infty }e^{-\lambda t}\overline{{\mathcal {B}}}(t,x) \mathrm{d}t&= \beta (x)\frac{\mu (x)}{d_m}\int \limits _0^\infty e^{-\lambda t}\int \limits _0^t e^{-(d_h+\rho )s}\int \limits _\Omega \Gamma _2(s,x,y)H(y)\\&\quad \times \int \limits _0^{t-s}\beta _1(a)\Pi (a)\int \limits _{\Omega } \Gamma _1(a,y,z)\overline{{\mathcal {B}}}(t-s-a,z)\mathrm{d}z\mathrm{d}a\mathrm{d}y\mathrm{d}s\mathrm{d}t. \end{aligned}$$

Interchanging the order of integration (t and s) and letting \(t-s={\tilde{t}}\) (for simplicity, still denote \({\tilde{t}}\) by t) yield

$$\begin{aligned} \int \limits _0^{\infty }e^{-\lambda t}\overline{{\mathcal {B}}}(t,x) \mathrm{d}t&= \beta (x)\frac{\mu (x)}{d_m}\int \limits _0^\infty e^{-(d_h +\rho )s}\int \limits _s^{\infty }e^{-\lambda t}\int \limits _\Omega \Gamma _2(s,x,y)H(y)\nonumber \\&\quad \times \int \limits _0^{t-s}\beta _1(a)\Pi (a)\int \limits _\Omega \Gamma _1(a,y,z)\overline{{\mathcal {B}}}(t-s-a,z)\mathrm{d}z\mathrm{d}a\mathrm{d}y\mathrm{d}t\mathrm{d}s\nonumber \\&=\beta (x)\frac{\mu (x)}{d_m}\int \limits _0^\infty e^{-(d_h+\rho )s} \int \limits _0^\infty e^{-\lambda (t+s)} \int \limits _\Omega \Gamma _2(s,x,y)H(y)\nonumber \\&\quad \times \int \limits _0^{t}\beta _1(a)\Pi (a)\int \limits _\Omega \Gamma _1(a,y,z) \overline{{\mathcal {B}}}(t-a,z)\mathrm{d}z\mathrm{d}a\mathrm{d}y\mathrm{d}t\mathrm{d}s. \end{aligned}$$

Interchanging the order of integration (t and a) and letting \(t-a={\bar{t}}\) (for simplicity, still denote \({\bar{t}}\) by t) yield

$$\begin{aligned} \int \limits _0^{\infty }e^{-\lambda t}\overline{{\mathcal {B}}}(t,x) \mathrm{d}t&=\beta (x)\frac{\mu (x)}{d_m}\int \limits _0^\infty e^{-(d_h+\rho )s}e^{-\lambda s} \int \limits _0^\infty \beta _1(a)\Pi (a) \int \limits _\Omega \Gamma _2(s,x,y)H(y)\nonumber \\&\quad \times \int \limits _a^{\infty }e^{-\lambda t} \int \limits _\Omega \Gamma _1(a,y,z)\overline{{\mathcal {B}}}(t-a,z)\mathrm{d}z\mathrm{d}t\mathrm{d}y\mathrm{d}a\mathrm{d}s\nonumber \\&=\beta (x)\frac{\mu (x)}{d_m}\int \limits _0^\infty e^{-(d_h+\rho )s}e^{-\lambda s} \int \limits _0^\infty \beta _1(a)\Pi (a) e^{-\lambda a}\int \limits _\Omega \Gamma _2(s,x,y)H(y)\nonumber \\&\quad \times \int \limits _0^{\infty }e^{-\lambda t} \int \limits _\Omega \Gamma _1(a,y,z)\overline{{\mathcal {B}}}(t,z)\mathrm{d}z\mathrm{d}t\mathrm{d}y\mathrm{d}a\mathrm{d}s. \end{aligned}$$

Consequently, interchanging the order of integration (t and z) yields

$$\begin{aligned} \int \limits _0^{\infty }e^{-\lambda t}\overline{{\mathcal {B}}}(t,x) \mathrm{d}t&=\beta (x)\frac{\mu (x)}{d_m}\int \limits _0^\infty e^{-(d_h+\rho )s}e^{-\lambda s} \int \limits _0^\infty \beta _1(a)\Pi (a)e^{-\lambda a} \int \limits _\Omega \Gamma _2(s,x,y)H(y)\nonumber \\&\quad \times \int \limits _\Omega \Gamma _1(a,y,z)\int \limits _0^{\infty }e^{-\lambda t}\overline{{\mathcal {B}}}(t,z) \mathrm{d}t\mathrm{d}z\mathrm{d}y\mathrm{d}a\mathrm{d}s. \end{aligned}$$
(3.4)

Setting \(\lambda =0\) leads to

$$\begin{aligned}&\int \limits _0^{\infty }\overline{{\mathcal {B}}}(t,x) \mathrm{d}t\nonumber \\&\quad =\beta (x)\frac{\mu (x)}{d_m}\int \limits _0^\infty e^{-(d_h+\rho )s}\int \limits _0^\infty \beta _1(a)\Pi (a)\int \limits _\Omega \Gamma _2(s,x,y)H(y)\int \limits _\Omega \Gamma _1(a,y,z)\int \limits _0^{\infty }\overline{{\mathcal {B}}}(t,z) \mathrm{d}t\mathrm{d}z\mathrm{d}y\mathrm{d}a\mathrm{d}s. \end{aligned}$$
(3.5)

Hence, the following operator \({\mathcal {L}}\) is termed as the next generation operator

$$\begin{aligned} {\mathcal {L}}[\psi ](x)=\beta (x)\frac{\mu (x)}{d_m}\int \limits _0^{\infty } e^{-(d_h+\rho )s}\int \limits _0^\infty \beta _1(a)\Pi (a)\int \limits _\Omega \Gamma _2(s,x,y)H(y)\int \limits _{\Omega } \Gamma _1(a,y,z)\psi (z)\mathrm{d}z\mathrm{d}y\mathrm{d}a\mathrm{d}s. \end{aligned}$$
(3.6)

The following result indicates that \({\mathcal {L}}\) defined in (3.6) is strictly positive and compact.

Lemma 3.1

Let \({\mathcal {L}}\) be defined in (3.6). Then, \({\mathcal {L}}\) is strictly positive and compact.

Proof

It is easily seen that the operator \({\mathcal {L}}\) is positive. Due to the properties of \(\Gamma _1\) and \(\Gamma _2\), we can select a bounded sequence \(\{\phi _n\}_{n\in {\mathbb {N}}}\) in \({\mathbb {X}}\) such that for \({\mathbb {M}} > 0\) and \(\forall x\in \Omega \),

$$\begin{aligned} |\phi _n|\le {\mathbb {M}},\ \text{ where }\ {\mathbb {M}} > 0. \end{aligned}$$

Hence,

$$\begin{aligned} {\mathcal {L}}[\phi _n](x)\le \frac{{\mathbb {M}}H^*\beta ^*\mu ^*}{d_m}\int \limits _0^{\infty }e^{-(d_h +\rho )s}\int \limits _0^\infty \beta _1(a)\Pi (a)\int \limits _\Omega \Gamma _2(s,x,y)\int \limits _{\Omega } \Gamma _1(a,y,z)\mathrm{d}z\mathrm{d}y\mathrm{d}a\mathrm{d}s, \end{aligned}$$

that is, \({\mathcal {L}}\) is uniformly bounded. According to the Arzelà–Ascoli theorem, it remains to confirm that \({\mathcal {L}}\) is equicontinuous. In fact, for any \(x,{\bar{x}}\in \Omega \) with \(|x-{\bar{x}}|<\delta \) and \(y\in {\overline{\Omega }}\),

$$\begin{aligned}&{\mathcal {L}}[\phi _n](x)-{\mathcal {L}}[\phi _n]({\bar{x}})\nonumber \\&\quad \le \frac{{\mathbb {M}}H^*\beta ^*\mu ^*}{d_m}\int \limits _0^{\infty }e^{-(d_h+\rho )s} \int \limits _0^\infty \beta _1(a)\Pi (a)\int \limits _\Omega \left| \Gamma _2(s,x,y) -\Gamma _2(s,{\bar{x}},y)\right| \int \limits _{\Omega } \Gamma _1(a,y,z)\mathrm{d}z\mathrm{d}y\mathrm{d}a\mathrm{d}s. \end{aligned}$$
(3.7)

Due to the compactness of the operator \(\Delta \) and the uniform continuity of \(\Gamma _2(s,x,y)\), finding \(\varepsilon _0>0\) ensures

$$\begin{aligned} \left| \Gamma _2(a,x,y)-\Gamma _2(a,{\bar{x}},y)\right| \le \frac{d_m\varepsilon _0}{{\mathbb {M}}H^*\beta ^*\mu ^*{\mathcal {M}}}, \end{aligned}$$

where \({\mathcal {M}}=|\Omega |\int \limits _0^{\infty }e^{-(d_h+\rho )s}ds\int \limits _0^\infty \beta _1(a)\Pi (a)\mathrm{d}a \) and \(|\Omega |\) is the volume of \(\Omega \). Using this \(\delta \) and \(\epsilon _0\), we know that \(|{\mathcal {L}}[\phi ](x)-{\mathcal {L}}[\phi ]({\bar{x}})|<\epsilon _0\), for all \(|x-{\bar{x}}|<\delta \). Consequently, the compactness of \({\mathcal {L}}\) directly follows (as \({\mathcal {L}}\) is uniformly bounded and equicontinuous). This proves Lemma 3.1. \(\square \)

