Loading [MathJax]/jax/output/SVG/jax.js
Research article Special Issues

Improved uniform persistence for partially diffusive models of infectious diseases: cases of avian influenza and Ebola virus disease


  • Past works on partially diffusive models of diseases typically rely on a strong assumption regarding the initial data of their infection-related compartments in order to demonstrate uniform persistence in the case that the basic reproduction number R0 is above 1. Such a model for avian influenza was proposed, and its uniform persistence was proven for the case R0>1 when all of the infected bird population, recovered bird population and virus concentration in water do not initially vanish. Similarly, a work regarding a model of the Ebola virus disease required that the infected human population does not initially vanish to show an analogous result. We introduce a modification on the standard method of proving uniform persistence, extending both of these results by weakening their respective assumptions to requiring that only one (rather than all) infection-related compartment is initially non-vanishing. That is, we show that, given R0>1, if either the infected bird population or the viral concentration are initially nonzero anywhere in the case of avian influenza, or if any of the infected human population, viral concentration or population of deceased individuals who are under care are initially nonzero anywhere in the case of the Ebola virus disease, then their respective models predict uniform persistence. The difficulty which we overcome here is the lack of diffusion, and hence the inability to apply the minimum principle, in the equations of the avian influenza virus concentration in water and of the population of the individuals deceased due to the Ebola virus disease who are still in the process of caring.

    Citation: Ryan Covington, Samuel Patton, Elliott Walker, Kazuo Yamazaki. Improved uniform persistence for partially diffusive models of infectious diseases: cases of avian influenza and Ebola virus disease[J]. Mathematical Biosciences and Engineering, 2023, 20(11): 19686-19709. doi: 10.3934/mbe.2023872

    Related Papers:

    [1] Ning Bai, Juan Zhang , Li Li, Zhen Jin . Evaluating the effect of virus mutation on the transmission of avian influenza H7N9 virus in China based on dynamical model. Mathematical Biosciences and Engineering, 2019, 16(5): 3393-3410. doi: 10.3934/mbe.2019170
    [2] Kazuo Yamazaki, Xueying Wang . Global stability and uniform persistence of the reaction-convection-diffusion cholera epidemic model. Mathematical Biosciences and Engineering, 2017, 14(2): 559-579. doi: 10.3934/mbe.2017033
    [3] Danfeng Pang, Yanni Xiao . The SIS model with diffusion of virus in the environment. Mathematical Biosciences and Engineering, 2019, 16(4): 2852-2874. doi: 10.3934/mbe.2019141
    [4] Jinzhe Suo, Bo Li . Analysis on a diffusive SIS epidemic system with linear source and frequency-dependent incidence function in a heterogeneous environment. Mathematical Biosciences and Engineering, 2020, 17(1): 418-441. doi: 10.3934/mbe.2020023
    [5] Jing-An Cui, Fangyuan Chen . Effects of isolation and slaughter strategies in different species on emerging zoonoses. Mathematical Biosciences and Engineering, 2017, 14(5&6): 1119-1140. doi: 10.3934/mbe.2017058
    [6] Liqiong Pu, Zhigui Lin . A diffusive SIS epidemic model in a heterogeneous and periodically evolvingenvironment. Mathematical Biosciences and Engineering, 2019, 16(4): 3094-3110. doi: 10.3934/mbe.2019153
    [7] C. Connell McCluskey . Global stability for an SEI epidemiological model with continuous age-structure in the exposed and infectious classes. Mathematical Biosciences and Engineering, 2012, 9(4): 819-841. doi: 10.3934/mbe.2012.9.819
    [8] Cuicui Jiang, Kaifa Wang, Lijuan Song . Global dynamics of a delay virus model with recruitment and saturation effects of immune responses. Mathematical Biosciences and Engineering, 2017, 14(5&6): 1233-1246. doi: 10.3934/mbe.2017063
    [9] Ming-Zhen Xin, Bin-Guo Wang, Yashi Wang . Stationary distribution and extinction of a stochastic influenza virus model with disease resistance. Mathematical Biosciences and Engineering, 2022, 19(9): 9125-9146. doi: 10.3934/mbe.2022424
    [10] Karen R. Ríos-Soto, Baojun Song, Carlos Castillo-Chavez . Epidemic spread of influenza viruses: The impact of transient populations on disease dynamics. Mathematical Biosciences and Engineering, 2011, 8(1): 199-222. doi: 10.3934/mbe.2011.8.199
  • Past works on partially diffusive models of diseases typically rely on a strong assumption regarding the initial data of their infection-related compartments in order to demonstrate uniform persistence in the case that the basic reproduction number R0 is above 1. Such a model for avian influenza was proposed, and its uniform persistence was proven for the case R0>1 when all of the infected bird population, recovered bird population and virus concentration in water do not initially vanish. Similarly, a work regarding a model of the Ebola virus disease required that the infected human population does not initially vanish to show an analogous result. We introduce a modification on the standard method of proving uniform persistence, extending both of these results by weakening their respective assumptions to requiring that only one (rather than all) infection-related compartment is initially non-vanishing. That is, we show that, given R0>1, if either the infected bird population or the viral concentration are initially nonzero anywhere in the case of avian influenza, or if any of the infected human population, viral concentration or population of deceased individuals who are under care are initially nonzero anywhere in the case of the Ebola virus disease, then their respective models predict uniform persistence. The difficulty which we overcome here is the lack of diffusion, and hence the inability to apply the minimum principle, in the equations of the avian influenza virus concentration in water and of the population of the individuals deceased due to the Ebola virus disease who are still in the process of caring.



    In the past few decades, there has been a growing interest in modeling real-world phenomena via systems of partial differential equations (PDEs) solved by functions of not only time but also space, rather than ordinary differential equations (ODEs). In the study of mathematical models of infectious diseases, the basic reproduction number R0 measures the expected number of secondary infections caused by one infectious individual during its infectious period in an otherwise susceptible population. Proposing and analyzing PDE models of infectious diseases have become very popular, led by important works such as [1,2,3].

    For clarity of discussions let us state the following definition precisely.

    Definition 1.1. Consider a system of m-many PDEs solved by (u1,,um)(x,t):

    tu1(x,t)=D1Δu1(x,t)+F1(x,t,u1,,um), (1.1a)
    tu2(x,t)=D2Δu2(x,t)+F2(x,t,u1,,um), (1.1b)
    (1.1c)
    tum(x,t)=DmΔum(x,t)+Fm(x,t,u1,,um), (1.1d)

    where we denoted t:=t. We say that this system (1.1) is

    1) "fully diffusive" if Dj>0 for all j{1,,m},

    2) "partially diffusive" if there exists some k{1,,m} such that Dk=0.

    The spatially diffusive terms such as DjΔuj within such PDE models can help capture the movement of human hosts and other entities such as viruses or bacteria. The diffusivity coefficients Dj's describe the distinct level of mobility for different population groups. It has been documented for some time (see [4,5]) that common diffusivity coefficients (i.e., D1=Dm) are desperately needed so that one can add equations together, cancel out the nonlinear terms and thereby verify the global existence of a unique solution, only after which asymptotic behavior of solutions such as global stability may be discussed. Thus, many models with fully diffusive PDEs have assumed common diffusivity coefficients; we refer to as examples

    1) [6, Eq (1.14)] for a dengue fever model,

    2) [7, Eqs (2),(4) and (5)] for a malaria model,

    3) [8, Eq (1.1)] for a Lyme disease model,

    4) [9, hypothesis of Theorem 2.2] and [10, hypothesis of Theorem 2.1] for a cholera model,

    5) [11, hypothesis of Theorem 2.1] for a Zika virus disease model.

    As examples of models that are not necessarily of infectious diseases but which assumed uniform diffusivity coefficients, we refer to

    1) [12, Eq (1.9)] for an internal storage model,

    2) [13, Eq (1.5)] for a periodically-pulsed bio-reactor model,

    3) [14, Eq (2.1)] for a hydraulic storage zone model,

    4) [15, Eq (3.1)] for a zooplankton and harmful algae model.

    Proving uniform persistence of disease when R0 is above 1 under additional conditions on initial data is an important topic that has received much attention (see [16]). The purpose of this work is to address the uniform persistence results in more extreme cases of partially diffusive models so that not all equations consist of diffusion. Primary examples of our concern are an avian influenza (AI) model from [17] in which only the equation of the concentration of the AI virus in water lacked diffusion (see (1.2)), and an Ebola virus disease (EVD) model from [18] in which only the equation of the population of individuals deceased due to the EVD lacked diffusion (see (1.4)). Although we will not study this issue directly in this work, one may also consider the evidence that the coronavirus of 2019 (COVID-19) survives in the environment for several days (see [19]), and thus including a non-diffusive equation of concentration of COVID-19 in the environmental reservoir may seem reasonable (see [20]).

    A lack of diffusion in general implies an inability to apply the minimum principle on its own across all equations. This seems to be the main reason why Vaidya et al. [17] obtained uniform persistence of AI when its R0 is above one under a strong assumption that all of the infected bird population, the recovered bird population and the AI virus concentration are initially non-vanishing, rather than when any of the infection-related compartments are initially non-vanishing (see Theorem 1.1). Similarly, Yamazaki in [18] obtained uniform persistence of EVD when its R0 is above one, and the infected human population is initially non-vanishing, but not when only the population of the human individuals who died due to EVD but are still in the process of caring are initially non-vanishing (see Theorem 1.2).