According to the general results as those in [30] and [7], we define the basic reproduction number as:

$$\begin{aligned} \mathfrak {R}_0 = r(\mathcal {{\mathcal {L}}}), \end{aligned}$$

where \(r(\cdot )\) is the spectral radius of the operator \({\mathcal {L}}\). As usual, it is difficult to obtain spectral radius of the next-generation operator \({\mathcal {L}}\), if not impossible, so that we cannot get further information on dynamical properties of (1.4). To proceed further, we will consider the homogeneous case that

$$\begin{aligned} \mu (x)\equiv \mu , \beta (x)\equiv \beta , H(x)\equiv H. \end{aligned}$$

4 Global stability problem of space-independent steady states

This section is spent on exploring the global stability problem of the space-independent equilibria of the following system:

$$\begin{aligned} \left\{ \begin{array}{ll} \displaystyle \frac{\partial S_m}{\partial t} - D_m\Delta S_m = \mu -\beta S_mI_h-d_m S_m,&{}\quad x\in \Omega ,\ t>0, \\ \displaystyle \frac{\partial i_m(t,a,x)}{\partial t}+ \frac{\partial i_m(t,a,x)}{\partial a} - D_m\Delta i_m(t,a,x) =-d_m i_m(t,a,x),&{}\quad x\in \Omega ,\ t>0,\ a\ge 0, \\ \displaystyle \frac{\partial I_h}{\partial t} - D_h\Delta I_h(t,x) =H\int \limits _0^\infty \beta _1(a)i_m(t,a,x)\mathrm{d}a-(d_h+\rho )I_h,&{}\quad x\in \Omega ,\ t>0, \\ \displaystyle i_m(t,0,x)=\beta S_mI_h,&{}\quad x\in \Omega ,\ t>0,\\ \displaystyle \frac{\partial S_m}{\partial n}=\frac{\partial i_m(t,a,x)}{\partial n}=\frac{\partial I_h}{\partial n}=0, &{}\quad x\in \partial \Omega ,\ t>0. \end{array}\right. \end{aligned}$$
(4.1)

Note that the assertions of model (1.4) in previous sections, including the existence and uniqueness of the solution, existence of global attractor, and definition of the basic reproduction number, still hold for (4.1). Obviously, (4.1) has a space-independent disease-free equilibrium \({\tilde{E}}_0 = (S_m^0,0,0)\), where \(S_m^0=\frac{\mu }{d_m}\). Similar to (1.3) and (3.2), \(i_m(t,a,x)\) takes the following form:

$$\begin{aligned} i_m(t,a,x)= \left\{ \begin{array}{ll} \displaystyle \int \limits _{\Omega } \Gamma _1(a,x,y) {\mathcal {B}}(t-a,y)\mathrm{d}y\Pi (a), &{}\quad t-a>0,\ x\in \Omega ,\\ \displaystyle \int \limits _{\Omega } \Gamma _1(t,x,y)\phi _2(a-t,y)\mathrm{d}y\frac{\Pi (a)}{\Pi (a-t)},&{}\quad a-t\ge 0,\ x\in \Omega , \end{array}\right. \end{aligned}$$
(4.2)

with \({\mathcal {B}}(t-a,y):=i_m(t-a,0,y)=\beta S_m(t-a,y)I_h(t-a,y)\).

Lemma 3.1, together with Krein–Rutman theorem (see, e.g., [2, Theorem 3.2]), implies that the basic reproduction number is the only positive eigenvalue of \({\mathcal {L}}\), with a positive eigenvector. Substituting \(\phi (x)\equiv \mathbf{1 } > 0\) in (3.6) and using \(\int \limits _{\Omega }\Gamma _i(\cdot ,x,y)\mathrm{d}y=1 (i=1,2)\), one gets

$$\begin{aligned} {\mathcal {L}}[\mathbf{1 }]=\beta \frac{\mu }{d_m}H\int \limits _0^\infty e^{-(d_h+\rho )s}\int \limits _0^\infty \beta _1(a)\Pi (a)\mathrm{d}a\mathrm{d}s[\mathbf{1 }]. \end{aligned}$$

In this setting, the basic reproduction number of (4.1), \([\mathfrak {R}_0]\), is

$$\begin{aligned} {[}\mathfrak {R}_0]=\beta \frac{\mu }{d_m}H\int \limits _0^\infty e^{-(d_h+\rho )s}ds\int \limits _0^\infty \beta _1(a)\Pi (a)\mathrm{d}a. \end{aligned}$$
(4.3)

Denote by \(E^* = (S_m^*, i_m^*(\cdot ), I_h^*)\) the space-independent endemic equilibrium of (4.1), if it exists. Then, it satisfies

$$\begin{aligned} \left\{ \begin{array}{ll} \displaystyle \mu -\beta S_m^*I_h^*-d_m S_m^*=0, \\ \displaystyle \frac{\partial i_m^*(a)}{\partial a} = -d_m i_m^*(a), \\ \displaystyle i_m^*(0)=\beta S_m^*I_h^*,\\ \displaystyle H\int \limits _0^\infty \beta _1(a)i_m^*(a)\mathrm{d}a-(d_h+\rho )I_h^*=0. \end{array}\right. \end{aligned}$$
(4.4)

Direct calculation gives

$$\begin{aligned} S_m^*=\frac{d_h +\rho }{\beta HK},\ i^*_m(a)=i^*_m(0)\Pi (a),\ \text{ and }\ I_h^*=\frac{K Hi_m^*(0)}{d_h+\rho }, \end{aligned}$$
(4.5)

where \(K=\int \limits _0^\infty \beta _1(a)\Pi (a)\mathrm{d}a\) and \(i_m^*(0)=\frac{d_m (d_h+\rho )([\mathfrak {R}_0]-1)}{HK\beta }\).

Obviously, we have the following assertion.

Lemma 4.1

If \([\mathfrak {R}_0] > 1\), (4.1) has a space-independent endemic equilibrium \(E^*=(S^*_m, i^*_m(0)\Pi (a), I_h^*)\), which is unique and defined by (4.5).

4.1 Local dynamics

This subsection is spent on the local stability problem of \({\tilde{E}}_0\) and \(E^*\). To this end, letting \(E=({\bar{S}}_m, {\bar{i}}_m(\cdot ), {\bar{I}}_h)\) be any equilibrium of (4.1). We linearize (4.1) around E

$$\begin{aligned} \left\{ \begin{array}{ll} \displaystyle \frac{\partial S_m}{\partial t}- D_m\Delta S_m =-\beta {\bar{S}}_mI_h(t,x)-\beta S_m(t,x){\bar{I}}_h-d_m S_m(t,x), \\ \displaystyle \frac{\partial i_m(t,a,x)}{\partial t}+ \frac{\partial i_m(t,a,x)}{\partial a} - D_m\Delta i_m(t,a,x) =-d_m i_m(t,a,x), \\ \displaystyle i_m(t,0,x)=\beta {\bar{S}}_mI_h+\beta S_m{\bar{I}}_h,\\ \displaystyle \frac{\partial I_h}{\partial t} - D_h\Delta I_h= H\int \limits _0^\infty \beta _1(a)i_m(t,a,x)\mathrm{d}a - (d_h+\rho )I_h, \\ \displaystyle \frac{\partial S_m}{\partial n}=\frac{\partial i_m(t,a,x)}{\partial n}=\frac{\partial I_h}{\partial n}=0. \end{array}\right. \end{aligned}$$
(4.6)

We proceed to determine the characteristic equation of E. Since the linear system contains Laplacian term, we introduce the related theory from [6]. Denote by \(\lambda _j (j= 1,2,\ldots )\) the eigenvalues of operator \(-\Delta \) on a bounded set \(\Omega \) with boundary condition (1.6), that is, \(\Delta \nu (x)= -\zeta _i \nu (x)\). Hence,

$$\begin{aligned} 0=\lambda _0<\lambda _1<\lambda _2<\cdot \cdot \cdot , \end{aligned}$$

corresponding to which there is the space of eigenfunctions in \(C^1(\Omega )\), denoted by \(E(\lambda _i)\).

Denote by \(\{\phi _{ij}|\ j =1,2,\ldots ,\dim E(\lambda _i)\}\) the orthogonal basis of \(E(\lambda _i)\). Further, let \({\mathbb {X}}_{ij} =\{c\phi _{ij}|\ c\in {\mathbb {R}}^3\}\) such that

$$\begin{aligned} {\tilde{\mathbb {X}}} = \bigoplus \limits _{i=0}^{\infty }{\mathbb {X}}_i,\ \text{ where }\ {\mathbb {X}}_i = \bigoplus _{j=1}^{\dim E(\lambda _i)}{\mathbb {X}}_{ij}. \end{aligned}$$

It is well known that the parabolic problem \(\frac{\partial u(t,x)}{\partial t}=\Delta u(t,x)\) with \(\frac{\partial u(t,x)}{\partial n} = 0\), admits the exponential solution \(u(t,x)= e^{\eta t}\nu (x)\), where \(\nu (x)\in {\mathbb {X}}_i\). Substituting \((S_m, i_m(t,a,x), I_h)=e^{\eta t}(\psi (x), \xi (a,x), \phi (x))\) in (4.6) gets

$$\begin{aligned} \left\{ \begin{array}{ll} \displaystyle \eta \psi (x)+D_m\lambda _i \psi (x)=-\beta {\bar{S}}_m\phi (x)-\beta {\bar{I}}_h\psi (x)-d_m \psi (x),\\ \displaystyle \eta \xi (a,x)+ \frac{\partial \xi (a,x)}{\partial a} +D_m\lambda _i\xi (a,x)= -d_m \xi (a,x),\\ \displaystyle \xi (0,x) = \beta {\bar{S}}_m\phi (x) + \beta {\bar{I}}_h\psi (x),\\ \displaystyle \eta \phi (x) + D_h\lambda _i \phi (x) =H\int \limits _0^\infty \beta _1(a)\xi (a,x)\mathrm{d}a - (d_h+\rho )\phi (x).\\ \end{array}\right. \end{aligned}$$
(4.7)

Combined with the second and third equation of (4.7), we get that

$$\begin{aligned} \xi (a,x)=\xi (0,x)e^{-\eta a}{\tilde{\Pi }}(a),\ \text{ where }\ {\tilde{\Pi }}(a)=e^{-D_m\lambda _ia}\Pi (a). \end{aligned}$$

We prove the following claim.

Claim

\(\eta \ne -(D_m\lambda _i+d_m)\) and \(\eta \ne -(D_h\lambda _i+d_h+\rho )\).

In fact, if \(\eta =-(D_m\lambda _i+d_m)\), together with the first equation of (4.7), implies that \( \xi (0,x)=0\). Hence, from the fourth equation of (4.7), we directly have \(\eta =-(D_h\lambda _i+d_h+\rho )\), which results in a contradiction. \(\eta \ne -(D_h\lambda _i+d_h+\rho )<0\) can be proved in a similar way.

This claim together with the first and fourth equation of (4.7) implies that

$$\begin{aligned} \psi (x)=-\frac{\xi (0,x)}{\eta +D_m\lambda _i+d_m}\ \text{ and }\ \phi (x)=\frac{\xi (0,x)f(\eta )}{\eta +D_h\lambda _i+d_h+\rho }, \end{aligned}$$
(4.8)

where \(f(\eta )=H\int \limits _0^\infty \beta _1(a)e^{-\eta a}{\tilde{\Pi }}(a)da\). Plugging (4.8) into the third equation of (4.7) and canceling \(\xi (0,x)\), we get

$$\begin{aligned} \left( 1+\frac{1}{\eta +D_m\lambda _i+d_m}\beta {\bar{I}}_h\right) =\beta {\bar{S}}_m\left( \frac{f(\eta )}{\eta +D_h\lambda _i+d_h+\rho }\right) . \end{aligned}$$
(4.9)

Hence, (4.9) indeed admits a principal eigenvalue \(\eta ^*\) (see, for example, [17, Lemma 2.2]).

The following result indicates that both \({\tilde{E}}_0\) and \(E^*\) are locally asymptotically stable (LAS) under threshold conditions, which is achieved by studying the distribution of characteristic roots of (4.9).

Theorem 4.2

Let \([\mathfrak {R}_0]\) be defined by (4.3).

  1. (i)

    If \([\mathfrak {R}_0]<1\), then \({\tilde{E}}_0\) is LAS;

  2. (ii)

    If \([\mathfrak {R}_0]>1\), then \(E^*\) is LAS.

Proof

Let us first prove (i). For \({\tilde{E}}_0\), that is, \({\bar{S}}_m=\frac{\mu }{d_m}\) and \({\bar{I}}_h=0\), (4.9) can be reduced to

$$\begin{aligned} 1=\beta \frac{\mu }{d_m}\frac{ f(\eta )}{\eta +D_h\lambda _i+d_h+\rho }. \end{aligned}$$
(4.10)

Suppose, by contradiction, (4.10) admits a real root \(\eta >0\). We estimate (4.10)

$$\begin{aligned} 1=\left| \ \frac{\mu }{d_m}\frac{\beta f(\eta )}{\eta +D_h\lambda _i+d_h+\rho }\right| \le \beta \frac{\mu }{d_m(d_h+\rho )}H\int \limits _0^\infty \beta _1(a)\Pi (a)\mathrm{d}a=[\mathfrak {R}_0], \end{aligned}$$

which leads to a contradiction with \([\mathfrak {R}_0]<1\). Hence, all the real roots of (4.10) are negative. On the other hand, if (4.10) admits a pair of complex roots, denote by \(\eta =c\pm di\) with \(c\ge 0\) and \(d>0\), we then have

$$\begin{aligned} 1&=\beta \frac{\mu }{d_m}H\left( \frac{(c+D_h\lambda _i+d_h+\rho )\int \limits _0^\infty \beta _1(a)e^{-c a}\cos (da){\tilde{\Pi }}(a)da-d\int \limits _0^\infty \beta _1(a)e^{-ca}\sin (da){\tilde{\Pi }}(a)da}{(c+D_h\lambda _i+d_h+\rho )^2+d^2}\right) \le [\mathfrak {R}_0], \end{aligned}$$
(4.11)

again a contradiction with \([\mathfrak {R}_0]<1\). This proves (i).

We next prove (ii). For \(E^*\), that is, \({\bar{S}}_m=S_m^*\) and \({\bar{I}}_h=I_h^*\), (4.9) can be reduced to

$$\begin{aligned} \left( 1+\frac{1}{\eta +D_m\lambda _i+d_m}\beta I^*_h\right) =\beta S^*_m\left( \frac{f(\eta )}{\eta +D_h\lambda _i+d_h+\rho }\right) . \end{aligned}$$
(4.12)

With \(\eta =c+di\) with \(c\ge 0\), we estimate the left-hand side (4.12),

$$\begin{aligned} \left| 1+\frac{1}{\eta +D_m\lambda _i+d_m}\beta I_h^*\right| =\frac{\sqrt{(c+D_m\lambda _i+d_m+\beta I_h^*)^2+d^2}}{\sqrt{(c+D_m\lambda _i+d_m)^2+d^2}}>1. \end{aligned}$$
(4.13)

The right-hand side (4.12) can be estimated as

$$\begin{aligned} \beta S_m^*\left| \frac{f(\eta )}{\eta +D_h\lambda _i+d_h+\rho }\right| = \beta S_m^* \frac{|f(\eta )|}{|\eta +D_h\lambda _i+d_h+\rho |}< \beta S_m^*\frac{HK}{d_h+\rho }=1, \end{aligned}$$
(4.14)

which is a contradiction. Consequently, if \([\mathfrak {R}_0]>1\), all roots of (4.12) have negative real parts. This proves (ii). \(\square \)

4.2 Global dynamics

This subsection is spent on the global stability problem of \({\tilde{E}}_0\) and \(E^*\). Combined with local asymptotic stability (in subsection (4.1)) and global attractivity of equilibria, we shall confirm that both \({\tilde{E}}_0\) and \(E^*\) are globally asymptotically stable (GAS).

Theorem 4.3

If \([\mathfrak {R}_0]<1\) and \(\phi =(\phi _1, \phi _2,\phi _3)\in {\mathcal {D}}\), then \({\tilde{E}}_0\) is GAS.

Proof

It is well known that function

$$\begin{aligned} G(u,v) = u - v - v\ln \frac{u}{v}\ge 0,\ \text{ for }\ u,v\in {\mathbb {R}}_+, \end{aligned}$$
(4.15)

and \(G(u,u)=0\). By using this function, we define

$$\begin{aligned} {\mathbb {L}}[S_m, i_m, I_h](t)=\int \limits _\Omega [{\mathbb {L}}_{S_m}(t,x)+{\mathbb {L}}_{i_m} (t,x)+{\mathbb {L}}_{I_h}(t,x)]\mathrm{d}x, \end{aligned}$$

where \({\mathbb {L}}_{S_m}=G(S_m(t,x), S_m^0)\), \({\mathbb {L}}_{i_m}=\int \limits _0^\infty \Theta _1(a)i_m(t,a,x)\mathrm{d}a\), \({\mathbb {L}}_{I_h}=\frac{S_m^0\beta }{d_h+\rho }I_h(t,x)\) and \(\Theta _1(a)\in L_1(\infty )\) will be determined later.