    It would be biologically meaningful if the condition on the initial data may be improved to an assumption that any, rather than all, of the infection-related compartments are initially non-vanishing. For example, one can try to prove the uniform persistence of AI when only the concentration of the AI virus in water is initially non-vanishing or the uniform persistence of the EVD when only the population of the deceased individuals due to EVD still in the process of caring is initially non-vanishing. As we will explain in detail, in the case of AI, in addition to transmission of AI from a bird to another bird, uninfected birds may drink water and ingest the virus through excretion of the viruses by infected birds (see [21,22,23]). In the case of EVD, we know that burial ceremonies that involve direct contact with the body of the deceased can also contribute to the transmission (see [24]).

    Interestingly, such uniform persistence results starting from initial data that only guarantees that at least one of the infection-related compartments is non-vanishing exist for systems of fully diffusive PDEs modeling various infectious diseases. For example, Yamazaki and Wang [10] proved uniform persistence of the cholera virus in the case R0>1 when either infectious population or the concentration of cholera bacteria in water such as rivers is initially non-vanishing; that is, only one of the infection-related compartments had to be initially non-vanishing for such uniform persistence result to hold. A key difference was that in this cholera model, the equation of the cholera bacteria had diffusion. Similarly, Yamazaki in [11] proved uniform persistence of the Zika virus disease when its R0 is above 1 and any of the exposed human female population, exposed human male population or exposed mosquito population, and hence any of the infection-related compartments, is initially non-vanishing. We emphasize again that all the equations of these infection-related compartments had diffusion. Proofs in [10,11] crucially relied on the minimum principle (see [25, Theorem 7.1.12]), which was available only because all the equations governing the infection-related compartments had diffusion.

    To the best of our knowledge, for this reason, uniform persistence results of partially diffusive models when the corresponding R0 is above one under a weak assumption that at least one of its infection-related compartments is initially non-vanishing has never been proven. We achieve this goal in the cases of the AI and the EVD models (see Theorems 2.1 and 2.2). Our approach is generalizable to other systems of PDEs too (as we do in [26]). The mathematical novelty consists of discovering the hidden ability of solutions to non-diffusive equations to induce strict positivity everywhere for not only their own compartment, but for all other infection-related solution components via an indirect approach (see Propositions 3.1 and 4.1).

    Let us describe the main equations, namely those of AI in (1.2) and EVD in (1.4). In both cases, we consider the spatial domain Ω which is a bounded, open and connected subset of Rn for nN with a smooth boundary Ω.

    As a first main model, we introduce the AI model from [17]. AI viruses are of major concern due to their potential socioeconomic impact and risks to wildlife conservation. Since October 2020, highly pathogenic AI viruses belonging to geese have been responsible for more than 70 million poultry deaths and 100 discrete infections in many wild mesocarnivore species (see [27]). It is known that aquatic birds are the primary reservoir of AI viruses and that AI viruses can be transmitted from bird to bird, but they may also be transmitted through excretion of the AI virus by infected birds followed by ingestion of the virus in the drinking water of uninfected birds (see [21,22,23]). For this reason, the authors in [17] chose to include an equation V(x,t) for the AI virus concentration in water. For further details of the AI virus and the derivation of the model, we refer to [17, Sections 1 and 2]. In addition to V(x,t), let us denote by S(x,t),I(x,t) and R(x,t) the population density of the bird population that are susceptible, infected and recovered, respectively. We refer to Table 1 for further notations.

    Table 1.  Definition of parameters in model (1.2a)–(1.2d).
    Parameter Definition
    α Shedding rate of infected bird hosts in their feces
    β1 Direct transmission rate from infectious bird
    β2 Indirect transmission rate concentration of AI virus
    d Natural death rate of birds
    D Diffusivity coefficients of S,I,R
    η Rate of bird hosts immunity loss
    γ Recovery rate of infected bird hosts
    λ Recruitment rate of susceptible bird hosts

     | Show Table
    DownLoad: CSV

    Considering the significant mobility of birds in contrast to that of the concentration of the AI virus in water, the following model was proposed in [17, Eqs (1)–(4)]:

    tS=DΔS+λ(β1I+β2V)SdS+ηR, (1.2a)
    tI=DΔI+(β1I+β2V)S(γ+d)I, (1.2b)
    tR=DΔR+γI(η+d)R, (1.2c)
    tV=αIc(x)V, (1.2d)

    subjected to an initial condition

    (S,I,R,V)(x,0)=(ϕ1,ϕ2,ϕ3,ϕ4)(x)

    and standard Neumann boundary condition. We observe that (1.2) has a disease-free equilibrium (DFE) of

    (S,I,R,V)=(mA,0,0,0) where mA:=λd. (1.3)

    As a second main model, we introduce the EVD model from [18] (its ODE version was initially studied in [28]). The EVD is transmitted from wild animals to susceptible humans, and infectious humans may spread to other humans as well. As recently as September 2022, the Sudan Ebola virus was confirmed by the Uganda Virus Research Institute in which a total of 164 cases with 77 deaths were discovered (see [29]). A special feature of the EVD model of concern from [18,28] is the inclusion of an equation of D(x,t), the population of the individuals deceased due to the EVD. This is due to the warm-hearted West African tradition and customs of caring for the deceased individuals by kissing them, washing them and dressing them up, even those who passed away due to the EVD despite the risk of infection. Infections during funeral and subsequent death have been documented in various articles (see [30]). For further details of EVD and the derivation of the model, we refer to [28, Sections 1 and 2] and [18, Sections 1 and 2]. In addition to D(x,t), let us denote S(x,t),I(x,t),R(x,t) and P(x,t) as the human population of susceptible, infected, recovered individuals and the Ebola virus pathogens in the environment, respectively. We refer to Table 2 for further notations.

    Table 2.  Definition of parameters in model (1.4).
    Parameter Definition
    α Shedding rate of deceased human individuals
    1b Mean caring duration of deceased human individuals
    β1 Effective contact rate of infectious human individuals
    β2 Effective contact rate of deceased human individuals
    Di,i=1,2,3,4 Diffusion coefficients of S,I,R,P respectively
    δ Disease-induced death rate of human individuals
    η Decay rate of Ebola virus in the environment
    γ Recovery rate of infectious human individuals
    λ Effective contact rate of Ebola virus
    μ Natural death rate of human individuals
    π Recruitment rate of susceptible human individuals
    ξ Shedding rate of infectious human individuals

     | Show Table
    DownLoad: CSV

    Considering the lack of mobility of the deceased individuals due to the EVD, the following model was proposed in [18, Eqs (1) and (2)]:

    tS=D1ΔS+π(β1I+β2D+λP)SμS, (1.4a)
    tI=D2ΔI+(β1I+β2D+λP)S(μ+δ+γ)I, (1.4b)
    tR=D3ΔR+γIμR, (1.4c)
    tP=D4ΔP+ξI+αDηP, (1.4d)
    tD=δIbD, (1.4e)

    subjected to the initial conditions of

    (S,I,R,P,D)(x,0)=(ϕ1,ϕ2,ϕ3,ϕ4,ϕ5)(x)

    and standard Neumann boundary conditions. We see that (1.4) has a DFE of

    (S,I,R,P,D)=(mE,0,0,0,0) where mE:=πμ (1.5)

    Remark 1.1. We observe that the diffusivity coefficients in (1.2) were same while those in (1.4) had the freedom to differ. We refer to [4] for technical details.

    Before we proceed further, let us fix the standard notation to represent the solution as u. Because no confusion occurs, by an abuse of notation, we write

    u:=(u1,u2,u3,u4):=(S,I,R,V) while u:=(u1,u2,u3,u4,u5):=(S,I,R,P,D) (1.6)

    as solutions to (1.2) and (1.4), respectively. We consider the space of continuous functions with supremum norm; i.e., fC(D):=supxD|f(x)| and define

    X:=C(¯Ω,Rk)=ki=1Xi where Xi:=C(¯Ω,R) and ˉΩ:=ΩΩ, (1.7)

    which is the space of Rk-valued functions that are continuous in x¯Ω and equipped with the supremum norm of uC(¯Ω):=ki=1uiC(¯Ω), where k=4 for the AI model while k=5 for the EVD model. Furthermore, we define

    X+:=C(¯Ω,Rk+)=ki=1X+i where X+i:={fC(¯Ω,R):f0}. (1.8)

    In this section, we describe previous results, and challenges brought by the absence of diffusion in some equations. First, let us recall two important definitions.

    Definition 1.2. Let Y be any metric space, Y0Y an open set, Y0:=YY0 and Ψt:YY be any semiflow on Y.

    1) ([31, p. 6172]; see also [16, Definition 1.3.1]) A continuous function q:Y[0,) is called a generalized distance function for Ψt if it satisfies q(Ψt(y))>0 for all t>0 whenever

    (a) q(y)=0 and yY0,

    (b) or q(y)>0,

    2) ([16, Definition 1.3.3]) Let q be a generalized distance function for Ψt. Then, Ψt is said to be uniformly persistent with respect to (w.r.t.) (Y0,Y0,q) if there exists η>0 such that

    lim infnq(Ψnt(ϕ))ηϕY0. (1.9)

    Remark 1.2. It is clear from (1.9) that the larger the Y0 chosen, the stronger uniform persistence result becomes.

    Let us list some of the main results obtained in [17,18].

    List 1.1. (a) For any ϕX+, both (1.2) and (1.4) admit a unique non-negative solution that remains in X+, and thus a semiflow Φt:X+X+ defined by Φt(ϕ):=u(t) for all t>0 such that u(x,0)=ϕ(x) (see [17, Theorem 3.3] and [18, Theorem 2.1]).

    (b) The basic reproduction number R0 was rigorously derived for both systems (1.2) and (1.4) from their corresponding infection-related compartments, namely I and V in (1.2) and I,P, and D in (1.4) (see [17, Eq (27)] and [18, Eq (34)]).

    (c) When R0<1, the corresponding DFE are globally attractive in X+ for both (1.2) and (1.4), although this result for (1.4) requires D1=D2=:ˉD (see [17, Theorem 3.8 (1)] and [18, Theorem 2.2]).

    In more detail, to deduce the result from List 1.1 (b), [17] considered the eigenvalue problem associated with the infection-related compartments I and V of (1.2), namely

    θψ2=DΔψ2+β1H(x)ψ2+β2H(x)ψ4(γ+d)ψ2, (1.10a)
    θψ4=αψ2cψ4, (1.10b)

    for an arbitrary H:ˉΩ(0,) (see [17, Eq (22)]) and proved the following result.

    Lemma 1.1. ([17, Lemma 3.4 and Proposition 1]) Let R0 be the basic reproduction number of the AI system (1.2) and (mA,0,0,0) its DFE from (1.3). Then, for any H:ˉΩ(0,), the eigenvalue problem (1.10) has a principal eigenvalue θ(H) associated with a strictly positive eigenfunction. Moreover, R01 and θ(mA) have the same sign.

    Similarly, to deduce the result from List 1.1 (b), [18] considered the eigenvalue problem associated with the infection-related compartments I,P and D of the EVD system (1.4), namely

    θψ2=¯DΔψ2+(β1ψ2+β2ψ5+λψ4)H(x)(μ+δ+γ)ψ2, (1.11a)
    θψ4=¯D4Δψ4+ξψ2+αψ5ηψ4, (1.11b)
    θψ5=δψ2bψ5, (1.11c)

    for an arbitrary H:ˉΩ[0,) (see [18, Eq (24)]) and proved the following result.

    Lemma 1.2. ([18, Propositions 4.3 and 4.4]) Let R0 be the basic reproduction number of the EVD system (1.4) and (mE,0,0,0,0) its DFE from (1.5). Then, for any H:ˉΩ(0,), the eigenvalue problem (1.11) has a principal eigenvalue θ(H) associated with a strictly positive eigenfunction. Moreover, R01 and θ(mE) have the same sign.

    Now, let us state the main results from [17, Theorem 3.8 (ⅱ)] and [18, Theorem 2.3].

    Theorem 1.1. ([17, Theorem 3.8 and its proof]) Define

    W0:={ψX+:ψ20,ψ30,andψ40}, (1.12a)
    W0:=X+W0={ψX+:ψ20orψ30orψ40}, (1.12b)

    and

    p:X+[0,)byp(ψ):=min{minxˉΩψ2(x),minxˉΩψ3(x),minxˉΩψ4(x)}. (1.13)

    If R0 defined by [17, Eq (27)] satisfies R0>1, then the AI system (1.2) admits at least one positive steady state ˆu and there exists σ>0 such that for any ϕX+ where ϕi0 for all i{2,3,4},

    lim inftui(x,t,ϕ)σi{1,2,3,4},xˉΩ, (1.14)

    where ui(x,0,ϕ)=ϕi(x); i.e., uniform persistence w.r.t. (W0,W0,p) holds for W0 and W0 defined in (1.12).

    Theorem 1.2. ([18, Theorem 2.3 and its proof]) Define

    W0:={ψX+:ψ20}, (1.15a)
    W0:=X+W0={ψX+:ψ20}, (1.15b)

    and

    p:X+[0,)byp(ψ):=min{minxˉΩψ2(x),minxˉΩψ3(x),minxˉΩϕ4(x),minxˉΩϕ5(x)}. (1.16)

    If D1=D2=:ˉD and R0 defined by [18, Eq (34)] satisfies R0>1, then the EVD system (1.4) admits at least one positive steady state ˆu and there exists σ>0 such that for any ϕX+ where ϕ20,

    lim inftui(x,t,ϕ)σi{1,2,3,4,5},xˉΩ, (1.17)

    where ui(x,0,ϕ)=ϕi(x); i.e., uniform persistence w.r.t. (W0,W0,p) holds for W0 and W0 defined in (1.15).

    As we pointed out already, the infection-related compartments of the AI system (1.2) are I and V, while those of the EVD system (1.4) are I,P and D. Further, for the fully diffusive systems of PDEs such as those for cholera [10] and the Zika virus disease [11], uniform persistence of the disease when the basic reproduction number was above 1 was successfully proven as long as any one of the infection-related compartments is initially non-vanishing. Therefore, it is a natural question to ask if we can improve W0 and consequently W0 in (1.12) to

    W0:={ψX+:ψ20 or ψ40}, (1.18a)
    W0:=X+W0={ψX+:ψ20 and ψ40}, (1.18b)

    and improve W0 in (1.15) to

    W0:={ψX+:ψ20 or ψ40 or ψ50}, (1.19a)
    W0:=X+W0={ψX+:ψ20,ψ40, and ψ50}; (1.19b)

    these are clear improvements based on Remark 1.2. In order to understand the difficulty of this aim, let us briefly review the proofs of Theorems 1.1 and 1.2. It turns out that both Theorems 1.1 and 1.2 rely crucially on the following consequence of the minimum principle.

    Lemma 1.3. ([17, Lemma 3.7]) Define mA by (1.3). Suppose that u(x,t,ϕ) is a solution to (1.2) such that u(,0,ϕ)=ϕ()X+.

    (1) For any i{2,3}, if there exists some ti00 such that ui(,ti0,ϕ)0, then ui(x,t,ϕ)>0 for all xˉΩ and all t>ti0.

    (2) For any ϕX+, u1(x,t,ϕ)>0 for all xˉΩ and all t>0, and

    lim inftu1(,t,ϕ)λΥA, (1.20)

    where

    ˜c:=minxˉΩc(x)andΥA:=2mAβ1+4αmA˜cβ2+d. (1.21)

    (3) If there exists some t200 such that u2(,t20,ϕ)0, then u4(x,t,ϕ)>0 for all xˉΩ and all t>t20.

    In particular, Lemma 1.3 (3) follows from solving for u4 of (1.2) directly as

    u4(x,t)=ec(x)(tt20)u4(x,t20)+αec(x)ttt20u2(x,s)ec(x)sdsxΩ,tt20

    and relying on Lemma 1.3 (1).

    Lemma 1.4. ([18, Proposition 4.7]) Define mE by (1.5). Suppose that D1=D2=:ˉD and u(x,t,ϕ) is a solution to (1.4) such that u(,0,ϕ)=ϕ()X+.

    (1) For any i{2,3,4}, if there exists some ti00 such that ui(,ti0,ϕ)0, then ui(x,t,ϕ)>0 for all xˉΩ and all t>ti0.

    (2) For any ϕX+, u1(x,t,ϕ)>0 for all xˉΩ and all t>0, and

    lim inftu1(,t,ϕ)πΥE, (1.22)

    where

    ΥE:=β12mE+β2(4mEδb)+λ(ξ4mE+[α8mEδb]η)+μ. (1.23)

    (3) If there exists some t200 such that u2(,t20,ϕ)0, then u5(x,t,ϕ)>0 for all xˉΩ and all t>t20.

    Proof of Lemma 1.4. The only improvement that we included here is in part (2); in contrast, [18, Proposition 4.7 (2)] stated that there exists T>0 such that u1(x,t,ϕ)>0 for all xˉΩ and t>T rather than u1(x,t,ϕ)>0 for all xˉΩ and t>0. As we see, it is also stated in Lemma 1.3 (2) (from [17, Lemma 3.7 (ⅲ)]) but its proof was omitted. We give a quick proof here. Suppose that there exists (x,t)Ω×(0,) at which u1(x,t,ϕ)=0. By non-negativity of the solution, this is a local minimum, and thus by Lemma A.1 we have ΔS(x,t)0 and thus (1.4a) gives us tS(x,t)π>0, which leads to a contradiction because, for some ϵ(0,t) sufficiently small, S(x,tϵ)<0. Therefore, S(x,t)>0 for all xΩ and all t>0. It follows immediately from this that S(x,t)>0 for all xˉΩ and all t>0 as claimed (see [32, Theorem 7.4.1]).

    Remark 1.3. We make a key observation from both Lemmas 1.3 and 1.4 that an assumption of the existence of t400 such that u4(,t40,ϕ)0 in (1.2) or an assumption of the existence of t500 such that u5(,t50,ϕ)0 in (1.4) does not immediately give us any positivity on any solution component, not even itself. This is completely reasonable; the proofs of Lemma 1.3 (1) and Lemma 1.4 (1) consist of a standard application of the minimum principle (see [25, Theorem 7.1.12]) and it is false in general that a solution to a non-diffusive equation can attain strict positivity everywhere for all t>t0 based only on the assumption that it does not vanish at time t0.

    Let us point out specifically two occasions where the difficulty would arise if we replace the W0 in (1.12a) by that in (1.18a) for the AI model (1.2) and similarly W0 in (1.15a) by that in (1.19a) for the EVD model (1.4).

    First, in the case of the AI model (1.2), in an effort to prove that the DFE is a uniform weak repeller for W0, they claim on [17, p. 2839]

    let ˜ψ:=(˜ψ2,˜ψ4) be the strict positive eigenfunction corresponding to [θ(mAδ0)>0]. Since ui(x,t,ϕ0)>0,xˉΩ,t>0,i=2,4, there exists ϵ0>0 such that [(u2(x,t1,ϕ0),u4(x,t1,ϕ0))ϵ0eθ(mAδ0)t1˜ψ].