We take the derivative of \({\mathbb {L}}_{S_m}\),

$$\begin{aligned} \frac{\partial {\mathbb {L}}_{S_m}}{\partial t}= D_m\left( 1-\frac{S_m^0}{S_m}\right) \Delta S_m- d_m\frac{(S_m-S_m^0)^2}{S_m} +S_m^0\beta I_h- \beta I_hS_m. \end{aligned}$$
(4.16)

With the help of (4.2), we rewrite \({\mathbb {L}}_{i_m}\)

$$\begin{aligned} {\mathbb {L}}_{i_m}=&\ \int \limits _0^t \Theta _1(t-a)\int \limits _\Omega \Gamma _1(t-a,x,y){\mathcal {B}}(a,y) \mathrm{d}y\Pi (t-a)\mathrm{d}a+\int \limits _0^\infty \Theta _1(a+t)\int \limits _\Omega \Gamma _1(t,x,y)\phi _2(a,y)\mathrm{d}y\Pi (t)\mathrm{d}a. \end{aligned}$$
(4.17)

We take the derivative of \({\mathbb {L}}_{i_m}\)

$$\begin{aligned} \frac{\partial {\mathbb {L}}_{i_m}}{\partial t}&= \Theta _1(0)\int \limits _\Omega \Gamma _1(0,x,y){\mathcal {B}}(t,y)\mathrm{d}y+\int \limits _0^t \frac{\mathrm{d}}{\mathrm{d}t}\Theta _1(t-a)\int \limits _\Omega \Gamma _1(t-a,x,y){\mathcal {B}}(a,y)\mathrm{d}y\Pi (t-a)\mathrm{d}a\nonumber \\&\quad +\int \limits _0^t \Theta _1(t-a)\int \limits _\Omega \frac{\partial }{\partial t}\Gamma _1(t-a,x,y){\mathcal {B}}(a,y)\mathrm{d}y\Pi (t-a)\mathrm{d}a\nonumber \\&\quad -d_m\int \limits _0^t \Theta _1(t-a)\int \limits _\Omega \Gamma _1(t-a,x,y){\mathcal {B}}(a,y)\mathrm{d}y\Pi (t-a)\mathrm{d}a\\&\quad +\int \limits _0^\infty \frac{\mathrm{d}}{\mathrm{d}t}\Theta _1(a+t)\int \limits _\Omega \Gamma _1(t,x,y)\phi _2(a,y)\mathrm{d}y\Pi (t)\mathrm{d}a\nonumber \\&\quad +\int \limits _0^\infty \Theta _1(a+t)\int \limits _\Omega \frac{\partial }{\partial t}\Gamma _1(t,x,y)\phi _2(a,y)\mathrm{d}y\Pi (t)\mathrm{d}a\nonumber \\&\quad -d_m\int \limits _0^\infty \Theta _1(a+t)\int \limits _\Omega \Gamma _1(t,x,y)\phi _2(a,y)\mathrm{d}y\Pi (t)\mathrm{d}a\nonumber . \end{aligned}$$
(4.18)

Collecting the terms with the form of (4.17), together with the fact that \(\frac{\partial \Gamma _1(t,x,y)}{\partial t}=D_m\Delta \Gamma _1(t,x,y)\) a.e. for \(x\in \Omega \), we get

$$\begin{aligned} \frac{\partial {\mathbb {L}}_{i_m}}{\partial t}&= \Theta _1(0){\mathcal {B}}(t,x)+\int \limits _0^\infty \left[ \frac{\mathrm{d}}{\mathrm{d}a}\Theta _1(a)-(d_m-D_m\Delta )\Psi (a)\right] i_m(t,a,x)\mathrm{d}a. \end{aligned}$$
(4.19)

We take the derivative of \({\mathbb {L}}_{I_h}(t,x)\)

$$\begin{aligned} \frac{\partial {\mathbb {L}}_{I_h}}{\partial t}=\beta \frac{S_m^0}{d_h+\rho }D_h\Delta I_h+\beta \frac{S_m^0}{d_h+\rho }H\int \limits _0^\infty \beta _1(a)i_m(t,a,x)\text {d}a-S_m^0 \beta I_h. \end{aligned}$$
(4.20)

With the help of (4.16), (4.18) and (4.20), we get the derivation of \({\mathbb {L}}(t)\)

$$\begin{aligned} \begin{aligned} \frac{\partial {\mathbb {L}}(t)}{\partial t}&= \int \limits _\Omega D_m\left( 1-\frac{S_m^0}{S_m}\right) \Delta S_m\mathrm{d}x -d_m\int \limits _\Omega \frac{(S_m-S^0)^2}{S_m}\mathrm{d}x+\int \limits _\Omega (\Theta _1(0)-1){\mathcal {B}}(t,x)\mathrm{d}x\\&\quad +\int \limits _\Omega \int \limits _0^\infty \left[ \beta \frac{S_m^0}{d_h +\rho }H\beta _1(a)+\frac{\mathrm{d}}{\mathrm{d}a}\Theta _1(a)-(d_m-D_m\Delta )\Theta _1(a)\right] i_m(t,a,x)\mathrm{d}a\mathrm{d}x. \end{aligned} \end{aligned}$$
(4.21)

Setting

$$\begin{aligned} \Theta _1(a)=\frac{1}{\Pi (a)}\int \limits _a^\infty \beta \frac{S_m^0 }{d_h+\rho }H\beta _1(\theta )\Pi (\theta )d\theta . \end{aligned}$$

Then, it satisfies

$$\begin{aligned} \left\{ \begin{array}{l} \displaystyle \frac{\mathrm{d}}{\mathrm{d}a}\Theta _1(a)=-\beta \frac{S_m^0 }{d_h+\rho }H\beta _1(a)+d_m\Theta _1(a),\\ \displaystyle \Theta _1(0)=[\mathfrak {R}_0], \end{array} \right. \end{aligned}$$

Hence, (4.21) becomes

$$\begin{aligned} \frac{\partial {\mathbb {L}}(t)}{\partial t} = - D_m\int \limits _\Omega \frac{|\nabla S_m|^2}{S^2_m}\mathrm{d}x - \int \limits _\Omega d_m\frac{(S_m-S^0)^2}{S_m}\mathrm{d}x + ([\mathfrak {R}_0]-1)\int \limits _\Omega {\mathcal {B}}(t,x)\mathrm{d}x. \end{aligned}$$

Hence, the global attractiveness of \({\tilde{E}}_0\) holds in \({\mathcal {D}}\) when \([\mathfrak {R}_0] < 1\) (see, e.g., [31, Theorem 4.2 ]). This together with Theorem 4.2 indicates that the \({\tilde{E}}_0\) is GAS. This proves Theorem 4.3. \(\square \)

Define

$$\begin{aligned} {\mathcal {D}}_0 = \left\{ (\phi _1, \phi _2, \phi _3)\in \mathbb {{\tilde{X}}}_+|\ \phi _3>0,\text{ for } \text{ some } \ x\in {\overline{\Omega }}\right\} . \end{aligned}$$
(4.22)

In the following, we pay attention to the persistence problem of system (4.1) when \([\mathfrak {R}_0] > 1\). We first give the following lemma.

Lemma 4.4

Let \({\mathcal {D}}_0\) be defined by (4.22). If \([\mathfrak {R}_0] > 1\), then there exists \(\epsilon _1>0\) such that

$$\begin{aligned} \lim \sup _{t\rightarrow \infty }|{\mathcal {B}}(t,\cdot )|_{X} > \epsilon _1,\ \forall \phi \in {\mathcal {D}}_0. \end{aligned}$$

Proof

Due to \([\mathfrak {R}_0] > 1\), we can select \(\epsilon _1 > 0\) such that

$$\begin{aligned} \beta \frac{(\mu - \epsilon _1)}{d_m}H\int \limits _0^\infty e^{-(d_h+\rho )s}ds\int \limits _0^\infty \beta _1(a)\Pi (a)\mathrm{d}a> 1. \end{aligned}$$
(4.23)

We will prove the assertion by contradiction. Assume, by contradiction, that there exists \(T_1>0\) such that for all \(t\ge T_1\) and \(x\in \Omega \), we have \({\mathcal {B}}(t,x)\le \epsilon _1\). By the inequality in (4.23), choosing \(T_2>T_1\) (denote by \(h=T_2-T_1\)) and \(\lambda >0\) small enough that

$$\begin{aligned} \bigodot =\beta \frac{(\mu - \epsilon _1)}{d_m}H(1 - e^{-d_m h})\int \limits _0^\infty e^{-(d_h+\rho )s}e^{-\lambda s}ds\int \limits _0^\infty \beta _1(a)\Pi (a)e^{-\lambda a}da> 1, \end{aligned}$$
(4.24)

Note that \(S_m\) satisfies

$$\begin{aligned} \left\{ \begin{array}{ll} \displaystyle \frac{\partial S_m}{\partial t} -D_m\Delta S_m \ge \mu -\epsilon _1-d_m S_m,&{}\quad x\in \Omega ,\ t\ge T_2,\\ \displaystyle \frac{\partial S_m}{\partial n}=0,&{}\quad x\in \partial \Omega ,\ t\ge T_2. \end{array}\right. \end{aligned}$$
(4.25)

Hence, direct integrating (4.25) yields

$$\begin{aligned} S_m&\ge \frac{\mu -\epsilon _1}{d_m}(1-e^{-d_m (t-T_1)})\ge \frac{\mu -\epsilon _1}{d_m}(1-e^{-d_m h}),\ \forall t\ge T_2. \end{aligned}$$

Further from (2.4),

$$\begin{aligned} I_h&\ge H\int \limits _0^t e^{-(d_h+\rho ) (t-s)}\int \limits _\Omega \Gamma _2(t-s,x,y)\int \limits _0^\infty \beta _1(a)i_m(s,a,y)\mathrm{d}a\mathrm{d}y\mathrm{d}s\\&\ge H\int \limits _0^t e^{-(d_h +\rho )s}\int \limits _\Omega \Gamma _2(s,x,y) \int \limits _0^{t-s}\beta _1(a)\Pi (a)\int \limits _\Omega \Gamma _1(a,y,z){\mathcal {B}}(t-s-a,z)\mathrm{d}z\mathrm{d}a\mathrm{d}y\mathrm{d}s. \end{aligned}$$

Combined with the above two inequalities, we have

$$\begin{aligned}&{\mathcal {B}}(t,x)\nonumber \\&\quad \ge \beta \frac{\mu - \epsilon _1}{d_m} H(1-e^{-d_m h})\int \limits _0^t e^{-(d_h +\rho )s}\int \limits _\Omega \Gamma _2(s,x,y)\int \limits _0^{t-s}\beta _1(a) \Pi (a)\int \limits _\Omega \Gamma _1(a,y,z){\mathcal {B}}(t-s-a,z)\mathrm{d}z\mathrm{d}a\mathrm{d}y\mathrm{d}s. \end{aligned}$$
(4.26)