    Here, if ϕW0 with W0 from (1.18a), then it is possible that the only available assumption is that ϕ40. That alone and Lemma 1.3 cannot guarantee us that both u2(x,t,ϕ) and u4(x,t,ϕ) are strictly positive on ˉΩ for all t>0; in fact, it cannot even guarantee the strictly positivity of either one. However, the strict positivity on ˉΩ for all t>0 is needed to achieve the claim (u2(x,t1,ϕ0),u4(x,t1,ϕ0))ϵ0eθ(mAδ0)t1˜ψ because ˜ψ is strictly positive on ˉΩ due to Lemma 1.1.

    In the case of the EVD model (1.4), an analogous situation is slightly improved as follows (see [18, p. 24]).

    Now, by hypothesis, we have ϕ=(ϕ1,ϕ2,ϕ3,ϕ4,ϕ5)W0 so that ϕ2()0. By [Lemma 1.4 (3)] this implies that u2(x,t,ϕ)>0,u5(x,t,ϕ)>0 for all xˉΩ,t>0. Suppose that for any t0>0, we have u4(,t0,ϕ)0 on ˉΩ. But, then by [(1.4d)], we deduce tP>0 for all (x,t)Ω×{t0}, which is a contradiction to [the non-negativity of the solution.] Therefore, we must have u4(,t0,ϕ)0 for all t0>0. By [Lemma 1.4 (1)], this implies u4(x,t,ϕ)>0 for all xˉΩ,t>0 by arbitrariness of t0>0. Thus, we see that in particular (u2,u4,u5)(x,t1,ϕ)0 so that we may find ϵ>0 sufficiently small such that (I,P,D)(x,t1,ϕ)ϵ˜ψ(x).

    Although this argument had a nice outcome that only ϕ20 led to not only u2,u5>0 due to Lemma 1.4 (3) but also u4>0, an analogous proof in the case that W0 in (1.15a) is replaced by that in (1.19a) seems hopeless. In particular, if all we know is that ϕ50, then Lemma 1.4 alone is not sufficient to guarantee the positivity of all of u2,u4 and u5 on ˉΩ and all t>0; in fact, it does not guarantee the strict positivity of any one of them. Yet, strict positivity of all of them is needed to achieve the desired (I,P,D)(x,t1,ϕ)ϵ˜ψ(x) considering that ˜ψ is strictly positive due to Lemma 1.2.

    Next, the following difficulty to be described next is even more dire than the first. In the case of the AI model (1.2), the function p defined in (1.13) is claimed to be a generalized distance function for the semiflow Φt for (1.2) (recall Definition 1.2).

    By [Lemma 1.3], it follows that... p has the property that if p(ϕ)>0 or ϕW0 with p(ϕ)=0, then p(Φtϕ)>0, t>0

    from [17, p. 2839]. Suppose we replace W0 in (1.12a) by that in (1.18a). Then, in the case that ϕW0 with p(ϕ)=0, the only assumption we can make may be that ϕ40, from which Lemma 1.3 alone cannot guarantee that P from (1.13) satisfies

    p(Φtϕ)=min{minxˉΩu2(x,t),minxˉΩu3(x,t),minxˉΩu4(x,t)}>0t>0.

    In the case of the EVD model (1.4), an analogous situation is again slightly improved as follows (see [18, p. 24]).

    The hypothesis that ϕW0 implies ϕ2()0. By the identical argument in the proof of [the DFE being a weak repeller], it follows that u2(x,t,ϕ)>0 which implies u5(x,t,ϕ)>0 by [Lemma 1.4] which leads to that u4(x,t,ϕ)>0 for all x¯Ω,t>0 as well. Moreover, suppose that for any t0>0,u3(,t0,ϕ)0. Then, by [(1.4c)], we obtain tR|t=t0>0,R(,t0)0 for all xΩ and hence contradiction to [the non-negativity of the solution]. Therefore, for all t0>0,u3(,t0,ϕ)0. By [Lemma 1.4], this implies R(x,t)>0 for all x¯Ω,t>0. Hence, p(Φt(ϕ))>0 for all t>0.

    The extra argument presented here allowed the only assumption of ϕ20 to verify that p from (1.16) satisfies

    p(Φtϕ)=min{minxˉΩu2(x,t),minxˉΩu3(x,t),minxˉΩu4(x,t),minxˉΩu5(x,t)}>0t>0;

    thus, it is a generalized distance function. Nevertheless, extending this argument to W0 from (1.19a) seems difficult; in particular, starting from only an assumption of ϕ50, it is not clear if we can guarantee that all of u2,u3,u4 and u5 are strictly positive on ˉΩ for all t>0. Again, Lemma 1.4 does not immediately guarantee that even one of them is strictly positive, even u5 itself.

    In this section we present our main results: Theorems 2.1 and 2.2. In sharp contrast to Remark 1.3, starting from the non-vanishing assumption at any t00 of the solution component of the non-diffusive equation, strict positivity of all infection-related compartments for all xˉΩ and all t>t0 can be shown (see Propositions 3.1 and 4.1). This ability of a solution to a non-diffusive equation such that it does not vanish at time t0 to induce strict positivity on ˉΩ for all t>0 on not only itself but even other infection-related components is surprising and it can be derived by taking advantage of the structure of the infectious disease models. Therefore, our approach is general and may prove to be suitable for many other models, such as the COVID-19 model in [26]. Consequently, Theorems 2.1 and 2.2 improve Theorems 1.1 and 1.2 on systems (1.2) and (1.4) by replacing W0 with those of (1.18a) and (1.19a), respectively.

    Theorem 2.1. Define W0 and W0 by (1.18), and p by (1.13). If R0 defined by [17, Eq (27)] satisfies R0>1, then (1.2) admits at least one positive steady state ˆu and there exists σ>0 such that for any ϕX+ such that ϕ20 or ϕ40, (1.14) is satisfied so that uniform persistence w.r.t. (W0,W0,p) holds for W0 and W0 defined in (1.18).

    Theorem 2.2. Let D1=D2=:ˉD in (1.4). Define W0 and W0 by (1.19), and p by (1.16). If R0 defined by [18, Eq (34)] satisfies R0>1, then (1.4) admits at least one positive steady state ˆu and there exists σ>0 such that for any ϕX+ such that ϕ20 or ϕ40 or ϕ50, (1.17) is satisfied so that uniform persistence w.r.t. (W0,W0,p) holds for W0 and W0 defined in (1.19).

    In comparison of Theorems 2.1 and 2.2 with Theorems 1.1 and 1.2, there is no direct improvement concerning the existence of a positive steady state ˆu. Nevertheless, its proof relies on the latter's uniform persistence results and therefore definitions of W0; thus, we included them for completeness.

    Before we proceed, we emphasize the novelties of this work and make comments.

    (1) Obviously, the W0 in (1.12a) is strictly embedded in the W0 from (1.18a); the same statement can be made concerning (1.15a) and (1.19a). Thus, both Theorems 2.1 and 2.2 are direct improvements of Theorems 1.1 and 1.2 according to Remark 1.2.

    (2) Theorems 2.1 and 2.2 are not only mathematically but also biologically meaningful. In words, Theorem 2.1 states that, in the case R0>1, if initially there are any concentration of AI virus in water, then the disease persists. Similarly, Theorem 2.2 states that, in the case R0>1, if initially there is any deceased individuals due to EVD still in the process of caring (recall 1b from Table 2), then the disease persists.

    The risks of infections from concentration of AI viruses in water to susceptible birds, as well as from the deceased individuals due to EVD to susceptible individuals during the burial ceremonies have been documented in the past, e.g., [21,22,23] in the case of the AI viruses and [24,28,30] in the case of the EVD. Theorems 2.1 and 2.2 are results with rigorous proofs that support the severity of such risks by predicting uniform persistence of the diseases even in the initial absence of other infection-related compartments.

    (3) Besides the actual results stated in Theorems 2.1 and 2.2, their proofs offer interesting features to researchers on PDEs in general. As we pointed out in Remark 1.3, it is false that a solution to an equation without diffusion has a property similar to Lemma 1.3 (1) or Lemma 1.4 (1). However, in the case of the AI model, u4 non-vanishing at t00 can, together with the fact that u1(x,t)>0 for all xˉΩ and t>0 due to Lemma 1.3 (2), show that u2(x,t)>0 for all xˉΩ and t>0, and this in turn actually shows that u4(x,t)>0 by Lemma 1.3 (3) (see Proposition 3.1 (1)f). Similarly in the case of EVD model, u5 non-vanishing at t00 can, together with the fact that u1(x,t)>0 for all xˉΩ and t>0 due to Lemma 1.4 (2), show that u2(x,t)>0 for all xˉΩ and t>0, and this, in turn shows that u5(x,t)>0 by Lemma 1.4 (3) (see Proposition 4.1 (1)).

    In subsequent sections, we prove Theorems 2.1 and 2.2 in order; due to their similarities, we will elaborate in detail for the proof of Theorem 2.1, while we will give main ideas only in the proof of {Theorem 2.2}. Throughout the rest of the work, for both (W0,W0) defined by (1.18) or (1.19) and Φt, the solution semiflow for (1.2) or (1.4) that is guaranteed to exist by List 1.1, we define

    M:={ϕW0:Φt(ϕ)W0t0} (2.1)

    and ω(ϕ) to be the omega limit set of the orbit O+(ϕ):={Φt(ϕ):t0}. Moreover, we denote the Kuratowski measure of non-compactness by κ (see Definition A.1 for precise definition). After proofs, we follow up with the conclusion; moreover, for readers' convenience, we leave some preliminaries and computations in the Appendix.

    Besides the results from List 1.1, Lemmas 1.1 and 1.3, let us recall some results from [17] that will be needed to prove Theorem 2.1.

    Lemma 3.1. ([17, Lemmas 3.5 and 3.6]) The semiflow Φt:X+X+ defined by Φt(ϕ)=u(t) guaranteed by List 1.1 (a) is a κ-contraction on X+; i.e., there exists a constant ˜c>0 such that κ(ΦtB)e˜ctκ(B) for all bounded sets BX+ such that κ(B)>0 so that limtκ(ΦtB)=0. Moreover, Φt is κ-condensing and consequently asymptotically smooth. Finally, Φt admits a global connected attractor A.

    The following result is an improvement of Lemma 1.3. In particular, it shows that a solution to the non-diffusive equation, namely u4 that solves (1.2d), starting from some t400 at which it does not vanish, does indeed become and remain strictly positive for all t>t0. Not only that, it has the ability to spread such strict positivity to the other infection-related compartment, namely u2.

    Proposition 3.1. Suppose that u(x,t,ϕ) is the solution to (1.2) such that u(,0,ϕ)=ϕ()X+.

    (1) If there exists ti00 such that ui(,ti0,ϕ)0 for i=2 or i=4, then uk(x,t,ϕ)>0 for all (x,t)ˉΩ×(ti0,) and all k{2,4}.

    (2) Moreover, Φt(W0)W0 for all t0.

    Proof of Proposition 3.1. We assume the existence of ti0 by hypothesis. First, consider the case i=2 so that u2(,t20,ϕ)0. Lemma 1.3 (1) immediately implies that u2(x,t,ϕ)>0 for all xˉΩ and all t>t20; in turn, Lemma 1.3 (3) now implies that u4(x,t,ϕ)>0 for all xˉΩ and t>t20.

    Next, consider the other case of u4(,t40,ϕ)0. This implies that there exists xˉΩ such that u4(x,t40,ϕ)>0. By continuity in space and time, we can assume that xΩ and t40>0 by relabeling if necessary. Then, suppose that u2(,t40,ϕ)0 on ˉΩ. Substituting this in (1.2b) and referencing the strict positivity of u1 for all xˉΩ and all t>0 from Lemma 1.3 (2), we see

    tu2(x,t40)=β2u1(x,t40)u4(x,t40)>0.

    By the same argument in the proof of Lemma 1.4, this contradicts the non-negativity of u2. Hence, instead, we must have u2(,t40,ϕ)0. Thus, by Lemma 1.3 (1) we have u2(x,t,ϕ)>0 for all xˉΩ and all t>t40. In turn, this implies u4(x,t,ϕ)>0 for all (x,t)ˉΩ×(t40,) by Lemma 1.3 (3). This completes the proof of Proposition 3.1 (1).

    Lastly, we will show that Φt(W0)W0 for all t0; the case t=0 is immediate because Φ0 is an identity and thus we focus on t>0. In fact, the proof is an immediate consequence of Proposition 3.1 (1) that we just verified. Let ϕW0 so that ϕ20 or ϕ40. Proposition 3.1 (1) immediately implies that u2(x,t,ϕ) and u4(x,t,ϕ) are both strictly positive on ˉΩ×(0,), and this is more than enough to conclude that Φt(ϕ)W0 for all t>0. This concludes the proof of Proposition 3.1 (2).

    The following result does not necessarily rely on Proposition 3.1; nevertheless, because we redefined W0 and consequently W0 and M in (1.18) and (2.1) differently from [17], we need to reprove it.

    Proposition 3.2. The omega limit set of (1.2) satisfies ω(ϕ)={(mA,0,0,0)} for all ϕM.

    Proof of Proposition 3.2. Let ϕM so that we have Φt(ϕ)W0 for all t0. Thus, u2(,t,ϕ)u4(,t,ϕ)0 for all t0. Then, from (1.2c) we see

    tu3=DΔu3(η+d)u3xΩ

    and thus limtu3(x,t,ϕ)=0 uniformly for xˉΩ by Lemma A.2. Therefore, (1.2a) is asymptotic to

    tY=DΔY+λdYxΩ,

    and hence by the theory of asymptotically autonomous semiflows (see for example [33, Corollary 4.3]) we have limtu1(x,t,ϕ)=mA uniformly for xˉΩ. Thus, we have limtu(x,t,ϕ)=(mA,0,0,0), and it follows that ω(ϕ)={(mA,0,0,0)} for all ϕM.

    The next result is the verification that the DFE is a uniform weak repeller for (1.2) that we described in Section 1.3 to be difficult with choice of W0 and W0 in (1.18).

    Proposition 3.3. If R0>1, then the DFE (mA,0,0,0) is a uniform weak repeller for W0 in the sense that there exists δ0>0 such that

    lim suptΦt(ϕ)(mA,0,0,0)C(ˉΩ)δ0ϕW0. (3.1)

    Proof of Proposition 3.3. By hypothesis, R0>1 so that by Lemma 1.1 we have θ(mA)>0. Suppose for purposes of contradiction that there exists ϕW0 such that for all δ0>0, and hence for δ0(0,mA), we have

    lim supt||Φt(ϕ)(mA,0,0,0)||C(¯Ω)<δ0. (3.2)

    Then, there exists t1>0 sufficiently large so that u1(x,t,ϕ)>mAδ0 for all (x,t)ˉΩ×[t1,). Thus, (1.2b) satisfies

    tu2DΔu2+β1(mAδ0)u2+β2(mAδ0)u4(γ+d)u2(x,t)Ω×[t1,).

    We consider simultaneously

    {tu2DΔu2+β1(mAδ0)u2+β2(mAδ0)u4(γ+d)u2,xΩ,tu4=αu2c(x)u4,xΩ,u2ν=u4ν=0,xΩ, (3.3)

    and

    {t^u2=DΔ^u2+β1(mAδ0)^u2+β2(mAδ0)^u4(γ+d)^u2,xΩ,t^u4=α^u2c(x)^u4,xΩ,^u2ν=^u4ν=0,xΩ, (3.4)

    for tt1 where ν is the outward unit normal vector on Ω. Consider them to have the same values at t1; i.e., (u2,u4)(,t1)=(^u2,^u4)(,t1). The second equations of (3.3) and (3.4) can be solved directly as

    u4(,t)=ec()(tt1)u4(,t1)+tt1αu2(,s)ec()(st)ds, (3.5a)
    ^u4(,t)=ec()(tt1)^u4(,t1)+tt1α^u2(,s)ec()(st)ds. (3.5b)

    Thus, we reduce (3.3) and (3.4) to the following systems for tt1:

    {tu2DΔu2+β1(mAδ0)u2+β2(mAδ0)(ec(tt1)u4(t1)+tt1αu2(s)ec(st)ds)(γ+d)u2xΩ,u2ν=0xΩ,

    and

    {t^u2=DΔ^u2+β1(mAδ0)^u2+β2(mAδ0)(ec(tt1)^u4(t1)+tt1α^u2(s)ec(st)ds)(γ+d)^u2xΩ,^u2ν=0xΩ.

    By the comparison principle ([32, Theorem 7.3.4]), it then follows that

    u2(x,t)^u2(x,t)(x,t)ˉΩ×[t1,). (3.6)

    Concerning the second equations of (3.3) and (3.4), (3.6) and (3.5) show that u4(x,t)^u4(x,t) for all (x,t)ˉΩ×[t1,). As mAδ0>0, we may rely on Lemma 1.1 with HmAδ0 to deduce that the eigenvalue problem corresponding to (3.4) has principle eigenvalue θ(mAδ0) with corresponding eigenfunction ˜ψ:=(~ψ2,~ψ4)0. Because θ(mA)>0, by taking δ0 smaller if needed we obtain θ(mAδ0)>0. Now, by assumption, we have ϕW0 so that ϕi0 for i=2 or i=4. By Proposition 3.1, this implies that uk(x,t,ϕ)>0 for all (x,t)ˉΩ×(0,) for both k{2,4}. Thus, there exists ϵ0>0 sufficiently small such that

    (u2(x,t1,ϕ),u4(x,t1,ϕ))ϵ0eθ(mAδ0)t1˜ψ.

    Finally, as ϵ0eθ(mAδ0)(tt1)˜ψ solves (3.4), we see

    (u2(x,t,ϕ),u4(x,t,ϕ))ϵ0eθ(mAδ0)(tt1)˜ψ(x,t)ˉΩ×[t1,).

    Since θ(mAδ0)>0 and ˜ψ0, it follows that u2(x,t,ϕ),u4(x,t,ϕ) as t, which contradicts (3.2), thereby proving our claim.

    Our next task is a verification that p defined by (1.13) is a generalized distance function for Φt that we described to be difficult in Section 1.3 with our choice of W0 and W0 in (1.18) for the AI model (1.2).

    Proposition 3.4. Define p by (1.13). If R0>1, then

    p1((0,))W0; (3.7)

    moreover, p is a generalized distance function for Φt.

    Proof of Proposition 3.4. First, if ψp1((0,)), then p(ψ)(0,) so that clearly ψW0 by (1.18), concluding (3.7).

    Next, recall from Definition 1.2 that, in order to verify that p is a generalized distance function for Φt, we must prove that

    p(Φt(ϕ))=min{minxˉΩu2(x,t),minxˉΩu3(x,t),minxˉΩu4(x,t)}>0t>0

    whenever either p(ϕ)=0 and ϕW0 or p(ϕ)>0.