Note that \(\int \limits _0^\infty e^{-\lambda t}{\mathcal {B}}(t,x)\mathrm{d}t < \infty \) for all \(x\in \Omega \). Choose \({\widetilde{x}}\in \Omega \) that \({\mathcal {B}}(t,{\widetilde{x}}) = \min _{x\in \Omega }{\mathcal {B}}(t,x)\). By (4.26), we have

$$\begin{aligned} \int \limits _0^\infty e^{-\lambda t}{\mathcal {B}}(t,{\widetilde{x}})\mathrm{d}t&\ge \beta \frac{\mu - \epsilon _1}{d_m}H(1-e^{-d_m h})\int \limits _0^\infty e^{-\lambda t}\int \limits _0^t e^{-(d_h +\rho )s}\int \limits _\Omega \Gamma _2(s,{\widetilde{x}},y)\int \limits _0^{t-s}\beta _1(a)\Pi (a)\\&\quad \times \int \limits _\Omega \Gamma _1(a,y,z){\mathcal {B}}(t-s-a,z)\mathrm{d}z\mathrm{d}a\mathrm{d}y\mathrm{d}s\mathrm{d}t. \end{aligned}$$

Interchanging the order of integration (t and s) and letting \(t-s={\bar{t}}\) (for simplicity, still denote \({\bar{t}}\) by t) yield

$$\begin{aligned} \int \limits _0^\infty e^{-\lambda t}{\mathcal {B}}(t,{\widetilde{x}})\mathrm{d}t&\ge \beta \frac{\mu - \epsilon _1}{d_m}H(1-e^{-d_m h})\int \limits _0^\infty e^{-(d_h +\rho )s}\int \limits _s^{\infty }e^{-\lambda t} \int \limits _\Omega \Gamma _2(s,{\widetilde{x}},y)\int \limits _0^{t-s}\beta _1(a)\Pi (a)\\&\quad \times \int \limits _\Omega \Gamma _1(a,y,z){\mathcal {B}}(t-s-a,z)\mathrm{d}z\mathrm{d}a\mathrm{d}y\mathrm{d}t\mathrm{d}s\\&=\beta \frac{\mu - \epsilon _1}{d_m}H(1-e^{-d_m T_1})\int \limits _0^\infty e^{-(d_h+\rho )s}e^{-\lambda s}\int \limits _0^\infty e^{-\lambda t} \int \limits _\Omega \Gamma _2(s,{\widetilde{x}},y)\int \limits _0^{t}\beta _1(a)\Pi (a)\\&\quad \times \int \limits _\Omega \Gamma _1(a,y,z) {\mathcal {B}}(t-a,z)\mathrm{d}z\mathrm{d}a\mathrm{d}y\mathrm{d}t\mathrm{d}s. \end{aligned}$$

Interchanging the order of integration (t and a) and letting \(t-a={\bar{t}}\) (for simplicity, still denote \({\bar{t}}\) by t) yield

$$\begin{aligned} \int \limits _0^\infty e^{-\lambda t}B(t,{\widetilde{x}})\mathrm{d}t&\ge \beta \frac{\mu - \epsilon _1}{d_m}H(1-e^{-d_m h})\int \limits _0^\infty e^{-(d_h+\rho )s}e^{-\lambda s}\int \limits _0^\infty \beta _1(a)\Pi (a)\int \limits _\Omega \Gamma _2(s,{\widetilde{x}},y)\int \limits _a^{\infty }e^{-\lambda t}\\&\quad \times \int \limits _\Omega \Gamma _1(a,y,z) {\mathcal {B}}(t-a,z)\mathrm{d}z\mathrm{d}t\mathrm{d}y\mathrm{d}a\mathrm{d}s\\&=\beta \frac{\mu - \epsilon _1}{d_m}H(1-e^{-d_m h})\int \limits _0^\infty e^{-(d_h+\rho )s}e^{-\lambda s}\int \limits _0^\infty \beta _1(a)\Pi (a)e^{-\lambda a}\int \limits _\Omega \Gamma _2(s,{\widetilde{x}},y)\int \limits _0^{\infty }e^{-\lambda t}\\&\quad \times \int \limits _\Omega \Gamma _1(a,y,z) {\mathcal {B}}(t,z)\mathrm{d}z\mathrm{d}t\mathrm{d}y\mathrm{d}a\mathrm{d}s. \end{aligned}$$

Consequently, we have

$$\begin{aligned} \int \limits _0^\infty e^{-\lambda t}{\mathcal {B}}(t,{\widetilde{x}})\mathrm{d}t&\ge \beta \frac{\mu - \epsilon _1}{d_m}H(1-e^{-d_m h})\int \limits _0^\infty e^{-(d_h+\rho )s}e^{-\lambda s}\int \limits _0^\infty \beta _1(a)\Pi (a)e^{-\lambda a}\int \limits _\Omega \Gamma _2(s,{\widetilde{x}},y)\nonumber \\&\quad \times \int \limits _\Omega \Gamma _1(a,y,z)\int \limits _0^{\infty }e^{-\lambda t}{\mathcal {B}}(t,z)\mathrm{d}t\mathrm{d}z\mathrm{d}y\mathrm{d}a\mathrm{d}s\nonumber \\&\ge \bigodot \int \limits _0^\infty e^{-\lambda t}{\mathcal {B}}(t,{\widetilde{x}})\mathrm{d}t, \end{aligned}$$

which results in a contradiction with (4.24). This proves Lemma 4.4. \(\square \)

Lemma 4.5

If \([\mathfrak {R}_0] > 1\), then \(\liminf _{t\rightarrow \infty }|{\mathcal {B}}(t,\cdot )|_{\mathbb {X}} > \epsilon _1\), \(\forall \phi \in {\mathcal {D}}_0\).

Proof

Suppose, by contradiction, that \(\lim \inf _{t\rightarrow \infty }|{\mathcal {B}}(t,\cdot )|_{\mathbb {X}} < \epsilon _1\). This together with Lemma 4.4 allows us to choose increasing sequences \(\{t_{1j}\}_{j=1}^\infty \), \(\{t_{2j}\}_{j=1}^\infty \), \(\{t_{3j}\}_{j=1}^\infty \) and a decreasing sequence \(\{t_{4j}\}_{j=1}^\infty \) such that \(t_{1j}>t_{2j}>t_{3j}\), \(\lim \limits _{j\rightarrow +\infty }t_{4j}=0\) that

$$\begin{aligned} \left\{ \begin{array}{ll} |{\mathcal {B}}(t,\cdot )|_{\mathbb {X}}>\epsilon _1, &{}\quad t=t_{3j}; \\ |{\mathcal {B}}(t,\cdot )|_{\mathbb {X}} =\epsilon _1, &{}\quad t=t_{2j}; \\ |{\mathcal {B}}(t,\cdot )|_{\mathbb {X}}<t_{4j}<\epsilon _1, &{}\quad t=t_{1j}; \end{array} \right. \end{aligned}$$
(4.27)

It is easy to see that \(|{\mathcal {B}}(t,\cdot )|_{\mathbb {X}}<\epsilon _1\), for all \(t\in (t_{2j},t_{1j})\). When \(t=t_{2j}\), we denote by \(\{{S_m}_j\}_{j=1}^\infty \), \(\{{\mathcal {B}}_j\}_{j=1}^\infty \) and \(\{{I_h}_j\}_{j=1}^\infty \) be functional sequences in \({\mathbb {X}}\), which satisfies \(S_{mj}=S_m(t_{2j},\cdot )\in {\mathbb {X}}\), \({\mathcal {B}}_j= {\mathcal {B}}(t_{2j},\cdot )\in {\mathbb {X}}\) and \(I_{hj}=I_h(t_{2j},\cdot )\in {\mathbb {X}}\), respectively. From the expressions in (2.2)-(2.4), we can conclude that there exist \((\breve{S}_m, \breve{{\mathcal {B}}}, \breve{I}_h)\in {\mathbb {X}}^+\times {\mathbb {Y}}^+\times {\mathbb {X}}^+\) such that

$$\begin{aligned} \lim \limits _{j\rightarrow +\infty } S_{mj} = \breve{S}_m,\ \lim \limits _{j\rightarrow +\infty } {\mathcal {B}}_{j} = \breve{{\mathcal {B}}},\ \lim \limits _{j\rightarrow +\infty } I_{hj} = \breve{I}_h. \end{aligned}$$

In this setting, for all \(a\ge 0\) and \(x\in \Omega \), denote by \((S_m^\diamond ,i_m^\diamond (t,a, x),I_h^\diamond )\) the solution of (4.1) with

$$\begin{aligned} \phi _{1}(x)=\breve{S}_m(x),\ \phi _{2}(a,x)=\int \limits _\Omega \Gamma _2(a,x,y)\breve{{\mathcal {B}}}(y)\mathrm{d}y\Pi (a),\ \phi _{3}(x)=\breve{I}_h(x). \end{aligned}$$

Here, the choice of \(\phi _2\) is based on (1.3). According to Lemma 4.4, there exists \(\tau ^*>0\) and \(\ell >0\) such that

$$\begin{aligned} |{\mathcal {B}}^\diamond (\tau ^*,\cdot )|_{\mathbb {X}}> \epsilon _1,\ \text{ and }\ |{\mathcal {B}}^\diamond (t,\cdot )|_{\mathbb {X}} > \ell ,\ \forall t\in (0,\tau ^*). \end{aligned}$$