    First, suppose p(ϕ)=0 and ϕW0. From (1.18), we deduce that ϕ20 or ϕ40. By Proposition 3.1, we see that u2(x,t,ϕ),u4(x,t,ϕ)>0 for all (x,t)ˉΩ×(0,). We wish to prove that u3(x,t,ϕ)>0 for all (x,t)ˉΩ×(0,). Due to Lemma 1.3 (1), we see that this is achieved if we verify that u3(t)0 for all t>0. Thus, suppose that u3(,t0,ϕ)0 for some t0>0. Then, (1.2c) gives us

    tu3(x,t0)=γu2(x,t0)>0xΩ,

    allowing us to find ϵ(0,t0) sufficiently small such that u3(x,t0ϵ)<0 for any xΩ, which contradicts the non-negativity of the solution. Hence, instead, we must have u3(,t,ϕ)0 for all t>0 due to the arbitrariness of t0>0. Therefore, we have proven that ui(x,t,ϕ)>0 for all xˉΩ, all t>0, and all i{2,3,4}; by (1.13), this implies p(Φt(ϕ))>0 for all t>0.

    Finally, suppose that p(ϕ)>0 so that

    min{minxˉΩϕ2(x),minxˉΩϕ3(x),minxˉΩϕ4(x)}>0

    by (1.13). This implies by Lemma 1.3 (1) that u2(x,t,ϕ),u3(x,t,ϕ)>0 for all xˉΩ and all t>0; consequently, by Lemma 1.3 (3), we have u4(x,t,ϕ)>0 for all xˉΩ and all t>0. Thus, we have that p(Φt(ϕ))>0 for all t>0. Hence, we conclude that p is indeed a generalized distance function for Φt.

    At last, we are ready to conclude the proof of Theorem 2.1.

    Proof of Theorem 2.1. With all the results we have obtained thus far, the proof of Theorem 2.1 follows the same line of reasoning used in previous works (see [17,18]). By Proposition 3.2 we know that any bounded orbit of Φt in M converges to the DFE (mA,0,0,0), which is isolated in X+. If we denote the stable set of the DFE by Ws((mA,0,0,0)), we now see that Ws((mA,0,0,0))W0=. Therefore, considering Proposition 3.2, it follows by [31, Lemma 1.3] (see also [16, Theorem 1.3.2]) that there exists σ>0 such that for all compact chain transitive sets L such that L{(mA,0,0,0)}, we have minϕLp(ϕ)>σ. By Lemma 1.3 (2), taking σ>0 smaller if necessary to satisfy σλΥA for ΥA from (1.21), we see then that

    lim inftui(,t,ϕ)σϕW0,i{1,2,3,4}, (3.8)

    which is precisely (1.14). Finally, by Lemma 3.1, we know that Φt:X+X+ has a global connected attractor A, so from [34, Theorem 3.7 and Remark 3.10], we see that Φt:W0W0 also has a global attractor A0. Because Φt(W0)W0 from Proposition 3.1 (2), and Φt is also κ-condensing by Lemma 3.1, it follows from [34, Theorem 4.7] that Φt has an equilibrium ˜uA0 and hence ˜uW0. It follows immediately from (3.8) that ˜u is a positive steady state of (1.2). This completes the proof of Theorem 2.1.

    Besides the results from List 1.1, Lemmas 1.2 and 1.4, let us recall some results from [18] that will be needed to prove Theorem 2.2.

    Lemma 4.1. ([18, Propositions 4.5 and 4.6 and their proof]) Let D1=D2=:ˉD in (1.4). Then, its semiflow Φt:X+X+ defined by Φt(ϕ)=u(t) guaranteed by List 1.1 (a) is a κ-contraction of order ebt on X+; i.e., κ(ΦtB)ebtκ(B) for all bounded sets BX+ such that κ(B)>0 so that limtκ(ΦtB)=0. Moreover, Φt is κ-condensing and consequently asymptotically smooth. Finally, Φt admits a global connected attractor A.

    The following result is an improvement of Lemma 1.4.

    Proposition 4.1. Let D1=D2=:ˉD in (1.4). Suppose that u(x,t,ϕ) is the solution to (1.4) such that u(,0,ϕ)=ϕ()X+.

    (1) If there exists ti00 such that ui(,ti0,ϕ)0 for any i{2,4,5}, then uk(x,t,ϕ)>0 for all (x,t)ˉΩ×(ti0,) and all k{2,4,5}.

    (2) Moreover, Φt(W0)W0 for all t0.

    Proof of Proposition 4.1. The proof is similar to that of Proposition 3.1. First, consider the case u2(,t20)0. It immediately follows by Lemma 1.4 (1) that u2(x,t)>0 for all (x,t)ˉΩ×(t20,) and hence by Lemma 1.4 (3) that u5(x,t)>0 for all (x,t)ˉΩ×(t20,). Moreover, suppose that u4(,t20)0. The hypothesis that u2(,t20)0 implies that there exists xˉΩ at which u2(x,t20)>0, relabeling if necessary. By continuity in space and time, we may assume that xΩ and t20>0. Then, we have from (1.4d)

    tu4(x,t20)=ξu2(x,t20)+αu5(x,t20)>0,

    indicating that we can find ϵ(0,t20) sufficiently small so that u4(x,t20ϵ)<0, which is a contradiction. Therefore, u4(,t20)0, and by Lemma 1.4 (1) this implies that u4(x,t)>0 for all (x,t)ˉΩ×(t20,).

    Now, consider the case ui(,ti0)0 for i=4 or i=5. This implies that there exists xˉΩ such that ui(x,ti0)>0. By spatial and temporal continuity again, we can assume that xΩ and ti0>0. Suppose that u2(,ti0)0. Then, from (1.4b)

    tu2(x,ti0)=(β2u5(x,ti0)+λu4(x,ti0))u1(x,ti0)>0,

    where the inequality follows by the hypothesis that at least one of u4 or u5 at (x,ti0) is strictly positive, and u1(x,t)>0 for all xˉΩ and t>0 due to Lemma 1.4 (2). Therefore, we can find ϵ(0,ti0) sufficiently small so that u2(x,ti0ϵ)<0, which is again a contradiction. Hence, u2(,ti0)0, which implies by Lemma 1.4 (1) that u2(x,t)>0 for all (x,t)ˉΩ×(ti0,). By the proof in the case u2(,t20)0, we now see that uj(x,t)>0 for all (x,t)ˉΩ×(ti0,) and j{4,5}{i} too. All necessary cases have been demonstrated; hence, Proposition 4.1 (1) was proven.

    Lastly we must show that Φt(W0)W0 for all t0; it suffices to prove the case t>0 as Φ0 is an identity. Let ϕW0 so that ϕ20 or ϕ40 or ϕ50. Proposition 4.1 (1) implies that ui(x,t,ϕ)>0 on ˉΩ×(0,) for all i{2,4,5}. This is more than enough to deduce that Φt(ϕ)W0 for all t>0, concluding the proof of Proposition 4.1 (2).

    The following result does not necessarily rely on Proposition 4.1; nevertheless, because we redefined W0 and consequently W0 and M in (1.19) and (2.1) differently from [18], it needs to be proven.

    Proposition 4.2. Let D1=D2=:ˉD in (1.4). Then, the omega limit set of (1.4) satisfies ω(ϕ)={(mE,0,0,0,0)} for all ϕM.

    Proof of Proposition 4.2. The proof is similar to that of Proposition 3.2; thus, we leave it in the Appendix for completeness.

    The next result is a verification that the DFE is a uniform weak repeller for (1.4) that we described in Section 1.3 to be difficult with the choice of W0 and W0 in (1.18).

    Proposition 4.3. If D1=D2=:ˉD and R0>1, then the DFE (mE,0,0,0,0) is a uniform weak repeller for W0 in the sense that there exists δ0>0 such that

    lim suptΦt(ϕ)(mE,0,0,0,0)C(ˉΩ)δ0ϕW0. (4.1)

    Proof of Proposition 4.3. Proof is similar to that of Proposition 3.3. By hypothesis, R0>1, so by Lemma 1.2 we have that θ(mE)>0. Suppose for purposes of contradiction that there exists ϕW0 such that, for all δ0>0, and hence for δ0(0,mE), we have

    lim supt||Ψt(ϕ)(mE,0,0,0,0)||C(ˉΩ)<δ0. (4.2)

    Then, there exists t1>0 sufficiently large so that u1(x,t,ϕ)>mEδ0 for all (x,t)ˉΩ×[t1,). Thus, (1.4b) leads us to consider simultaneously

    {tu2ˉDΔu2+(β1u2+β2u5+λu4)(mEδ0)(μ+δ+γ)u2,xΩ,tu4=D4Δu4+ξu2+αu5ηu4,xΩ,tu5=δu2bu5,xΩ,u2ν=u4ν=u5ν=0,xΩ, (4.3)

    and

    {t^u2=ˉDΔ^u2+(β1^u2+β2^u5+λ^u4)(mEδ0)(μ+δ+γ)^u2,xΩ,t^u4=D4Δ^u4+ξ^u2+α^u5η^u4,xΩ,t^u5=δ^u2b^u5,xΩ,^u2ν=^u4ν=^u5ν=0,xΩ, (4.4)

    with tt1, starting from same values at t1; i.e., (u2,u4,u5)(,t1)=(^u2,^u4,^u5)(,t1). Solving for u5 and ^u5 as

    u5(,t)=u5(,t1)eb(t1t)+tt1δu2(,s)eb(st)ds,^u5(,t)=^u5(,t1)eb(t1t)+tt1δ^u2(,s)eb(st)ds,

    and using the comparison theorem argument give us

    u2(x,t)^u2(x,t),u4(x,t)^u4(x,t), and u5(x,t)^u5(x,t)(x,t)ˉΩ×[t1,). (4.5)

    As mEδ0>0, we may rely on Lemma 1.2 with HmEδ0 to deduce that the eigenvalue problem corresponding to (4.4) has principle eigenvalue θ(mEδ0) with corresponding eigenfunction ˜ψ:=(~ψ2,~ψ4,~ψ5)0. Because θ(mE)>0, by taking δ0 smaller if needed we obtain θ(mEδ0)>0. By assumption, we have ϕW0 so that ϕi0 for i=2 or i=4 or i=5. By Proposition 4.1, this implies that uk(x,t,ϕ)>0 for all (x,t)ˉΩ×(0,) for all k{2,4,5}. Therefore, there exists ϵ0>0 sufficiently small such that

    (u2(x,t1,ϕ),u4(x,t1,ϕ),u5(x,t1,ϕ))ϵ0eθ(mEδ0)t1˜ψ.