Denote by \({\mathcal {B}}_j^\diamond (t,\cdot ) = {\mathcal {B}}(t+\theta _j,\cdot )\). The above inequalities imply that for sufficiently large j that

$$\begin{aligned} \left\{ \begin{array}{ll} |{\mathcal {B}}_j^\diamond (t,\cdot )|_{\mathbb {X}}> \epsilon _1, &{}\quad t=\tau ^*; \\ |{\mathcal {B}}_j^\diamond (t,\cdot )|_{\mathbb {X}}>\ell >e_j, &{} \quad t\in (0,\tau ^*). \end{array} \right. \end{aligned}$$
(4.28)

Corresponding to this, however, by letting \({\widetilde{t}}_j:=t_{1j}-t_{2j}\), (4.27) indicates that

$$\begin{aligned} \left\{ \begin{array}{ll} |{\mathcal {B}}_j^\diamond (t,\cdot )|_{\mathbb {X}}< e_j< \epsilon _1, &{}\quad t={\widetilde{t}}_{j}; \\ |{\mathcal {B}}_j^\diamond (t,\cdot )|_{\mathbb {X}} < \epsilon _1, &{} \quad t\in (0,{\widetilde{t}}_j). \end{array} \right. \end{aligned}$$
(4.29)

For the cases where \(\tau ^*<{\widetilde{t}}_j\) and \(\tau ^*>{\widetilde{t}}_j\), it is easily seen that (4.28) and (4.29) contradict each other. This proves Lemma 4.5. \(\square \)

Combined with Lemmas 4.4 and 4.5, we directly have:

Theorem 4.6

If \([\mathfrak {R}_0] > 1\), system (4.1) is uniformly strongly persistent, that is, there exists \(\epsilon _0>0\) such that for \(x\in \Omega \) and \(a\ge 0\),

$$\begin{aligned} \liminf \limits _{t\rightarrow \infty ,x\in \Omega } U>\epsilon _0, \end{aligned}$$
(4.30)

where \(U=S_m, i_m(t,a,x), I_h\), respectively.

Proof

For all \(a\ge 0, x\in \Omega \), Lemma 4.5 implies that there exist \(\epsilon _1>0\) and \(t_2>0\) such that \(i_m(t,a,x)>\epsilon _1 \Pi (a):=\epsilon _{i_m}\) for all \(t > t_2\). By using this estimate, the \(I_h\) equation of (4.1) satisfies

$$\begin{aligned} \left\{ \begin{array}{ll} \displaystyle \frac{\partial I_h}{\partial t} > D_h\Delta I_h+ \epsilon _1HK - (d_h + \rho )I_h,&{}\quad x\in \Omega ,\ t\ge t_2,\\ \displaystyle \frac{\partial I_h}{\partial n}=0,&{}\quad x\in \partial \Omega , ~t\ge t_2. \end{array}\right. \end{aligned}$$

By a comparison principle, we have

$$\begin{aligned} \liminf \limits _{t\rightarrow \infty ,x\in \Omega } I_h(t,x)>\frac{\epsilon _1HK}{d_h + \rho }:=\epsilon _{I_h}. \end{aligned}$$

where \(\frac{\epsilon _2HK}{d_h + \rho }\) is the unique positive steady state of the following system

$$\begin{aligned} \left\{ \begin{array}{ll} \displaystyle \frac{\partial I_h}{\partial t} = D_h\Delta I_h + \epsilon _1HK - (d_h + \rho )I_h,&{}\quad x\in \Omega ,\ t\ge t_2,\\ \displaystyle \frac{\partial I_h}{\partial n}=0,&{}\quad x\in \partial \Omega ,~t\ge t_2. \end{array}\right. \end{aligned}$$

In view of \(S_m\) equation of (4.1) and (2.9), there exists \(t_3>0\) such that

$$\begin{aligned} \left\{ \begin{array}{ll} \displaystyle \frac{\partial S_m}{\partial t}\ge \ D_m\Delta S_m+\mu -(\beta M_{I_h} + d_m)S_m&{}\quad x\in \Omega ,\ t\ge t_3,\\ \displaystyle \frac{\partial S_m}{\partial n}=0,&{}\quad x\in \partial \Omega ,~t\ge t_3. \end{array}\right. \end{aligned}$$

Again, from comparison principle, we have

$$\begin{aligned} \liminf \limits _{t\rightarrow \infty ,x\in \Omega } S_m(t,x)>\frac{\mu }{\beta M_{I_h} + d_m}:=\epsilon _{S_m}. \end{aligned}$$

Consequently, the theorem is proved by letting \(\epsilon _0=\min \{\epsilon _{S_m}, \epsilon _{i_m}, \epsilon _{I_h}\}\). \(\square \)

Theorem 4.7

If \([\mathfrak {R}_0] > 1\), then \(E^*\) is GAS for \(\phi \in {\mathcal {D}}_0\).

Proof

Let us define

$$\begin{aligned} \overline{{\mathbb {L}}}[S_m, i_m, I_h](t)=\int \limits _\Omega [\overline{{\mathbb {L}}}_{S_m}(t,x) +\overline{{\mathbb {L}}}_{i_m}(t,x)+\overline{{\mathbb {L}}}_{I_h}(t,x)]\mathrm{d}x, \end{aligned}$$

where

$$\begin{aligned} \left\{ \begin{array}{ll} \overline{{\mathbb {L}}}_{S_m}=G[S_m,S_m^*],\\ \overline{{\mathbb {L}}}_{i_m}=\int \limits _0^\infty \Psi _1(a)G[i_m(t,a,x),i_m^*(a)]\mathrm{d}a,\\ \overline{{\mathbb {L}}}_{I_h}=\frac{S_m^*\beta }{d_h+\rho }G[I_h,I_h^*)], \end{array}\right. \end{aligned}$$

and \(G(u,v) = u - v - v\ln \frac{u}{v},\ \text{ for }\ u,v>0\), and \(\Psi _1(a)\) will be determined later. We next wish to calculate the derivative of \(\overline{{\mathbb {L}}}_{S_m}\), \(\overline{{\mathbb {L}}}_{i_m}\), and \(\overline{{\mathbb {L}}}_{I_h}\), respectively.

We derivative \(\overline{{\mathbb {L}}}_{S_m}\) along the solution of (4.1)

$$\begin{aligned} \frac{\partial \overline{{\mathbb {L}}}_{S_m}}{\partial t}&= \left( 1 - \frac{S_m^*}{S_m}\right) (D_m\Delta S_m + \mu - d_mS_m - i_m(t,0,x))\nonumber \\&= \left( 1 - \frac{S_m^*}{S_m}\right) D_m\Delta S_m -\frac{d_m}{S_m}(S_m - S_m^*)^2 + i_m^*(0) - i_m(t,0,x) \nonumber \\&\quad - i_m^*(0)\frac{S_m^*}{S_m} + i_m(t,0,x)\frac{S_m^*}{S_m}. \end{aligned}$$
(4.31)

Here, the equality, \(\mu = d_mS_m^*+i_m^*(0)\), is used. Using (4.2), rewriting \(\overline{{\mathbb {L}}}_{i_m}\) as

$$\begin{aligned} \overline{{\mathbb {L}}}_{i_m}&= \int \limits _0^t \Psi _1(t-r) G\left[ \int \limits _{\Omega }\Gamma _1(t-r,x,y)i_m(r, 0,y)\mathrm{d}y\Pi (t-r), i^*_m(t-r)\right] \text {d}r\\&\quad +\int \limits _0^\infty \Psi _1(t+r) G\left[ \int \limits _{\Omega }\Gamma _1(t,x,y)\phi _2(r,y)\mathrm{d}y\Pi (t), i^*_m(t+r)\right] \text {d}r. \end{aligned}$$

For convenience, let

$$\begin{aligned} u_1=\int \limits _{\Omega }\Gamma _1(t-r,x,y)i_m(r, 0,y)\mathrm{d}y\Pi (t-r),\ v_1=i^*_m(t-r) \end{aligned}$$

and

$$\begin{aligned} u_2=\int \limits _{\Omega }\Gamma _1(t,x,y)\phi _2(r,y)\mathrm{d}y\Pi (t),\ v_2=i^*_m(t+r). \end{aligned}$$

It follows from the relations \(i^*(a) = i^*(0)e^{-d_ma}\) and \(\Pi (0)=1\) that

$$\begin{aligned} \frac{\partial \overline{{\mathbb {L}}}_{i_m}}{\partial t}&= \Psi _1(0) G\left[ \int \limits _{\Omega }\Gamma _1(0,x,y)i_m(t, 0,y)\mathrm{d}y, i^*_m(0)\right] \\&\quad +\int \limits _0^t \frac{\mathrm{d}}{\mathrm{d}t}\Psi _1(t-r) G\left[ u_1, v_1\right] \text {d}r+\int \limits _0^\infty \frac{\mathrm{d}}{\mathrm{d}t}\Psi _1(t+r) G\left[ u_2, v_2\right] dr\\&\quad +\int \limits _0^t \Psi _1(t-r) \bigg \{\bigg [\int \limits _{\Omega }\frac{\partial }{\partial t}\Gamma _1(t-r,x,y)i_m(r,0,y)\mathrm{d}y\Pi (t-r)-d_mu_1\bigg ]\frac{\partial }{\partial u_1}G\left[ u_1, v_1\right] \\&\quad -d_mi^*_m(t-r)\frac{\partial }{\partial v_1}G\left[ u_1, v_1\right] \bigg \}dr\\&\quad +\int \limits _0^\infty \Psi _1(t+r) \bigg \{\bigg [\int \limits _{\Omega }\frac{\partial }{\partial t}\Gamma _1(t,x,y)\phi (r,y)\mathrm{d}y\Pi (t)-d_mu_2\bigg ]\frac{\partial }{\partial u_2} G\left[ u_2, v_2\right] \\&\quad -d_mi^*_m(t+r)\frac{\partial }{\partial v_2}G\left[ u_2, v_2\right] \bigg \}dr. \end{aligned}$$