    Finally, as ϵ0eθ(mEδ0)(tt1)˜ψ solves (4.4), we see

    (u2(x,t,ϕ),u4(x,t,ϕ),u5(x,t,ϕ))ϵ0eθ(mEδ0)(tt1)˜ψ(x,t)ˉΩ×[t1,).

    Since θ(mEδ0)>0 and ˜ψ0, it follows that u2(x,t,ϕ),u4(x,t,ϕ),u5(x,t,ϕ) as t, which contradicts (4.2), thereby proving our claim.

    Our next task is a verification that p defined by (1.16) is a generalized distance function for Φt that we described in Section 1.3 to be difficult with our choice of W0 and W0 in (1.19) for the EVD model (1.4).

    Proposition 4.4. Define p by (1.16). If D1=D2=:ˉD and R0>1, then

    p1((0,))W0; (4.6)

    moreover, p is a generalized distance function for Φt.

    Proof of Proposition 4.4. The proof is similar to that of Proposition 3.4. First, if ψp1((0,)), then p(ψ)(0,) so that clearly ψW0 by (1.19), concluding (4.6).

    Next, we must prove that p(Φt(ϕ))>0 for all t>0 whenever either p(ϕ)=0 and ϕW0 or p(ϕ)>0.

    First, suppose p(ϕ)=0 and ϕW0. From (1.19), we deduce that ϕ20 or ϕ40 or ϕ50. By Proposition 4.1 we see that u2(x,t,ϕ),u4(x,t,ϕ),u5(x,t,ϕ)>0 for all (x,t)ˉΩ×(0,). To prove that u3(x,t,ϕ)>0 for all (x,t)ˉΩ×(0,), due to Lemma 1.4 (1), it suffices to verify that u3(t)0 for all t>0. Suppose that u3(,t0,ϕ)0 for some t0>0. Then, (1.4c) gives us

    tu3(x,t0)=γu2(x,t0)>0xΩ,

    allowing us to find ϵ(0,t0) sufficiently small such that u3(x,t0ϵ)<0 for any xΩ, which contradicts the non-negativity of the solution. Hence, instead, we must have u3(,t,ϕ)0 for all t>0 due to the arbitrariness of t0>0. Therefore, we have proven that ui(x,t,ϕ)>0 for all xˉΩ, all t>0 and all i{2,3,4,5} so that p(Φt(ϕ))>0 for all t>0 by (1.16).

    At last, suppose that we have p(ϕ)>0. Then, by Lemma 1.4 (1), we can obtain u2(x,t,ϕ),u3(x,t,ϕ),u4(x,t,ϕ)>0 for all xˉΩ and all t>0. Consequently by Lemma 1.4 (3), we have u5(x,t,ϕ)>0 for all xˉΩ and all t>0. Thus, we have that p(Φt(ϕ))>0 for all t>0. Therefore, we conclude that p is indeed a generalized distance function for Φt.

    As we mentioned, with all the results obtained thus far, the proof of Theorem 2.2 follows the same line of reasoning in the literature, and it is similar to the proof of Theorem 2.1. Thus, we leave this in the Appendix for completeness.

    As PDE models of infectious diseases continue to attract attention from researchers, taking into account distinct mobilities among the solution components and therefore considering partially diffusive systems of PDEs seems inevitable. In this work, we improved uniform persistence results of such systems of PDEs, namely the AI model from [17] and the EVD model from [18]. To do so, we weaken the assumption of initial data so that any (rather than all) of the infection-related components must be initially non-vanishing. The mathematical novelty is the discovery of the ability of the solution to a non-diffusive equation that does not vanish at time t00 to induce strict positivity on ˉΩ for all t>t0 on not only itself but all other infection-related components (see Propositions 3.1 and 4.1). This takes the place of the minimum principle, which is standard to apply in a fully diffusive system. Specifically, the key in both cases is to make use of the fact that S(x,t)>0 for all xˉΩ and all t>0 and verify that, as a consequence, I(x,t)>0 for all xˉΩ and all t>0. Ultimately, this closes the gap between the uniform persistence results of the fully diffusive systems of PDEs and the partially diffusive systems of PDEs, with such results having previously been known for the former but not the latter. Our approach seems general and applies to other partially diffusive models such as the COVID-19 model that needs to take into account of the environmental modes of transmission.

    The authors declare they have not used Artificial Intelligence (AI) tools in the creation of this article.

    Yamazaki is supported by the Simons Foundation (962572, KY). The authors thank Prof. Jin Wang for motivating discussions.

    The authors declare no conflict of interest.

    The following well-known fact proved to be useful.

    Lemma A.1. Let f:UR be twice differentiable, where U is an open subset of Rn. If f has a local minimum at aU, then Δf(a)0.

    We recall the definition of Kuratowski measure of non-compactness:

    Definition A.1. ([16, p. 3]) Given any metric space Y, the Kuratowski measure of non-compactness for any bounded set B of Y is defined by

    κ(B):=inf{r:B has a finite cover of diameter r}.

    We refer to [16, p. 3], [35, Proposition 7.2] and [36, Lemma 2.3.5] for various properties concerning κ.

    Lemma A.2. ([37, Proposition 4.1]) Consider a spatial domain ΩRn for nN that is bounded with smooth boundary Ω and the following equation:

    tw(x,t)=˜D(x)Δw(x,t)(¯U(x))w(x,t)+g(x)λw(x,t),(n)w(x,t)|Ω=0fort>0,andw(x,0)=ψ(x)forx¯Ω,

    where ˉUC2(ˉΩ), ˜D(x) is continuous and ˜D(x)M>0 for all x¯Ω, g(x)>0 is a continuous function, and n is an outward unit normal vector. Then, for all ψC(¯Ω,R+), there exists a unique positive steady state w which is globally attractive in C(¯Ω,R). Moreover, if g(x)g, then wgλ.

    Let ϕM so that we have Φt(ϕ)W0 for all t0. Thus, u2(,t,ϕ)u4(,t,ϕ)u5(,t,ϕ)0 for all t0. From (1.4c), we see

    tu3=D3Δu3μu3

    and thus limtu3(x,t,ϕ)=0 uniformly for xˉΩ due to Lemma A.2. Therefore, now (1.4a) is asymptotic to

    tY=ˉDΔY+πμY

    and hence by the theory of asymptotically autonomous semiflows (see [33, Corollary 4.3]) we have limtu1(x,t,ϕ)=mE uniformly for all xˉΩ. Thus, we see limtu(x,t,ϕ)=(mE,0,0,0,0), and it follows that ω(ϕ)={(mE,0,0,0,0)}.

    By Proposition 4.2 we know that any bounded orbit of Φt in M converges to the DFE (mE,0,0,0,0), which is isolated in X+. Considering Proposition 4.2, it follows by [31, Lemma 3] (see also [16, Theorem 1.3.2]) that there exists σ>0 such that for all compact chain transitive sets L such that L{(mE,0,0,0,0)}, we have minϕLp(ϕ)>σ. By Lemma 1.4 (2), taking σ>0 smaller if necessary to satisfy σπΥE for ΥE from (1.23), we see then that

    lim inftui(,t,ϕ)σϕW0,i{1,2,3,4,5}, (C.1)

    which is precisely (1.17). Finally, by Lemma 4.1 we know that Φt:X+X+ has a global connected attractor A. Thus, we see from [34, Theorem 3.7 and Remark 3.10] that Φt:W0W0 also has a global attractor A0. Because Φt(W0)W0 from Proposition 4.1 (2), and Φt is also κ-condensing by Lemma 4.1, it follows from [34, Theorem 4.7] that Φt has an equilibrium ˜uA0 and hence ˜uW0. It follows from (C.1) that ˜u is a positive steady state of (1.4).