With the equality that \(u\frac{\partial }{\partial u}G[u, v] + v\frac{\partial }{\partial v}G[u, v] =G[u, v]\), we have

$$\begin{aligned} \frac{\partial \overline{{\mathbb {L}}}_{i_m}}{\partial t}&= \Psi _1(0) G\left[ \int \limits _{\Omega }\Gamma _1(0,x,y)i_m(t, 0,y)\mathrm{d}y, i^*_m(0)\right] \\&\quad +\int \limits _0^\infty \bigg [\frac{\mathrm{d}}{\mathrm{d}a}\Psi _1(a)-d_m\Psi (a)\bigg ] G[i_m(t,a,x), i_m^*(a)]\mathrm{d}a\\&\quad +\int \limits _0^t \Psi _1(t-r)\bigg [\int \limits _{\Omega }\frac{\partial }{\partial t}\Gamma _1(t-r,x,y)i_m(r,0,y)\mathrm{d}y\Pi (t-r)\bigg ]\frac{\partial }{\partial u_1}G\left[ u_1, v_1\right] dr\\&\quad +\int \limits _0^\infty \Psi _1(t+r)\bigg [\int \limits _{\Omega }\frac{\partial }{\partial t}\Gamma _1(t,x,y)\phi (r,y)\mathrm{d}y\Pi (t)\bigg ]\frac{\partial }{\partial u_2}G\left[ u_2, v_2\right] dr. \end{aligned}$$

Note that \(\frac{\partial }{\partial t}\Gamma _1=D_m\Delta \Gamma _1\), \(\frac{\partial }{\partial u}G[u,v]=1-\frac{v}{u}\) and semigroup \((T(0)[\phi ])(x)=\int \limits _\Omega \Gamma _1(0,x,y)\phi (y)\mathrm{d}y\) is an unit semigroup. It follows that

$$\begin{aligned} \frac{\partial \overline{{\mathbb {L}}}_{i_m}}{\partial t}&= \Psi _1(0) G\left[ i_m(t, 0,x), i^*_m(0)\right] +\int \limits _0^\infty \bigg [\frac{\mathrm{d}}{\mathrm{d}a}\Psi _1(a)-d_m\Psi (a)\bigg ] G[i_m(t,a,x), i_m^*(a)]\text {d}a\nonumber \\&\quad +\int \limits _0^\infty \Psi _1(a)D_m\Delta i_m(t, a,x)\left[ 1-\frac{i^*_m(a)}{i_m(t,a,x)}\right] \text {d}a. \end{aligned}$$
(4.32)

Now we let

$$\begin{aligned} \Psi _1(a) = \frac{HS_m^*\beta }{d_h + \rho }\frac{1}{\Pi (a)}\int \limits _a^\infty \beta _1(\theta )e^{-d_m\theta }d\theta . \end{aligned}$$
(4.33)

Then, it satisfies

$$\begin{aligned} \left\{ \begin{array}{l} \displaystyle \frac{\mathrm{d}}{\mathrm{d}a}\Psi _1(a)=-\frac{HS_m^*\beta }{d_h + \rho } \beta _1(a)+d_m\Psi _1(a),\\ \displaystyle \Psi _1(0)=\beta \frac{HS_m^*}{d_h + \rho }K=S_m^*\frac{[\mathfrak {R}_0]}{S_m^0}=1. \end{array} \right. \end{aligned}$$

Hence, (4.32) becomes

$$\begin{aligned} \frac{\partial \overline{{\mathbb {L}}}_{i_m}}{\partial t}&= \frac{HS_m^*\beta }{d_h + \rho }\int \limits _0^\infty \beta _1(a)\left[ i_m^*(a) - i_m(t,a,x) + i_m^*(a)\ln \frac{i_m(t,a,x)}{i_m^*(a)}\right] \mathrm{d}a\\&\quad +G[i_m(t,0,x),i_m^*(0)]+ \int \limits _0^\infty \Psi _1(a)\left[ 1-\frac{i_m^*(a)}{i_m(t,a,x)}\right] D_m\Delta i_m(t,a,x)\mathrm{d}a\nonumber . \end{aligned}$$
(4.34)

Taking derivative of \(\overline{{\mathbb {L}}}_{I_h}\) yields

$$\begin{aligned} \frac{\partial \overline{{\mathbb {L}}}_{I_h}}{\partial t} = \left[ 1-\frac{I^*_h}{I_h}\right] \left[ \frac{S_m^*\beta }{d_h+\rho }D_h\Delta I_h +\frac{HS_m^*\beta }{d_h+\rho }\int \limits _0^\infty \beta _1(a)i_m(t,a,x)\mathrm{d}a - S_m^*\beta I_h\right] . \end{aligned}$$
(4.35)

For simplicity, we let

$$\begin{aligned} \widetilde{{\mathbb {L}}}(t,x) = \overline{{\mathbb {L}}}_{S_m}(t,x) + \overline{{\mathbb {L}}}_{i_m}(t,x) + \overline{{\mathbb {L}}}_{I_h}(t,x). \end{aligned}$$

With the help of (4.31), (4.34) and (4.35), we then have

$$\begin{aligned} \frac{\partial \overline{{\mathbb {L}}}}{\partial t}&= {\mathbb {W}}+i_m^*(0)-i_m(t,0,x)- i_m^*(0)\frac{S_m^*}{S_m} + i_m(t,0,x)\frac{S_m^*}{S_m}\\&\quad +G[i_m(t,0,x),i_m^*(0)]+\frac{HS_m^*\beta }{d_h + \rho } \int \limits _0^\infty \beta _1(a)\left[ i_m^*(a) - i_m(t,a,x) + i_m^*(a)\ln \frac{i_m(t,a,x)}{i_m^*(a)}\right] \mathrm{d}a\\&\quad +\left[ 1-\frac{I^*_h}{I_h}\right] \left[ \frac{HS_m^*\beta }{d_h+\rho } \int \limits _0^\infty \beta _1(a)i_m(t,a,x)\mathrm{d}a - S_m^*\beta I_h\right] , \end{aligned}$$

where

$$\begin{aligned} {\mathbb {W}}&= \left[ 1 - \frac{S_m^*}{S_m}\right] D_m\Delta S_m -\frac{d_m}{S_m}(S_m - S_m^*)^2\\&\quad + \int \limits _0^\infty \Psi _1(a)\left[ 1-\frac{i_m^*(a)}{i_m(t,a,x)}\right] D_m\Delta i_m(t,a,x)\mathrm{d}a+ \left[ 1 - \frac{I^*_h}{I_h}\right] \frac{S_m^*\beta }{d_h+\rho }D_h\Delta I_h. \end{aligned}$$

Collecting terms with \(i_m^*(0)\), \(i_m(t,0,x)\) and \(\frac{HS_m^*\beta }{d_h + \rho }\), respectively,

$$\begin{aligned} \frac{\partial \overline{{\mathbb {L}}}}{\partial t}&= {\mathbb {W}}+i_m^*(0)\left[ 1- \frac{S_m^*}{S_m}\right] -i_m(t,0,x)\left[ 1- \frac{S_m^*}{S_m}\right] \\&\quad +G[i_m(t,0,x),i_m^*(0)]+\frac{HS_m^*\beta }{d_h + \rho } \int \limits _0^\infty \beta _1(a)\left[ i_m^*(a) - i_m(t,a,x) + i_m^*(a)\ln \frac{i_m(t,a,x)}{i_m^*(a)}\right] \mathrm{d}a\\&\quad +\frac{HS_m^*\beta }{d_h+\rho }\int \limits _0^\infty \left[ 1-\frac{I^*_h}{I_h}\right] \beta _1(a)i_m(t,a,x)\mathrm{d}a - i_m(t,0,x)\frac{S_m^*}{S_m}+i_m^*(0).\\ \end{aligned}$$

Here, we have used the fact that \(\left[ 1-\frac{I^*_h}{I_h}\right] S_m^*\beta I_h=i_m(t,0,x)\frac{S_m^*}{S_m}-i_m^*(0)\). Canceling the zero terms, together with \(i^*_m(0)=\frac{HS_m^*\beta }{d_h + \rho } \int \limits _0^\infty \beta _1(a)i^*_m(a)\mathrm{d}a\), yields

$$\begin{aligned} \frac{\partial \overline{{\mathbb {L}}}}{\partial t}&= {\mathbb {W}}+i_m^*(0)\left[ 1-\frac{S_m^*}{S_m}-\ln \frac{i_m(t,0,x)}{i_m^*(0)}\right] \\&\quad + \frac{HS_m^*\beta }{d_h + \rho } \int \limits _0^\infty \beta _1(a)i_m^*(a)\left[ 1+ \ln \frac{i_m(t,a,x)}{i_m^*(a)} - \frac{I^*_hi_m(t,a,x)}{I_hi_m^*(a)}\right] \mathrm{d}a\\&= {\mathbb {W}}+ \frac{HS_m^*\beta }{d_h + \rho } \int \limits _0^\infty \beta _1(a)i_m^*(a)\left[ 2 + \ln \frac{i_m(t,a,x)}{i_m^*(a)} - \frac{I^*_hi_m(t,a,x)}{I_hi_m^*(a)} - \frac{S_m^*}{S_m} - \ln \frac{i_m(t,0,x)}{i_m^*(0)}\right] \mathrm{d}a. \end{aligned}$$

By a zero trick that \(1-\frac{S_mI_hi_m^*(0)}{S_m^*I^*_h i_m(t,0,x)}=0\), we have

$$\begin{aligned} \frac{\partial \overline{{\mathbb {L}}}}{\partial t}&= {\mathbb {W}}- \frac{HS_m^*\beta }{d_h + \rho }\int \limits _0^\infty \beta _1(a)i_m^*(a)\left[ g\left( \frac{S_m^*}{S_m}\right) + g\left( \frac{I^*_hi_m(t,a,x)}{I_hi_m^*(a)}\right) + g\left( \frac{S_mI_hi_m^*(0)}{S_m^*I^*_h i_m(t,0,x)}\right) \right] \mathrm{d}a, \end{aligned}$$

where \(g(\alpha )=\alpha -1-\ln \alpha , \alpha \in {\mathbb {R}}_+\) possesses the properties that \(g(\alpha )\ge 0\) when \(\alpha \ge 1\) and \(g(1)=0\). Hence,

$$\begin{aligned} \frac{\partial \overline{{\mathbb {L}}}}{\partial t}&= - D_mS_m^*\int \limits _\Omega \frac{\left| \nabla S_m\right| ^2}{S^2_m}\mathrm{d}x - \int \limits _0^\infty D_m\Psi _1(a)i_m^*(a)\int \limits _\Omega \frac{\left| \nabla i_m(t,a,x)\right| ^2}{i^2_m(t,a,x)}\mathrm{d}x\mathrm{d}a\\&\quad - \frac{D_hS_m^*\beta }{d_h+\rho }\int \limits _\Omega \frac{\left| \nabla I_h\right| ^2}{I^2_h}\mathrm{d}x - \int \limits _\Omega \frac{d_m}{S_m}(S_m - S_m^*)^2\mathrm{d}x\\&\quad -\int \limits _\Omega \frac{HS_m^*\beta }{d_h + \rho }\int \limits _0^\infty \beta _1(a)i_m^*(a)\left[ g\left( \frac{S_m^*}{S_m}\right) + g\left( \frac{I^*_hi_m(t,a,x)}{I_hi_m^*(a)}\right) + g\left( \frac{S_mI_hi_m^*(0)}{S_m^*I^*_h i_m(t,0,x)}\right) \right] \mathrm{d}a\mathrm{d}x\\&\le 0. \end{aligned}$$

Hence, with the help of the property of g, \(\frac{\partial \overline{{\mathbb {L}}}(t,x)}{\partial t} = 0\) if and only if \(S_m = S_m^*\), \(i_m(t,a,x) = i_m^*(a)\) and \(I_h= I^*_h\). According to the invariance principle (see, e.g., [31, Theorem 4.2]), the global attractivity of \(E^*\) directly follows. This together with the local stability of \(E^*\), as stated in Theorem 4.2, implies that \(E^*\) is GAS for all \(a\ge 0, x\in \Omega \). This proves Theorem. \(\square \)

5 Conclusion and discussion

This paper provides the complete analysis on the threshold dynamics of an age-space structured malaria epidemic model. Unlike in [21] where the spatial movement of mosquitoes in EIP will result in non-local infection, here we intend to incorporate an infection age to be a continuous variable and the fixed incubation period is ignored. We conducted a complete analysis of model (1.4) by adopting the density of susceptible population to be H(x) and the mass action incidence, which can be regarded as a continuous work of [21]. In Lemma 2.2, we confirmed that (1.4) admits a unique local solution through investigating the fixed point problem, which is defined in (2.6). It is also proved that the local solution of (1.4) is positive for initial conditions (see Lemma 2.3). The positivity of \(i_m(t,a,x)\) is implied by the positivity of \({\mathcal {B}}(t,x)\), which is proved by the methods of Picard sequences and iteration. The proof of which is indeed not trivial, as \({\mathcal {B}}(t,x)\) involves the product of \(S_m\) and \(I_h\), as demonstrated in (2.6). In addition, we have extended the interval of existence of the local solution to \([0,+\infty )\). To achieve this, we verified that in a finite time interval, the solution does not blow up (see Lemma 2.4). Consequently, the solution semiflow generated by the solution of (1.4) possesses a global attractor in \(\mathbb {{\tilde{X}}}_+\).

By introducing the renewal equation and Laplace transformation, we identified that the next-generation operator \({\mathcal {L}}\) defined in (3.6) is strictly positive and compact (see Lemma 3.1), which allow us to define \(\mathfrak {R}_0\). It is important to mention that it is difficult to obtain spectral radius of \({\mathcal {L}}\), if not impossible, so that we cannot get further information on dynamical properties of (1.4). To proceed further, we considered the special case where parameters are all independent of x. In such a setting, the Krein–Rutman theorem ensures that \([\mathfrak {R}_0]\) can be explicitly obtained if the positive eigenvector (corresponding to \([\mathfrak {R}_0]\)) is a constant. It is also identified that if \([\mathfrak {R}_0]<1\), then space-independent disease-free equilibrium \({\tilde{E}}_0 = (S_m^0,0,0)\), where \(S_m^0=\frac{\mu }{d_m}\), is GAS (see Theorem 4.3); if \([\mathfrak {R}_0]>1\), then the model is uniformly persistent (see Theorem 4.6) and the space-independent endemic equilibrium \(E^*=(S^*_m, i^*_m(0)\Pi (a), I_h^*)\) is GAS (see Theorem 4.7). The local stability of \(E_0\) and \(E^*\) is achieved by studying the distribution of characteristic roots of characteristic equation (4.9) (see Theorem 4.2). The strong persistence result is implied by the weak persistence (see Lemmas 4.4 and 4.5), which is achieved by making the Laplace transformation and the way of contraction. Both \({\tilde{E}}_0\) and \(E^*\) are proved to be GAS by Lyapunov functions in different cases (see Theorem 4.3 and 4.7).

If the average incubation period is fixed as \(\tau \), and

$$\begin{aligned} \beta _1(a) = \left\{ \begin{array}{ll} 0, &{}\quad 0\le a <\tau ; \\ \beta ^*, &{}\quad A\ge \tau , \end{array} \right. \end{aligned}$$

as standard arguments in [13, 21], our model (1.4) can be reformulated into, for \(x\in \Omega ,\ t>0\),

$$\begin{aligned} \left\{ \begin{array}{ll} \displaystyle \frac{\partial S_m}{\partial t} - D_m\Delta S_m = \mu (x)-\beta _2(x)S_mI_h-d_m S_m, \\ \displaystyle \frac{\partial I_m}{\partial t} -D_m\Delta I_m = e^{-d_m\tau }\int \limits _\Omega \Gamma _2(D_m\tau ,x,y)\beta _2(y)S_m(t-\tau ,y)I_h(t-\tau ,y)\mathrm{d}y- d_mI_m,\\ \displaystyle \frac{\partial I_h}{\partial t} - D_h\Delta I_h = H(x)\beta ^*I - (d_h+\rho )I_h, \end{array}\right. \end{aligned}$$
(5.1)

with

$$\begin{aligned} \frac{\partial S_m}{\partial n} = \frac{\partial I_m}{\partial n}= \frac{\partial I_h}{\partial n} = 0,\ x\in \partial \Omega ,\ t>0, \end{aligned}$$

where \(I_m = \int \limits _\tau ^\infty i_m(t,a,x)\mathrm{d}a\). As studied in [21, Section 4], with additional conditions, the global attractivity of endemic equilibrium in homogeneous case can be proved by using a fluctuation method.

If we further take seasonality into account, by using the standard arguments as those in [4, 36], our model (1.4) can be reformulated by following system, for \(x\in \Omega ,\ t>0\),

$$\begin{aligned} \left\{ \begin{array}{ll} \displaystyle \frac{\partial S_m}{\partial t} - D_m\Delta S_m = \mu (t,x)-\beta _2(t,x)S_mI_h-d_m(t,x) S_m, \\ \displaystyle \frac{\partial I_m}{\partial t} - D_m\Delta I_m = e^{-d_m\tau }\int \limits _\Omega \Gamma _2(D_m\tau ,x,y)\beta _2(t,y)S_m(t-\tau ,y)I_h(t-\tau ,y)\mathrm{d}y- d_m(t,x)I_m,\\ \displaystyle \frac{\partial I_h}{\partial t} - D_h\Delta I_h = H(x)\beta ^*I - (d_h+\rho )I_h, \end{array}\right. \end{aligned}$$
(5.2)

with

$$\begin{aligned} \frac{\partial S_m}{\partial n} = \frac{\partial I_m}{\partial n}= \frac{\partial I_h}{\partial n} = 0,\ x\in \partial \Omega ,\ t>0. \end{aligned}$$

Here, \(\mu (t,x)\), \(\beta _2(t,x)\) and \(d_m(t,x)\) are the recruitment rate, biting rate and mortality rate involving the seasonality, respectively. The dynamics of (5.1) and (5.2) will be left for future investigation.