    [1] P. van den Driessche, J. Watmough, Reproduction numbers and sub-threshold endemic equilibria for compartmental models of disease transmission, Math. Biosci., 180 (2002), 29–48. https://doi.org/10.1016/S0025-5564(02)00108-6 doi: 10.1016/S0025-5564(02)00108-6
    [2] H. R. Thieme, Spectral bound and reproduction number for infinite-dimensional population structure and time heterogeneity, SIAM J. Appl. Math., 70 (2009), 188–211. https://doi.org/10.1137/080732870 doi: 10.1137/080732870
    [3] W. Wang, X. Q. Zhao, Basic reproduction numbers for reaction-diffusion epidemic models, SIAM J. Appl. Dyn. Syst., 11 (2012), 1652–1673. https://doi.org/10.1137/120872942 doi: 10.1137/120872942
    [4] K. Yamazaki, Global well-posedness of infectious disease models without life-time immunity: the cases of cholera and avian influenza, Math. Med. Biol., 35 (2018), 427–445. https://doi.org/10.1093/imammb/dqx016 doi: 10.1093/imammb/dqx016
    [5] H. M. Yin, On a reaction-diffusion system modelling infectious diseases without lifetime immunity, Eur. J. Appl. Math., 33 (2021), 803–827. https://doi.org/10.1017/S0956792521000231 doi: 10.1017/S0956792521000231
    [6] W. Wang, X. Q. Zhao, A nonlocal and time-delayed reaction-diffusion model of dengue transmission, SIAM J. Appl. Math., 71 (2011), 147–168. https://doi.org/10.1137/090775890 doi: 10.1137/090775890
    [7] Y. Lou, X. Q. Zhao, A reaction-diffusion malaria model with incubation period in the vector population, J. Math. Biol., 62 (2011), 543–568. https://doi.org/10.1007/s00285-010-0346-8 doi: 10.1007/s00285-010-0346-8
    [8] X. Q. Zhao, Global dynamics of a reaction and diffusion model for Lyme disease, J. Math. Biol., 65 (2012), 787–808. https://doi.org/10.1007/s00285-011-0482-9 doi: 10.1007/s00285-011-0482-9
    [9] K. Yamazaki, X. Wang, Global well-posedness and asymptotic behavior of solutions to a reaction-convection-diffusion cholera epidemic model, Discrete Contin. Dyn. Syst. Ser. -B, 21 (2016), 1297–1316. https://doi.org/10.3934/dcdsb.2016.21.1297 doi: 10.3934/dcdsb.2016.21.1297
    [10] K. Yamazaki, X. Wang, Global stability and uniform persistence of the reaction-convection-diffusion cholera epidemic model, Math. Biosci. Eng., 14 (2017), 559–579. https://doi.org/10.3934/mbe.2017033 doi: 10.3934/mbe.2017033
    [11] K. Yamazaki, Zika virus dynamics partial differential equations model with sexual transmission route, Nonlinear Anal. Real World Appl., 50 (2019), 290–315. https://doi.org/10.1016/j.nonrwa.2019.05.003 doi: 10.1016/j.nonrwa.2019.05.003
    [12] S. B. Hsu, J. Jiang, F. B. Wang, On a system of reaction-diffusion equations arising from competition with internal storage in an unstirred chemostat, J. Differ. Equations, 248 (2010), 2470–2496. https://doi.org/10.1016/j.jde.2009.12.014 doi: 10.1016/j.jde.2009.12.014
    [13] S. B. Hsu, F. B. Wang, X. Q. Zhao, Dynamics of a periodically pulsed bio-reactor model with a hydraulic storage zone, J. Dyn. Differ. Equations, 23 (2011), 817–842. https://doi.org/10.1007/s10884-011-9224-3 doi: 10.1007/s10884-011-9224-3
    [14] J. P. Grover, S. B. Hsu, F. B. Wang, Competition and coexistence in flowing habitats with a hydraulic storage zone, Math. Biosci., 222 (2009), 42–52. https://doi.org/10.1016/j.mbs.2009.08.006 doi: 10.1016/j.mbs.2009.08.006
    [15] S. B. Hsu, F. B. Wang, X. Q. Zhao, Global dynamics of zooplankton and harmful algae in flowing habitats, J. Differ. Equations, 255 (2013), 265–297. https://doi.org/10.1016/j.jde.2013.04.006 doi: 10.1016/j.jde.2013.04.006
    [16] X. Q. Zhao, Dynamical Systems in Population Biology, Springer-Verlag, New York, Inc., 2003. https://doi.org/10.1007/978-0-387-21761-1
    [17] N. K. Vaidya, F. B. Wang, X. Zou, Avian influenza dynamics in wild birds with bird mobility and spatial heterogeneous environment, Discrete Contin. Dyn. Syst. Ser. -B, 17 (2012), 2829–2848. https://doi.org/10.3934/dcdsb.2012.17.2829 doi: 10.3934/dcdsb.2012.17.2829
    [18] K. Yamazaki, Threshold dynamics of reaction-diffusion partial differential equations model of Ebola virus disease, Int. J. Biomath., 11 (2018), 1850108. https://doi.org/10.1142/S1793524518501085 doi: 10.1142/S1793524518501085
    [19] Z. J. Cheng, J. Shan, Novel coronavirus: where we are and what we know, Infection, 48 (2020), 155–163. https://doi.org/10.1007/s15010-020-01401-y doi: 10.1007/s15010-020-01401-y
    [20] C. Yang, J. Wang, A mathematical model for the novel coronavirus epidemic in Wuhan, China, Math. Biosci. Eng., 17 (2020), 2708–2724. https://doi.org/10.3934/mbe.2020148 doi: 10.3934/mbe.2020148
    [21] J. D. Brown, G. Goekjian, R. Poulson, S. Valeika, D. E. Stallknecht, Avian influenza virus in water: Infectivity is dependent on pH, salinity and temperature, Vet. Microbiol., 136 (2009), 20–26. https://doi.org/10.1016/j.vetmic.2008.10.027 doi: 10.1016/j.vetmic.2008.10.027
    [22] V. S. Hinshaw, R. G. Webster, B. Turner, The perpetuation of orthomyxoviruses and paramyxoviruses in Canadian waterfowl, Can. J. Microbiol., 26 (1980), 622–629. https://doi.org/10.1139/m80-108 doi: 10.1139/m80-108
    [23] R. G. Webster, M. Yakhno, V. S. Hinshaw, W. J. Bean, K. G. Murti, Intestinal influenza: replication and characterization of influenza viruses in ducks, Virology, 84 (1978), 268–278. https://doi.org/10.1016/0042-6822(78)90247-7 doi: 10.1016/0042-6822(78)90247-7
    [24] World Health Organization, Ebola virus disease, 2023. Available from: https://www.who.int/news-room/fact-sheets/detail/ebola-virus-disease.
    [25] L. Evans, Partial Differential Equations, American Mathematics Society, Providence, Rhode Island, 1998.
    [26] R. Covington, S. Patton, E. Walker, K. Yamazaki, Stability analysis on a partially diffusive model of the coronavirus disease of 2019, submitted.
    [27] W. Puryear, K. Sawatzki, N. Hill, A. Foss, J. J. Stone, L. Doughty, et al., Pathogenic Avian Influenza A(H5N1) virus outbreak in New England Seals, United States, Emerging Infect. Dis., 29 (2023), 786–791. https://doi.org/10.3201/eid2904.221538 doi: 10.3201/eid2904.221538
    [28] T. Berge, J. M. S. Luburma, G. M. Moremedi, N. Morris, R. Kondera-Shava, A simple mathematical model for Ebola in Africa, J. Biol. Dyn., 11 (2017), 42–74. https://doi.org/10.1080/17513758.2016.1229817 doi: 10.1080/17513758.2016.1229817
    [29] World Health Organization, Ebola disease caused by Sudan ebolavirus - Uganda, 2023. Available from: https://www.who.int/emergencies/disease-outbreak-news/item/2023-DON433.
    [30] C. Freeman, Meet the world's bravest undertakers - Liberia's Ebola burial squad, The Telegraph, 2014. Available from: https://www.telegraph.co.uk/news/worldnews/ebola/11024042/Meet-the-worlds-bravest-undertakers-Liberias-Ebola-burial-squad.html.
    [31] H. L. Smith, X. Q. Zhao, Robust persistence for semidynamical systems, Nonlinear Anal. Theory Methods Appl., 47 (2001), 6169–6179. https://doi.org/10.1016/S0362-546X(01)00678-2 doi: 10.1016/S0362-546X(01)00678-2
    [32] H. L. Smith, Monotone Dynamical Systems: An Introduction to the Theory of Competitive and Cooperative Systems, Mathematical Surveys and Monographs, American Mathematical Society, Providence, Rhode Island, 41 (1995).
    [33] H. R. Thieme, Convergence results and a Poincarˊe-Bendixson trichotomy for asymptotically autonomous differential equations, J. Math. Biol., 30 (1992), 755–763. https://doi.org/10.1007/BF00173267 doi: 10.1007/BF00173267
    [34] P. Magal, X. Q. Zhao, Global attractors and steady states for uniformly persistent dynamical systems, SIAM J. Math. Anal., 37 (2005), 251–275. https://doi.org/10.1137/S0036141003439173 doi: 10.1137/S0036141003439173
    [35] K. Deimling, Nonlinear Functional Analysis, Dover Publications, Inc., Mineola, New York, 1985. https://doi.org/10.1007/978-3-662-00547-7
    [36] J. K. Hale, Asymptotic Behavior of Dissipative Systems, Mathematical Surveys and Monographs, American Mathematics Society, Providence, Rhode Island, 1988. https://doi.org/10.1090/surv/025
    [37] J. Wang, C. Yang, K. Yamazaki, A partially diffusive cholera model based on a general second-order differential operator, J. Math. Anal. Appl., 501 (2021), 125181. https://doi.org/10.1016/j.jmaa.2021.125181 doi: 10.1016/j.jmaa.2021.125181
  • Reader Comments
  • © 2023 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(2194) PDF downloads(56) Cited by(0)

Figures and Tables

Tables(2)

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog