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

Can infectious diseases eradicate host species? The effect of infection-age structure

  • Dedicated to Stephan Luckhaus on the occasion of his 70th birthday.
  • It is a fundamental question in mathematical epidemiology whether deadly infectious diseases only lead to a mere decline of their host populations or whether they can cause their complete disappearance. Upper density-dependent incidences do not lead to host extinction in simple, deterministic SI or SIS (susceptible-infectious) epidemic models. Infection-age structure is introduced into SIS models because of the biological accuracy offered by considering arbitrarily distributed infectious periods. In an SIS model with infection-age structure, survival of the susceptible host population is established for incidences that depend on the infection-age density in a general way. This confirms previous host persistence results without infection-age for incidence functions that are not generalizations of frequency-dependent transmission. For certain power incidences, hosts persist if some infected individuals leave the infected class and become susceptible again and the return rate dominates the infection-age dependent infectivity in a sufficient way. The hosts may be driven into extinction by the infectious disease if there is no return into the susceptible class at all.

    Citation: Joan Ponce, Horst R. Thieme. Can infectious diseases eradicate host species? The effect of infection-age structure[J]. Mathematical Biosciences and Engineering, 2023, 20(10): 18717-18760. doi: 10.3934/mbe.2023830

    Related Papers:

    [1] Maria Guadalupe Vazquez-Peña, Cruz Vargas-De-León, Jorge Velázquez-Castro . Global stability for a mosquito-borne disease model with continuous-time age structure in the susceptible and relapsed host classes. Mathematical Biosciences and Engineering, 2024, 21(11): 7582-7600. doi: 10.3934/mbe.2024333
    [2] 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
    [3] 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
    [4] Paul L. Salceanu . Robust uniform persistence in discrete and continuous dynamical systems using Lyapunov exponents. Mathematical Biosciences and Engineering, 2011, 8(3): 807-825. doi: 10.3934/mbe.2011.8.807
    [5] Hongbin Guo, Michael Yi Li . Global dynamics of a staged progression model for infectious diseases. Mathematical Biosciences and Engineering, 2006, 3(3): 513-525. doi: 10.3934/mbe.2006.3.513
    [6] Abulajiang Aili, Zhidong Teng, Long Zhang . Dynamical behavior of a coupling SEIR epidemic model with transmission in body and vitro, incubation and environmental effects. Mathematical Biosciences and Engineering, 2023, 20(1): 505-533. doi: 10.3934/mbe.2023023
    [7] Yan-Xia Dang, Zhi-Peng Qiu, Xue-Zhi Li, Maia Martcheva . Global dynamics of a vector-host epidemic model with age of infection. Mathematical Biosciences and Engineering, 2017, 14(5&6): 1159-1186. doi: 10.3934/mbe.2017060
    [8] Peter Rashkov, Ezio Venturino, Maira Aguiar, Nico Stollenwerk, Bob W. Kooi . On the role of vector modeling in a minimalistic epidemic model. Mathematical Biosciences and Engineering, 2019, 16(5): 4314-4338. doi: 10.3934/mbe.2019215
    [9] Britnee Crawford, Christopher Kribs-Zaleta . A metapopulation model for sylvatic T. cruzi transmission with vector migration. Mathematical Biosciences and Engineering, 2014, 11(3): 471-509. doi: 10.3934/mbe.2014.11.471
    [10] Ali Moussaoui, Vitaly Volpert . The impact of immune cell interactions on virus quasi-species formation. Mathematical Biosciences and Engineering, 2024, 21(11): 7530-7553. doi: 10.3934/mbe.2024331
  • It is a fundamental question in mathematical epidemiology whether deadly infectious diseases only lead to a mere decline of their host populations or whether they can cause their complete disappearance. Upper density-dependent incidences do not lead to host extinction in simple, deterministic SI or SIS (susceptible-infectious) epidemic models. Infection-age structure is introduced into SIS models because of the biological accuracy offered by considering arbitrarily distributed infectious periods. In an SIS model with infection-age structure, survival of the susceptible host population is established for incidences that depend on the infection-age density in a general way. This confirms previous host persistence results without infection-age for incidence functions that are not generalizations of frequency-dependent transmission. For certain power incidences, hosts persist if some infected individuals leave the infected class and become susceptible again and the return rate dominates the infection-age dependent infectivity in a sufficient way. The hosts may be driven into extinction by the infectious disease if there is no return into the susceptible class at all.



    It is a fundamental question in mathematical epidemiology whether deadly infectious diseases only lead to a mere decline of their host populations or whether they can also cause their complete disappearance. This question has become particularly relevant for the possible role of infectious diseases in amphibian decline and disappearance (see [1,2] and the references therein).

    Deterministic mathematical models have shown the importance of how the disease incidence, the rate of new infections, denoted by f(S,I), depends on the numbers of susceptible, S, and infected individuals, I. The function f is called the incidence function.

    Density dependent (or mass action) incidence f(S,I)=SI typically does not lead to host extinction [3,4] unless there is a reservoir for the disease [5,6,7] or the population is subject to an Allee effect [8,9,10].

    Frequency dependent incidence is of the form f(S,I)=SIN, where N is the total population size, and f(S,I)=SIS+I if the population only consists of susceptible and infective individuals (i.e. the model is of SI or SIS type). Infectious disease models with frequency-dependent (or standard) incidence are mathematically known to show both decline and disappearance of the host in different parameter regions, [4,11,12,13,14,15,16,17,18,19,20,21], and [22, Sec.4.4]. However, empirical evidence of a species or population eliminated by frequency-dependent parasites seems elusive [23,24].

    Many more incidence forms have been suggested, and many of those, though not all, have been fitted to the infection experiments involving tiger salamander larvae and ranavirus in [25]. See also [1]. Both density- and frequency-dependent incidence do not fare well in these fits.

    In [1] and [26], epidemic models are considered that consist of two ordinary differential equations for the densities of susceptible and infective hosts; they have general incidence functions and rather general host population dynamics. In [1], all infective hosts die from the disease (unless they die from natural causes first), while in [26] some infective hosts recover from the disease and become susceptible again.

    Two types of incidence functions were identified:

    ● Incidence functions that are mildly upper density-dependent,

    f(0,I)=0=f(S,0),S,I0,f(S,I)S+I0,S,I0,0<S+I0, (1.1)

    (i.e., f is differentiable at (0,0) with zero derivative).

    ● Incidence functions that are positively homogeneous of degree one.

    An incidence function f is positively homogeneous of degree ζ>0 [27, Sec.6.53] if

    f(rS,rI)=rζf(S,I),S,I,r0. (1.2)

    The density-dependent incidence f(S,I)=SI is mildly upper density-dependent, while the frequency-dependent incidence f(S,I)=SIS+I is homogeneous of degree one.

    In these ODE models, incidence functions that are mildly upper density-dependent lead to mere (though possibly substantial) host decline (with one subtlety for certain power incidences), while incidence functions that are positively homogeneous of degree one can cause host extinction.

    For power incidence functions f(S,I)=SqIp with p,q>0, q<1<q+p (which are mildly upper density-dependent), though, it makes a big difference whether or not infective individuals can become susceptible again. If the disease is universally fatal, there is always initial-value dependent extinction of both host and disease [1, Thm.5.8]. If infective individuals can become susceptible again, both host and disease always persist [26].

    In these ODE models, infective individuals die or become susceptible again at constant per capita rates so that the sojourn of hosts in the infectious stage (which is influenced by the extra death rate due the disease and, possibly, by the return rate into the susceptible class) is exponentially distributed. It is long known that this is far from reality. Sardwell [28,29], for instance, has fitted lognormal distributions to the length of incubation periods of various diseases. The distribution of the time from infection to disease death in infection experiments involving salamander larvae and ranavirus [25] looks very different from an exponential distribution [2]. A lognormal distribution fits better than a Gamma distribution which in turn fits better than a Weibull distribution. The case has been made that assuming exponential stage distributions rather than more realistic distributions can lead to wrong epidemiological conclusions [30].

    We therefore will study whether the feature that mildly upper density-dependent incidence functions do not lead to host disappearance remains valid if the length of the infectious period has an arbitrary distribution. In other words, we will introduce infection-age structure into the model. Such model have already been considered by Kermack and McKendrick [31,32] though they apparently did not use the term "infection-age" but talked about "the duration of the illness". They also (as we will) let the infectivity of infective hosts depend on their infection-age ("The infectivity of the disease is a function of the period of illness"), which turned out to be an important feature of HIV infection [33].

    Kermack and McKendrick [31,32] even let the susceptibility of a recovered host be a "function of the period which has elapsed since the time of" recovery (of the recovery-age) which we will not assume. See [34, Sec.8.2] [35, 13.3] [36]). Infection-age and recovery-age are special cases of epidemiological class-age [37]. Covid-19 models with two or more infection-ages have been considered for households of two hosts [38] and for groups with successive multiple infections [39]. Class-age needs to be distinguished from chronological age, time since birth. While there can be good reason to consider both age types [34, Sec.6.5] [40, Sec.5.3.2], we will concentrate on infection-age in this paper. One reason is that, if both type of ages are considered, our message will drown in technicalities. Another is that chronological age may be epidemiologically less important for some animal populations (like ungulates affected by an infectious eye disease [41], see end of Section 1.6) than for human populations. For amphibian populations, different developmental stages would be more important [42].

    Finally, let us mention that Kermack and McKendrick [31] did not only already consider epidemic outbreaks like in [43,44] but also the "Problem of Endemicity" which in our model is caused by a per capita host growth rate.

    Let P(a) denote the probability that a previously infected host is still infected (i.e., has not died from the disease or from other causes and not returned into the susceptible stage) at infection age a, i.e. a time units after infection with the disease. Then the sojourn function P:R+[0,1] of the infected stage is decreasing,

    P(0)=1,D(0)=0P(a)da<. (1.3)

    D(0) is the expected duration of the infected stage. See [45, Chap.12] [2].

    There are two cases:

    (i) P(a)>0 for all aR+.

    (ii) There is some b(0,) (which is uniquely determined) such that P(a)>0 for a[0,b) and P(a)=0 for a(b,).

    If b(0,), b has the interpretation of the maximal infection age.

    To consider both cases simultaneously, we allow b=. If the duration of the infected stage is Gamma, lognormal or Weibull distributed, b=. For numerical purposes, it may be convenient to assume b(0,).

    Let I(t) denote the number of infected hosts at time t0 where t=0 marks the start of the epidemic. If infection-age is considered,

    I(t)=b0u(t,a)da,t0, (1.4)

    where u(t,a), as a function of a, is the infection-age density, loosely speaking the number of infected hosts at time t with infection-age a, i.e., those hosts that have been infected a time units ago.

    The number of infected hosts needs to be distinguished from the infective impact,

    J(t)=b0κ(a)u(t,a)da, (1.5)

    where κ(a) is the infectivity of an infected host with infection-age a, for instance the amount of infective particles that an infective host releases. Our notation follows [46, chap.9]. In [26], i has been used instead of u and W instead of J. In [2], I and J have been switched and j been used instead of u.

    The incidence function f provides the functional relationship between the rate of new infections at time t, B(t), on the one hand and the number of susceptible hosts, S(t), and the infection-age density of infected hosts, u(t,), on the other hand,

    B(t)=f(S(t),u(t,)). (1.6)

    Here, f:R+×L1+[0,b)R+. X=L1[0,b) is the Banach space of Lebesgue integrable functions on [0,b) with values in [,] (more precisely of the appropriate a.e. equivalence classes) and X+=L1+[0,b) is the cone of nonnegative functions in X. The norm on X is w1=b0|w(a)|da, wX+, which is the number of infected hosts if wX+.

    We assume that f:R+×X+R is continuous, where the norm on R×X is (S,w)=|S|+w1, SR+ and wX+, which is the number of hosts.

    If B is the disease incidence, the rate of new infections, and u0X+ is the density of infected host at time 0, we have the relation

    u(t,a)={B(ta)P(a),0a<tu0(at)P(a)P(at),0t<a}t0,a[0,b). (1.7)

    P(a)P(at) is the conditional probability to be still infected at infection age a provided one was already infected at infection-age at. See [45, Sec.13.1].

    If B is continuous and u0X+, then u(t,) is a continuous function of t with values in X+ [2] (Lemma 7.1).

    The straightforward extension of the density-dependent incidence function to infection-age models is

    f(S,w)=SJ,J=b0κ(a)w(a)da,SR+,wX+, (1.8)

    with infective impact J and infection-age dependent per capita infectivity κL+[0,b). Notice that if u(t,) is a continuous function of t with values in X+, then the infective impact J(t)=b0κ(a)u(t,a)da is also a continuous function of t.

    L[0,b) is the Banach space of essentially bounded Borel measurable functions (more precisely, their equivalence classes) with the essential supremum norm w, wL[0,b).

    Density-dependent incidence yields a very poor fit in [1,25,47]. Another possible incidence function is

    f(S,w)=SJ/(1+kI+1+2kI),SR+,wX+,I=b0w(a)da,J=b0κ(a)w(a)da. (1.9)

    In the special case without infection age, where J=σI, this incidence function is an adaptation of the contact incidence in [48] and performs quite well in the fitting comparisons in [1,47]. Notice that, compared to [2], J and I have been switched. In [26], incidence functions

    f(S,w)=SJh(S+I),I=b0w(a)da,J=b0κ(a)w(a)da, (1.10)

    with a decreasing continuous function h:R+R+ were considered. Cf. [2, (1.6)].

    Since we will consider general incidence functions in this paper, survival of the susceptible host population is established in the form of uniform weak persistence and not in form of the more desirable uniform persistence.

    Uniform persistence means that there is an initial-value independent level above which the susceptible host density settles given enough time:

    There is some ϵ>0 such that lim inftS(t)ϵ for any nonnegative solution on R+ with S(0)>0.

    Uniform weak persistence (in which the lim inf is replaced by the lim sup) means that there is an initial-value independent level above which the host density always returns though it can become arbitrarily low at times.

    Definition 1.1. The susceptible hosts persist uniformly weakly if there exists some ϵ>0 such that for all solutions of our model system on R+ the number of susceptible hosts S(t) at time t0 satisfies the following implication:

    S(0)>0lim suptS(t)>ϵ.

    In detailed words, the susceptible hosts S persist uniformly weakly if there is some ϵ>0 such that, whenever S(0)>0 and s>0, there is some t>s such that S(t)>ϵ.

    For general persistence concepts, see [46,49,50] and the references therein. See also Section 8.1 about possible roads from uniform weak to uniform persistence. To give a taste of what we want to accomplish in this paper, we present a special case of Theorem 4.1 which covers many incidence functions, in particular those we just mentioned. More general results would involve more terminology than we want to present in the introduction (Section 2.5).

    Corollary 1.2. Assume that there exists a function ˜f:R+×X+R+ such that

    f(S,w)=S˜f(S,w),S,w0,SR+,wX+,

    where ˜f is continuous at (0,0) and is bounded on bounded sets and ˜f(0,0)=0.

    Then the susceptible hosts persist uniformly weakly.

    The function ˜f is often called the force of infection [51] [52, Sec.1] [53,54] [35, 6.2.2], not to be confused with infective impact, (1.5), which is a term specific for infection-age structured models; for density-dependent incidence, the two can coincide [34, Sec.5.3].

    Once again, the Laplace transform will be a useful tool for proving uniform weak persistence in age-structured population models [45,46,55]. Corollary 1.2 follows from Corollary 2.5.

    Example 1.3. Let

    f(S,w)=SJ11+κ0S+J2,Jj=b0κj(a)w(a)da,j=1,2.

    Here κ0(0,) is the mixing activity of a susceptible host. κ2(a) is the mixing activity of an infected host, κ1(a)=σ(a)κ2(a) is the effective infectivity of an infected host at infection-age a and σ(a) is the amount of infectious particles that an infected host releases at infection-age a. We assume that σ,κjL+[0,b), j=1,2.

    The fraction J21+κ0S+J2 is the chance that a random contact of a susceptible host actually occurs with an infected host and J1/J2 is the average effective infectivity of active infected hosts. The force of infection is

    ˜f(S,w)=J11+κ0S+J2,˜f(0,0)=0,

    ˜f is continuous at (0, 0) and bounded on bounded subsets on X+. By Corollary 1.2, the susceptible hosts persist uniformly weakly.

    These examples should not distract from the fact that the incidence function is allowed to depend on the infection-age distribution in a very general way which we do not remember to have seen in the literature. However an incidence function is derived, by first principles or from data, chances are that it is covered by this paper unless it is a generalization of frequency-dependent incidence. In spite of this generality, the outcome is the same as for the analogous model without infection-age: the disease does not drive the host into extinction, as long as the incidence function leads to a proper force of infection, like in Corollary 1.2 and differently from power incidences

    f(S,w)=SqJp,J=b0κ(a)w(a)da, with q(0,1) or p(0,1). (1.11)

    As a trade-off for the generality concerning the incidence function, our assumptions concerning the effect of the disease on the growth of the host population are somewhat restrictive: Only susceptible hosts reproduce and compete for resources effectively. These assumptions are sometimes made (the first more frequently than the second) in ecoepidemic models, which have a predator in addition to a parasite and to a host that is also a prey. See [41,56] [35, 10.4.2] and the literature cited therein; see [57] for other assumptions. To our knowledge, ecoepidemic models rarely mention concrete predator – prey/host – parasite systems. An exception is a highly infective Mycoplasma disease affecting the eyes of wild ungulates (chamois, ibex, mouflon) on high mountain pastures [41]. All age groups and both sexes are affected. The disease can either heal or lead to total blindness. In the case of wild ungulates, the disease leads to a change of behavior as the affected animals can no longer follow their herd. They become insecure and unable to feed and may fall from precipices. The disease should make it difficult to find a mate and, if reproduction occurs nevertheless, to keep offspring alive. Finally, it seems a valid assumption that diseased hosts have little or no impact on the food supply of susceptible hosts.

    The assumption that infected individuals do not effectively reproduce or compete seems acceptable for diseases that do not last for a long time because either recovery or disease death occurs fast or that cause severe symptoms interfering with feeding, mating, giving birth, or keeping neonates alive.

    This assumption does not only save us from mathematical difficulties but also from some modeling issues. While it is not too difficult to model the influence of infected individuals on population growth without infection age [51,57,58,59], the incorporation of infection-age raises many questions such as, if there is vertical transmission (from mother to unborn child), how is infection-age inherited?

    The epidemic model will be introduced in three equivalent versions in Section 2. In Section 3, we derive a priori properties of the model solutions. First, we establish upper bounds for nonnegative solutions; interestingly enough, they do not depend on the incidence function. Incidentally, the assumption that infected individuals do not compete for resources makes finding upper bounds more difficult. This was already challenging without infection-age [26] but is more so with infection-age.

    Secondly, we find conditions for the number of susceptible hosts to stay strictly positive. Clearly, this helps the persistence of susceptible hosts; those conditions involve the incidence function.

    Section 4 is the core of this paper, where we establish conditions for uniform weak host persistence. If this were a movie, we could have called it "The return of the Laplace transform".

    In Section 5 and 6, we present two special cases in which our assumptions for host persistence turn out to be necessary.

    Finally, in Section 7, we establish existence and uniqueness for nonnegative solutions of our model system. We do this at the end of the paper for two reasons. First, the proofs turn out to be the most technical of this paper, though as one expects, they ultimately rely on standard arguments: the contraction mapping theorem, an approximation procedure and Arzela-Ascoli type arguments. Therefore, we want to convince ourselves that we can gain epidemiological insights which give us enough motivation to tackle them. Second, the a priori considerations in Section 3 will be helpful. As the incidence function is allowed to depend on the infection-age distribution in a general way, we make a fixed point argument within a fixed point argument, which may be a novel feature.

    We will show existence but not uniqueness of solutions for power incidences (1.11) because they are not Lipschitz continuous at the origin. In spite of this mathematical deficiency, power incidence functions keep popping up in the literature (see [52,60,61,62] and the references therein). They fit the infection experiments in [25] very well [1]. With due caution, one has to say, though, that it is difficult to estimate incidence functions for very few susceptible and infective hosts, which is the range relevant for host survival.

    For introductions to mathematical epidemiology in book form, we refer to [22,30,34,35,40,51,52,63,64,65,66,67] and [68, Chap.6], [69, Chap.9, 10], [45, Part 3].

    We assume that, with respect to the infectious disease, the hosts are in one of two stages:

    They are susceptible to the disease or they are infected by the disease. They leave the susceptible stage by being infected and leave the infected stage either by dying from natural causes, dying from the disease or by becoming susceptible again.

    We do not consider the case where infected individuals recover from the disease and have acquired permanent immunity to it. Infected hosts are not necessarily infective: they may be still in a latent period or have recovered enough to be no longer infective and no longer subject to death from the disease. But we assume that they still have symptoms severe enough that they do not reproduce or are not able to keep their offspring alive and that prevent them from competing for resources.

    Rather than introducing classes of infected hosts that are in the latency period or have recovered from the disease (without having become susceptible again yet), in the spirit of [31,32] we introduce infection-age as the time since the moment of infection and assume that the infectivity of infected hosts depends on their infection-age. We also assume that the per capita rate at which infected hosts die from the disease or become susceptible again depends on their infection-age.

    Let P:R+[0,1] be the sojourn function of the infected stage (Section 1.3). Since P is decreasing, P is differentiable a.e. on R+ with nonpositive a.e. derivative P and P is finitely integrable [70, Thm.6.4]. Recall the maximal infection age b(0,] (Section 1.3). If b<, P(a)=0 for a.a. a(b,).

    Assumption 2.1. We assume throughout this paper that P is absolutely continuous on R+ and P(0)=1, i.e.,

    P(a)P(0)=P(a)1=a0P(s)ds,aR+. (2.1)

    Lemma 2.2. If b< and P:R+[0,1] is continuous, then P is absolutely continuous on R+ if and only if P is absolutely continuous on [0,b).

    Example 2.3 ([70, Thm.6.6]). Let b(0,]. P is absolutely continuous on R+ if P is continuous on R+ and Lipschitz continuous on every proper subinterval of [0,b).

    Assume that a0, h>0. Then P(a)P(a+h)P(a) is the probability of leaving the infected stage during the age interval [a,a+h). So, for almost all a0,

    γ(a):=P(a)/P(a), a.e. a[0,b), (2.2)

    is the rate at which an infected individual leaves the infected stage at infection age a. By a chain rule for absolute continuity (Lemma 8.1), lnP is absolutely continuous on [0,b) and (lnP)=γ a.e. on [0,b),

    P(a)=exp(a0γ(s)ds),a[0,b).

    Infected individuals leave the infected stage either by returning to the susceptible class, by death from the disease, or by disease-unrelated ("natural") death,

    γ(a)=α(a)+ϑ(a)+ν,a0. (2.3)

    Here α(a) is the per capita rate of becoming susceptible again, ϑ(a) the per capita disease death-rate and ν the natural death rate (which does not depend on infection-age).

    The function γ,α:[0,b)R+ are assumed to be Borel measurable, and γ is assumed to be bounded away from 0 a.e.

    If we do not like the explicit formula (2.6), we simply assume that there is some ν>0 such that

    γ(a)α(a)+ν,a[0,b), (2.4)

    where ν>0 is related to the natural mortality of hosts. As for susceptible hosts, the natural mortality has been incorporated into g.

    Let, at time t0, denote

    S(t) the amount of susceptible hosts,
    I(t) the amount of infected hosts,
    u(t,) the infection-age density of infected hosts,
    R(t) the rate of infected hosts returning into the susceptible class,
    B(t) the incidence, the rate of new infections.

    We choose the letter B for the incidence because it is the birth rate of the disease. See Table 1 for a more complete list of symbols. The disease dynamics can be mathematically described in various equivalent ways.

    Table 1.  Parameters and model variables.
    Symbol Definition
    S(t) number of susceptible hosts at time t
    I(t) number of infected hosts at time t
    R(t) rate of infected hosts returning to the S class at time t
    B(t) incidence, rate of new infections at time t
    J(t) infective impact at time t
    P(a) probability that an infectious host remains
    infected at infection-age a
    b maximal infection-age
    g(S) per capita growth rate of S susceptible hosts
    f incidence function
    u(t,) infection-age density of infected hosts at time t
    κ(a) infectivity of an infected host at infection-age a
    γ(a) rate at which an infected individual leaves the
    infected stage at infection-age a
    α(a) per capita rate at which an infected host ceases to be
    infected at infection-age a and returns to the S class
    ϑ(a) is the per capita rate at which infected hosts die from
    the disease at infection-age a
    ν per capita natural death rate
    K population's carrying capacity, g(K)=0

     | Show Table
    DownLoad: CSV

    This model formulation consists of an integro-differential equation and a partial differential equation,

    S=Sg(S)B+Ra.e. on (0,),(t+a)u(t,a)=γ(a)u(t,a),tR+,a[0,b),at,u(t,0)=B(t)=f(S(t),u(t,)),R(t)=b0α(a)u(t,a)da. (2.5)

    Further, there are initial conditions,

    S(0)=S0,u(0,a)=u0(a),a[0,b).

    In (2.5), S is assumed to be absolutely continuous. There, γ(a) is the rate at which infected hosts, at infection age a, cease to be infected by either natural death, disease death, or by becoming susceptible again; α(a) is the per capita rate at which infected hosts cease to be infected at infection age a and become susceptible again; more precisely

    γ(a)=α(a)+ϑ(a)+ν,a0, (2.6)

    where ϑ is the per capita rate at which infected hosts die from the disease and ν>0 is the per capita rate of natural death, which does not depend on infection age.

    For S0, g(S) is the per capita growth rate of susceptible hosts if their amount is S. The model formulation implicitly assumes that infected individuals do not reproduce or are not able to keep their offspring alive and that they do not compete for resources (cf. [10]). This assumption allows to reduce the problem to one for just two independent variables: the number of susceptibles, S, and the incidence of the disease, B.

    As described in Section 1.5, f:R+×X+R+, X+=L1+[0,b), is the incidence function, i.e., the entity f(S,w) is the rate of new infections if the amount of susceptible hosts is SR+ and the infection-age density of infected hosts is wX+.

    The per capita host growth rate g has the following properties.

    Assumption 2.4. The function g:R+R is Lipschitz continuous on bounded subsets of R+: For any c(0,), there exists some Λc(0,) such that

    |g(S)g(˜S)|Λc|S˜S|,S,˜S[0,c].

    Further g is decreasing, g(0)>0, and g(S)<0 for large S>0.

    In fact, we assume that there is a unique K>0 such that g(K)=0 and that g is continuously differentiable on (0,K+δ) with some δ>0, g(S)<0 for S(0,K] and Sg(S)0 as S0+.

    In particular, g is a decreasing function of the size of the susceptible part of the population due to resource competition.

    K is called the carrying capacity for the host population. The term "carrying capacity" originates from the fact that the host population size converges to it in the absence of the disease.

    If f(K,u)=0 when u is the zero function, (K,u) with u=0 is an equilibrium of (2.5), called the disease-free equilibrium, DFE, for short.

    This formulation takes the sojourn function P for the infection stage as a starting point. P(a) is the probability at which a host that has been infected is still alive and infected a time units after the moment of infection. See Section 1.3.

    Let u0:[0,b)R+ be the initial density of infected hosts, b0u0(a)da<. The model is a system consisting of a differential equation and Volterra functional equations,

    S=Sg(S)B+Ra.e. on (0,),B(t)=f(S(t),u(t,)),u(t,a)={B(ta)P(a),t>a0u0(at)P(a)P(at),0t<a}0a<b,R(t)=b0α(a)u(t,a)da. (2.7)

    Further, there is the initial condition for the susceptible hosts, S(0)=S0.

    Since there are a lot of variables in this system, let us clarify that S0, u0, P, α, g, f are given and a continuous function B and an absolutely continuous function S are wanted. The auxiliary variable u is determined by B and u0 in the spirit of Lotka [72]. See (1.7).

    The assumption (2.4) takes the form

    γ(a):=P(a)P(a)α(a)+νfor a.a. a[0,b). (2.8)

    The Lotka formulation can be derived from the McKendrick formulation by integration along characteristics [45, Sec.13.2]. The function u satisfies the PDE in the McKendrick formulation if B, u0 and P are differentiable and γ is given by (2.8) [45, Sec.13.2]. Otherwise, the PDE is satisfied in an appropriate generalized sense.

    As a special case of (2.7), we consider

    f(S,w)=˘f(S,J1,,Jn),wL1+[0,b),˘f:Rn+1+R+,Jj=b0κj(a)w(a)da,κjL+[0,b). (2.9)

    The Jj are various impacts by infectious hosts associated with infection-age dependent per capita infectivities κjL+[0,b). If κ1=1, then J1=I is the amount of infective individuals.

    Several infectious impacts may occur if the infection can take different routes like the eye disease of chamois, ibex and mouflon [41] that we mentioned towards the end of Section 1.5: It spreads not only through direct contact but also by airborne particles and through vectors such as flies. Also see Example 1.3.

    If b(0,), we extend κj and α to R+ by κj(a)=0=α(a), a(b,), b(0,]. We substitute the formula for u in (2.7) into Jj(t)=b0κj(a)u(t,a)da and obtain a system of one differential equation and several Volterra integral equations,

    S=Sg(S)B+Ra.e. on (0,),B(t)=˘f(S(t),J1(t),,Jn(t)),Jj(t)=t0κj(a)P(a)B(ta)da+Jj0(t),R(t)=t0α(a)P(a)B(ta)da+R0(t), (2.10)

    with the given functions

    Jj0(t)=b0κj(s+t)P(s+t)P(s)u0(s)ds,R0(t)=b0α(s+t)P(s+t)P(s)u0(s)ds. (2.11)

    If 0<t<b<, since P(a)=0 for a>b,

    R0(t)=bt0α(s+t)P(s+t)P(s)u0(s)ds,

    and similar formulas hold for Jj0. If bt<, R0(t)=0=Jj0(t).

    Under our assumptions so far, it cannot be excluded that R0 takes infinite values; however R0 is finitely integrable on R+. Indeed, by Tonelli's theorem and (2.8),

    0R0(t)dt0(b0γ(s+t)P(s+t)P(s)u0(s)ds)dt=b0(0P(s+t)dt)u0(s)P(s)ds=b0u0(s)ds. (2.12)

    We give this special case extra consideration because we could not prove existence of solutions to (2.7) without a Lipschitz conditions for f:R+×X+R+ (Section 7.1) whereas (2.10) can be solved for ˘f:Rn+1+R+ that are just continuous (Section 7.2). See Example 2.8.

    We now have developed enough terminology to give a preview of some more results. The following one generalizes Corollary 1.2 and follows from Theorem 4.1. Recall that X+=L1+[0,b).

    Corollary 2.5. Assume that there exists a function ˜f:R+×X+R+ such that f(S,w)=S˜f(S,w) for all S0, wX+, where ˜f is continuous at (0,0) and is bounded on bounded sets. Then, if

    g(0)>(1b0α(a)P(a)da)˜f(0,0), (2.13)

    the susceptible hosts persist uniformly weakly. In particular, if ˜f(0,0)=0, the susceptible hosts persist uniformly weakly.

    The function ˜f is often called the force of infection [51] [52, Sec.1] [54] [35, 6.2.2], not to be confused with infective impact which is a term specific for infection-age structured models; for density-dependent incidence, the two can coincide [34, Sec.5.3]. The term ˜f(S,0) represents the force of infection coming from reservoirs of the disease.

    Remark 2.6. By (2.8),

    :=b0α(a)P(a)dab0νP(a)dab0P(a)da.

    By the fundamental theorem of calculus,

    1νD<1, (2.14)

    where D is the expected sojourn time in the infection period (natural death included). Recall that 1/ν>D is the life expectation of a healthy host. We owe to one of the referees the insight that can be interpreted as the probability that a susceptible host just infected will turn susceptible again and how to reformulate (2.13) in terms of a basic host reproduction number, H0. The number 1(0,1]) is the probability that a susceptible host just infected will not turn susceptible again. Let g(0)=βμ with μ the per capita death rate of healthy hosts and β the per capita birth rate without competition for resources. Then (2.13) can be rewritten as

    1<H0=β˜f(0,0)(1)+μ. (2.15)

    Notice that (˜f(0,0)(1)+μ)1 is the average time a host is alive and susceptible if very few hosts (susceptible or infective) are around.

    Theorem 2.7. Assume that f(S,w)=˘f(S,J) with J=b0κ(a)w(a)da, where κL[0,b) and ˘f:R2+R is differentiable at (0,0), ˘f(0,J)=0 for all J0 and 1˘f(0,0)=0.

    Further assume that there is some η0>0 such that

    α(a)η0κ(a),a[0,b). (2.16)

    Then the susceptible hosts, S, persist uniformly weakly.

    This theorem is proved at the end of Section 4.2.

    Condition 2.16 implies that, at whatever infection-age, if some infected individuals are infectious there are also some infected individuals that leave the infected class and become susceptible again.

    Example 2.8. Let f(S,w)=SqJp=˘f(S,J) with q(0,1) and p>0 and q+p>1, J=b0κ(a)w(a)da, SR+, wX+; further κL+[0,b).

    Assume that there is some η0>0 such that

    α(a)η0κ(a),for a.a. a[0,b).

    Since

    ˘f(S,J)(S+J)p+q,S,J0,

    with p+q>1, ˘f is differentiable at (0, 0) with zero derivative, and the susceptible hosts persist uniformly weakly under this incidence function and assumption (2.16).

    We mention that the model in Section 2 has non-negative solutions under these assumption (Section 7.2) though they may not be unique.

    This result should be compared with the next one in which host persistence fails if assumption (2.16) is violated in the extreme form that α(a)=0 for all a0 and κ is a positive constant.

    Theorem 2.9. Let α,γ,κ be nonnegative constants and α=0<γ and f(S,w)SqJp, S0, J=κb0w(a)da with q(0,1) and p0. Assume that S(0)>0 and I(0)>0 satisfy

    (S(0))1q(I(0))p<(1q)κp(1q)g(0)+pγ.

    Then S(t) becomes 0 at finite time and stays zero thereafter, and I(t)0 as t.

    For the proof see Section 6. Theorem 2.9 shows that the assumption 2.16 in Theorem 2.7 cannot be completely abandoned.

    One may like to turn to numerical calculations to explore the grey zone between Theorem 2.7 and Theorem 2.9.

    We fix a natural per capita host mortality rate ν>0, i.e., 1/ν is the life expectation of a healthy host. Then we choose a sojourn function Q for the infection period (natural death disregarded) such that P(a)=eνaQ(a). To find the per capita rates of disease death on the one hand and of return into the susceptible class on the other hand, we use the concept of a case fatality ratio (usually but falsely called a rate) [16], which we assume to be independent of infection age and is denoted by ϖ(0,1): If an individual leaves the infection period, it leaves by disease death with probability 1ϖ and leaves by return into the susceptible class with probability ϖ. This means that

    α(a)=ϖQ(a)Q(a),0a<b. (2.17)

    If b=, boundedness of α on R+ may not be a bad assumption. Let us take a look at the sojourn functions which have been fitted in [2] to the infection experiments involving salamander larvae and ranavirus [25]. We should mention that all infected larvae (their number was 446) died from the disease and that they were all dead by the 23rd day. Since they were larvae and not yet terrestrial salamanders, reproduction was not an issue.

    If the duration of the infection stage (neglecting natural death) is lognormally or Gamma distributed Q/Q is bounded on R+ for the associated sojourn function Q.

    If the length of the infection period (neglecting natural death) is lognormally distributed, the probability density of the sojourn function is given by

    (2π)1/2Q(a)=eb(a)2/21ζa,b(a)=(1/ζ)ln(a/m),a0,

    with parameters ζ,m>0 [45, Sec.12.8]. By the chain rule,

    (2π)1/2Q(a)=eb(a)2/2(b(a)1(ζa)2+1ζa2)=eb(a)2/2ζ3a2(ln(a/m)+ζ2).

    Notice that Q(a), Q(a) and Q(a) are different from 0 if a/m>eζ2. Further, Q(a)0, Q(a)0 and Q(a)0 as a. Since Q is decreasing, actually Q(a)<0 for all a>0. So

    limaQ(a)Q(a)=limaQ(a)Q(a)=0

    by a l'Hôpital type argument [2, Thm.7.2]. This implies that QQ is bounded on R+.

    Similar arguments show that this also holds if the infection period is Gamma-distributed.

    However, for the Weibull distribution, Q(a)=eξaζ,a0, with ξ>0 and ζ1, we have

    Q(a)Q(a)=ξζaζ1,a>0,

    which is unbounded on (0,), unless ζ=1, the special case of an exponential distribution.

    If b(0,), after normalization b=1, then Q(a)=(1am)n with n,m>1 would have a plausible shape. (See the figures in [2]). Then

    Q(a)Q(a)=n(1am)1mam1,a[0,1),

    goes to infinity if a1.

    In view of these findings, we will try to avoid assuming essential boundedness of α as much as we can. Instead, we will try to exploit (2.8) which gives us for 0a1<a2 that

    a2a1α(a)P(a)daa2a1γ(a)P(a)daνa2a1P(a)da=a2a1P(a)daνa2a1P(a)da=P(a1)P(a2)νa2a1P(a)da.

    In the limit a10 and a2b,

    0α(a)P(a)da1νb0P(a)da. (2.18)

    Assuming that α is not essentially bounded, has the unpleasant technical complication that the return rate R(t) is not continuous in t and that the differential equation for S in (2.7) holds only a.e..

    Before we discuss the existence of solutions of either of the systems, we derive some of their properties provided that they exist. We will need these properties to discuss the dynamics of the system, and later this information will help us in setting up the existence proof.

    Since u(t,) is the infection-age density of infected hosts, the total number of infected hosts is given by

    I(t)=b0u(t,a)da,t0. (3.1)

    The total number of hosts, H, is

    H(t)=S(t)+I(t),t0. (3.2)

    By (2.7),

    I(t)=t0P(ts)B(s)ds+b0P(t+s)P(s)u0(s)ds. (3.3)

    In particular,

    I(0)=b0u0(a)da. (3.4)

    Assume that the number of susceptibles S is absolutely continuous and that the disease incidence B is continuous. Further, we assume that S, B, u0 are nonnegative. Then I in (3.3) and R in (2.10) are nonnegative.

    Since P is absolutely continuous, I is absolutely continuous and, for a.a. t0, by (3.3),

    I(t)=B(t)+t0P(ts)B(s)ds+b0P(t+s)P(s)u0(s)ds. (3.5)

    As for the first two expressions after the equality sign, one checks that integrating them in time yields the first expression in (3.3) by Tonelli's theorem. Indeed, since the integrand is non-positive a.e.,

    T0(t0P(ts)B(s)ds)dt=T0(TsP(ts)dt)B(s)ds=T0(P(Ts)P(0))B(s)ds=T0P(Ts)B(s)dsT0B(s)ds.

    As for the third expression, one checks that integrating it in time yields the second expression in (3.3) after the equality sign by Tonelli's theorem. Indeed, for any T>0, since the integrand is non-negative, by Tonelli's theorem,

    T0(b0P(t+s)P(s)u0(s)ds)dt=b0u0(s)P(s)(T0P(t+s)dt)ds=b0u0(s)P(s)(P(s)P(T+s))ds=b0u0(s)dsb0P(T+s)P(s)u0(s)ds.

    By (1.7) and (3.5), for a.a. t>0,

    I(t)=B(t)t0γ(a)P(a)B(ta)dsb0γ(t+s)P(t+s)P(s)u0(s)ds=B(t)b0γ(a)u(t,a)da.

    Since we have assumed that S is absolutely continuous, H is absolutely continuous and by (2.7), for a.a. t>0,

    H(t)=S(t)g(S(t))+b0[α(a)γ(a)]u(t,a)da.

    By (2.8),

    H(t)S(t)g(S(t))νI(t)a.e.

    By (3.2), this can be rewritten as

    HS(g(S)+ν)νHa.e.

    Let

    ψ(t)=S(t)[g(S(t))+ν],t0. (3.6)

    Then, for a.a. t>0,

    H(t)+νH(t)ψ(t).

    This implies

    t0eν(ts)ψ(s)dst0eν(ts)H(s)ds+t0νeν(ts)H(s)ds.

    By the integration by parts formula for absolutely continuous functions [73, Cor.18.20], for all t0,

    t0eν(ts)ψ(s)dsH(t)H(0)eνt. (3.7)

    Let T>0 and ψT=supψ([0,T]) and HT=supH([0,T]). For t[0,T],

    H(t)ψT1ν(1eνt)+H(0)eνtmax{1νψT,H(0)}.

    Taking the maximum for t[0,T],

    HTmax{1νψT,H(0)}.

    By (3.6),

    HTmax{H(0),1νmax0SSTS(g(S)+ν)}.

    Suppose that HT>H(0). Then there exists some ˜S[0,ST] such that

    HT1ν˜S(g(˜S)+ν).

    Suppose that ˜S>K. Then g(˜S)<0 and HT<˜SST, a contradiction. So ˜SK and

    HTmax{H(0),1νmax0SKS(g(S)+ν)}. (3.8)

    Since the right hand side of this inequality does not depend on T>0, H is bounded on R+ and

    H(t)max{H(0),1νmax0SKS(g(S)+ν)},t0,H(0)=S0+0u0(a)da. (3.9)

    Once we know that H is bounded on R+, we also know that SH is bounded and ψ in (3.6) is bounded.

    In the following, we use the following notation:

    H=lim suptH(t)=infs0suptsH(t). (3.10)

    S and ψ and J are defined analogously. By (3.7),

    Hlim suptt0eνsψ(ts)ds.

    This can be written as

    Hlim supt0eνsH0(ts)ψ(ts)ds

    with the Heaviside function H0(s)=1 if s0 and H0(s)=0 if s<0. By a generalization of the Lebesgue-Fatou lemma [45, p.468],

    H0lim suptH0(ts)ψ(ts)eνsdsψ0eνsds=ψν. (3.11)

    By (3.6),

    ψ=lim suptS(t)[g(S(t))+ν]=limnS(tn)[g(S(tn))+ν]

    for an appropriate sequence (tn) with tn as n. Since S is bounded, after choosing a subsequence, S(tn)˜S as n for some ˜S[0,S]. So, ψ=˜S[g(˜S)+ν]. By (3.11),

    H1ν˜S[g(˜S)+ν] (3.12)

    with some ˜S[0,S]. If ˜S>K, then g(˜S)<0 and H<˜SS, a contradiction because H(t)S(t) for all t0. Hence, ˜SK and

    H1νmax0SKS[g(S)+ν]=:H. (3.13)

    We also get from (3.12) that

    HSν[g(0)+ν]. (3.14)

    Assume κL+[0,b) and

    κ=κ. (3.15)

    Then the infective impact J associated with κ via (1.5) is bounded and

    J(t)κI(t),t0. (3.16)

    By (3.13),

    Jκνmax0SKS[g(S)+ν]. (3.17)

    By (3.14),

    JSκν[g(0)+ν]. (3.18)

    Uniform weak persistence of susceptible hosts is certainly facilitated if all nonnegative solutions S and B of (2.7) on R+ with S(0)>0 satisfy S(t)>0 for all tR+. In this section we will derive conditions for this to hold.

    Lemma 3.1. Assume that for any c>0 there is some ˜c>0 such that

    f(S,w)˜cS,0S,w1c,wX+. (3.19)

    Then the following hold.

    (a) If T(0,) and there is a continuous nonnegative solution of (2.7) on [0,T] with S(t)>0 for all t[0,T), then S(T)>0.

    (b) If there is a continuous solution of (2.7) on [0,) and S(0)>0, then S(t)>0 for all t0.

    Proof. (a) Let S(0)>0 and S(t)>0 for t[0,T).

    Suppose that S(T)=0. By (3.8), there is some c>0 such that S(t)c,I(t)=b0u(t,a)dac for all t[0,T]. By assumption (3.19), there is some ˜c>0 such that the function ψ(t)=f(S(t),u(t,))/S(t), t[0,T), satisfies ψ(t)˜c for all t[0,T). Further, by (2.7),

    S(t)S(t)[g(S(t))˜c], for a.a. t[0,T].

    We use an integrating factor,

    S(t)S(0)exp(t0[g(S(s))˜c]ds)>0=S(T),0t<T,

    contradicting the continuity of S at T.

    Part (b) follows from part (a).

    The next positivity result involves the per capita return rate α in (2.7) of infected hosts into the susceptible class. Return into the susceptible class keeps the number of susceptible hosts away from zero provided there is some relation between the infectivity and the per capita return rate of infected hosts. See Theorem 2.7 and Example 2.8. Recall that we do not assume that αL+[0,b) (Section 2.6).

    Lemma 3.2. Let κL+[0,b) and η0(0,1) such that

    α(a)η0κ(a), for a.a. a[0,b). (3.20)

    Assume that there are c,δ>0 such that

    SR+,wX+,J=b0κ(a)w(a)da,S+Jδf(S,w)cS+η0J. (3.21)

    Then, if T(0,) and S,B is a continuous nonnegative solution of (2.7) on [0,T] and S(t)>0 for all t[0,T), then S(T)>0.

    Proof. Let T(0,) and S(t)>0 for all t[0,T). Suppose S(T)=0.

    By (3.20),

    R(t)=b0α(a)u(t,a)dab0η0κ(a)u(t,a)da=η0J(t),t[0,T]. (3.22)

    We integrate the differential equation for S in (2.7); for all t(0,T),

    0S(T)S(t)TtS(s)g(S(s))dsTtB(s)ds+η0TtJ(s)ds.

    We divide by Tt>0 and take the limit tT. Since S, J and B are continuous and J(T)0=S(T),

    0S(T)g(S(T))B(T)+η0J(T).

    Since B(T)=f(S(T),u(T,))=0, we have J(T)=0.

    Since S and J are continuous, there is some s(0,T) such that S(t)+J(t)<δ for all t[s,T] with δ>0 from (3.21).

    By (3.21) and (3.22),

    f(S(t),u(t,))cS(t)+η0J(t)cS(t)+R(t),t[s,T].

    By (2.7),

    S(t)cS(t) for almost all t[s,T].

    Hence S(T)S(s)ec(sT)>0, a contradiction.

    Recall the definition of uniform weak persistence of susceptible hosts (Definition 1.6). The following persistence considerations are very much facilitated by using the Laplace transform.

    Let S:R+R+ be Borel measurable and be bounded or finitely integrable. Then

    ˆS(λ)=0eλtS(t)dt,λ>0. (4.1)

    The Laplace transform is useful in this context because it converts convolutions into products.

    Let B:R+R+ be Borel measurable and be bounded or finitely integrable and κL1+(R+).

    Then the convolution of B and κ is defined by

    (Bκ)(t)=t0B(ts)κ(s)ds,tR+. (4.2)

    Respectively, Bκ is bounded or finitely integrable on R+ and

    ^Bκ(λ)=ˆB(λ)ˆκ(λ),λ>0. (4.3)

    See [74, Sec.2.2].

    Theorem 4.1. Let Assumption 2.4 be satisfied. Assume that for any c>0 there is some ˜c>0 such that f(S,w)˜cS for all S[0,c] and all wX+ with w1=b0w(a)dac. Further assume that

    (1b0α(a)P(a)da)lim supS,w0f(S,w)S<g(0). (4.4)

    Then the susceptible hosts persist uniformly weakly.

    Condition (4.4) is equivalent to the existence of some δ(0,g(0)) such that

    (1b0α(a)P(a)da)f(S,w)Sg(0)δ,0<Sδ,wX+,w1δ. (4.5)

    Remark 4.2. Under the assumptions of Theorem 4.1, for any S0(0,) and any u0X+, there exists a unique solution to (2.7) provided that f:R+×X+R+ is Lipschitz continuous on bounded subsets of R+×X+:

    For any c(0,), there exists some Λc(0,) (called a Lipschitz constant) such that

    |f(S,w)f(˜S,˜w)|Λc(|S˜S|+|w˜w|),S,˜S[0,c],w,˜wX+,w˜w1c. (4.6)

    See Section 7.1. f is called Lipschitz continuous if there is some Λ=Λc that does not depend on c>0.

    Proof of Theorem 4.1. Suppose that the susceptible hosts do not persist uniformly weakly. Then, for any ϵ>0, there exists a solution with S(0)>0 and S<ϵ. If b(0,), extend α to R+ by α(a)=0 for a[b,). By assumption (4.5), there exists some δ>0 (by making it smaller if needed) such that

    (10α(a)P(a)da)f(S,w)S<g(δ)δ,S(0,δ],wX+,w1δ. (4.7)

    Since the susceptible hosts do not persist uniformly weakly, there exists a solution with S(0)>0 and S<ϵ. By (3.14), if ϵ(0,δ) is chosen small enough, S<δ, I<δ,I(t)=b0u(t,a)da. By Lemma 3.1, S(t)>0 for all t0. So there exists some r>0 such that 0<S(t)<δ and I(t)<δ for all tr. We write some of the model equations in the form

    S=Sg(S)B+R,R(t)t0α(a)P(a)B(ta)da. (4.8)

    We introduce the translations Sr(t)=S(r+t), Br(t)=B(r+t), Ir(t)=I(r+t) for r,t0. Further x(t)X+ is defined by x(t)(a)=u(t,a) for t0, a[0,b) and xr(t)=x(r+t) for t,r0. Notice that Ir(t)=xr(t)1. By (4.8), since B is nonnegative,

    δSr(t),xr(t)1,t0,Sr=Srg(Sr)Br+Rr,Rr(t)t0α(a)P(a)Br(ta)da. (4.9)

    We take Laplace transforms which convert convolutions into products, (4.3),

    ˆSr(λ)=0eλtSr(t)dt,λ>0.

    For all λ>0,

    λˆSr(λ)^Srg(Sr)(λ)ˆBr(λ)+ˆRr(λ),ˆRr(λ)^αP(λ)ˆBr(λ).

    We substitute the second inequality into the first,

    λˆSr(λ)^Srg(Sr)(λ)(1^αP(λ))ˆBr(λ). (4.10)

    Since Sr(t),xr(t)1δ and Sr(t)>0 for all t0,

    ˆBr(λ)=0eλtSr(t)f(Sr(t),xr(t))Sr(t)dtˆSr(λ)supt0f(Sr(t),xr(t))Sr(t)ˆSr(λ)sup0<Sδ,w1δf(S,w)S.

    Similarly, since g is decreasing,

    ^Srg(Sr)(λ)ˆSr(λ)g(δ),λ>0.

    We substitute these inequalities into (4.10) and divide by 0<ˆSr(λ)<,

    λg(δ)(1^αP(λ))sup0<Sδ,w1δf(S,w)S,λ>0.

    This inequality does not depend on the solution of our system and holds for all λ>0. We take the limit of this inequality as λ0,

    0g(δ)(10α(a)P(a)da)sup0<Sδ,0w1δf(S,w)Sδ>0,

    with the last, contradictory, inequality following from (4.7).

    Corollary 2.5 is a consequence of Theorem 4.1.

    Remark 4.3. If f(S,w)=SqJp, J=b0κ(a)w(a)da, 0<q<1, p>0, κL[0,b), κ>0, then

    lim supS,w0f(S,w)S=.

    So this power incidence function does not satisfy (4.4) and Theorem 4.1 does not apply. Compare Example 2.8 which will follow from our next next persistence result (Theorem 4.4).

    Proof. Choose a sequence (wn) in X+=L1+[0,b) with

    0<Jn0,n,Jn=b0κ(a)wn(a)da.

    This is possible because κ>0. Choose Sn=Jrn with some r>0 to be chosen later. Then

    f(Sn,Jn)Sn=Sq1nJn=Jr(q1)+1n.

    If r>0 is chosen large enough, r(q1)+1<0 and the right hand side of the last inequality tends to infinity as n.

    Remark 4.3 motivates the next general persistence result.

    Theorem 4.4. Assume that any continuous nonnegative solutions S,B of (2.7 with S(0)>0 satisfies S(tn)>0 for a sequence (tn) in R+ with tn as n.

    Assume there is some ζ(0,) with the following properties:

    (i) For any η>0 there exists some ˜ϵ>0 such that

    SR+,wX+,S+w1˜ϵf(S,w)(ζ+η)S+ηR,R=b0α(a)w(a)da; (4.11)

    (ii)

    ζ(1b0α(a)P(a)da)<g(0). (4.12)

    Then the susceptible hosts persist uniformly weakly.

    Notice that, by γ=P/P and P(0)=1,

    1b0α(a)P(a)da=b0[γ(a)α(a)]P(a)da>0. (4.13)

    Proof. If b(0,), we extend α:[0,b)R+ to R+ by α(a)=0 for a[b,). By (4.12), there is some δ(0,1)(0,g(0)), such that

    (ζ+δ)(1(1δ)0α(a)P(a)da)<g(δ)δ. (4.14)

    We write some of the model equations in the form

    S=Sg(S)B+δR+(1δ)R a.e.,R(t)=b0α(a)u(t,a)dat0α(a)P(a)B(ta)da. (4.15)

    Suppose that the susceptibles do not persist uniformly weakly.

    Then, for arbitrary ϵ>0, which can be chosen as small as needed to generate a contradiction, there exists a solution such that S(0)>0 and S<ϵ. By assumption, S(tn)>0 for a sequence (tn) with tn as n. By (3.14),

    (S+I)<ϵν[g(0)+ν]. (4.16)

    We can assume ϵ<δ. See (4.14). There exists r>0 (r=tn for some sufficiently large n) such that S(r)>0 and

    0<S(t)<ϵ,I(t)=u(t,)1<ϵν[g(0)+ν],tr, (4.17)

    We introduce that translations Sr(t)=S(r+t), Rr(t)=R(r+t), Br(t)=B(r+t) for r,t0. By (4.15),

    SrSrg(δ)Br+δRr+(1δ)Rr a.e.,Rr(t)t0α(a)P(a)Br(ta)da. (4.18)

    We take Laplace transforms which convert convolutions into products. For all λ>0, we obtain

    ˆRr(λ)ˆBr(λ)^αP(λ) (4.19)

    and

    λˆSr(λ)ˆSr(λ)g(δ)ˆBr(λ)(1(1δ)^αP)+δˆRr(λ). (4.20)

    By (4.11) there exists some ˜ϵ>0 such that

    f(S,w)(ζ+η)S+ηR,S,w1[0,˜ϵ]. (4.21)

    Let ϵ(0,˜ϵ) and ϵν[g(0)+ν]<˜ϵ. By (4.17), for all t0, 0<Sr(t)<˜ϵ and Ir(t)<˜ϵ and, by (4.21),

    Br(t)(ζ+η)Sr(t)+ηRr(t).

    Since the Laplace transform is a linear operation,

    ˆBr(λ)(ζ+η)ˆSr(λ)+ηˆRr(λ).

    We substitute this inequality into (4.20),

    λˆSr(λ)ˆSr(λ)g(ϵ)((ζ+η)ˆSrηˆRr(λ))(1(1δ)^αP)+ηˆRr(λ).

    This implies

    λˆSr(λ)ˆSr(λ)g(ϵ)((ζ+η)ˆSr(λ))(1(1δ)^αP). (4.22)

    Since ˆSr(λ)>0 because Sr(0)>0, we can divide by it,

    λg(ϵ)((ζ+η))(1(1δ)^αP). (4.23)

    Now, this inequality holds for any λ>0 and so it holds in the limit λ=0,

    0g(ϵ)(ζ+η)(1(1δ)0α(a)P(a)da). (4.24)

    Since ϵδ and ηδ, this inequality contradicts inequality (4.14).

    We combine Lemma 3.2 and Theorem 4.4.

    Corollary 4.5. Let κL+[0,b) and η0(0,1) such that

    α(a)η0κ(a),fora.a.a[0,b). (4.25)

    Assume there is some ζ(0,) with the following properties:

    (a) For any ˘η>0 there exists some ˘ϵ>0 such that

    SR+,wX+,S+J˘ϵ,J=b0κ(a)w(a)daf(S,w)(ζ+˘η)S+˘ηJ;

    (b)

    ζ(1b0α(a)P(a)da)<g(0).

    Then the susceptible hosts persist uniformly weakly.

    Proof. Observe that for all wX+, we have Jκw1 and, by (4.25), b0α(a)w(a)da=Rη0J.

    Some tedious but elementary considerations then show that the assumptions (a) and (b) imply the assumptions of Lemma 3.2 and Theorem 4.4.

    Lemma 3.2 implies that S(0)>0 implies S(t)>0 for all t>0 for all solutions of (2.7) on R+. Apply Theorem 4.4.

    We assume that

    f(S,w)=˘f(S,J),SR+,wX+,J=b0κ(a)w(a)da, (4.26)

    where κL+[0,b),

    ˘f(0,J)=0,JR+. (4.27)

    Then the partial derivative 0=2f(0,0) with respect to J exists.

    Corollary 4.6. Assume that the partial derivative 1˘f with respect to S exists on R2+ and is bounded on bounded subsets of R+ and is continuous at (0,0) with

    (1b0α(a)P(a)da)1˘f(0,0)<g(0). (4.28)

    Then the susceptible hosts persists uniformly weakly.

    Proof. By the mean value theorem, for S>0 and J0,

    ˘f(S,J)S=1˘f(˜S,J) (4.29)

    with some ˜S(0,S). Let c>0. Since 1˘f is bounded on bounded sets, there exists some ˜c>0 such that 1˘f(˜S,J)˜c for all ˜S,J[0,c]. By (4.29), ˘f(S,J)˜cS for all S,J[0,c]. Since 1˘f is assumed to be continuous at (0,0),

    ˘f(S,J)S1˘f(0,0),S,J0,

    and (4.28) implies (4.4). The assertion now follows from Theorem 4.1.

    Proof of Theorem 2.7. The assumptions of Corollary 4.5 are satisfied with ζ=1˘f(0,0)=0.

    The constant risk incidence has the form

    f(S,w)=σS,S,wX+. (5.1)

    This means that the per capita risk of a susceptible host to be infected does not depend on the number of infected hosts. It is the only incidence that is differentiable at (0,0) and is positively homogeneous of degree 1. The constant risk incidence is a simplistic way to model an infection reservoir. Surprisingly, it did not fare so bad in the fits in [1] of the infection experiments in [25] as one might expect. As unrealistic as constant risk may appear, we consider this incidence in order to put our other results into perspective and to present an example in which our persistence results are sharp. Notice that the constant risk incidence is infinitely often differentiable and has the partial derivatives

    1f(S,w)=σ,2f(S,w)=0. (5.2)

    As a special case of Corollary 2.5 or 4.6 we obtain the following result.

    Theorem 5.1. Let g(0)>σ(10α(a)P(a)da). Then the susceptible hosts persist uniformly weakly.

    In the following we show that this result is sharp. The system (2.10) decouples

    S=Sg(S)σS+R(t),R(t)=t0α(a)σS(ta)P(a)da+R0(t),R0(t)=b0α(s+t)u0(s)P(s+t)P(s)ds. (5.3)

    The equation for the infected hosts takes the form

    I(t)=t0σS(ta)P(a)da+I0(t),I0(t)=0u0(s)P(s+t)P(s)ds. (5.4)

    Since the solutions are bounded by (3.9), we can apply the fluctuation method [75] [45, Prop.A.22]: There exists a sequence (tn) with tn and S(tn)S and S(tn)0 as n. This yields

    0=Sg(S)σS+limntn0α(a)σS(tna)P(a)da+lim supnR0(tn).

    By the Lebesgue-Fatou's lemma [45, p.468],

    0Sg(S)σS+0α(a)σlim supnS(tna)P(a)da+R0.

    Now lim supnS(tna)S for any a0. If b<, R0(t)=0 for tb. If α is bounded, R0(tn)0 as n by the almost everywhere version of the dominated convergence theorem [70, Thm.4.9]. In either case,

    0S(g(S)σ+σ0α(a)P(a)da). (5.5)

    Recall (2.14),

    =0α(a)P(a)da1νD,

    where D is the expected sojourn time in the infection period (natural death included). Remember that 1/ν>D is the life expectation of a healthy host.

    Theorem 5.2. Let g(0)σ(10α(a)P(a)da) and let b< or α be bounded. Then the hosts die out.

    Proof. Suppose the susceptible hosts do not die out. Then S>0 and g(0)>g(S) and g(S)σ(10α(a)P(a)da) by (5.5), a contradiction to the assumption of this theorem. So S=0. Another application of Fatou's lemma, this time to (5.4), shows that I=0.

    If the rates γ, α and κ do not depend on infection-age, we have R(t)=αI(t) and J(t)=κI(t) with I given by (3.1), and P(a)=eγa. The system (2.10) takes the form

    S=Sg(S)˘f(S,κI)+αI,I(t)=t0eγ(ta)˘f(S(a),κI(a))da+I0(t),I0(t)=eγt0u0(s)ds. (6.1)

    S and I satisfy the ODE system

    S=Sg(S)ˇf(S,κI)+αI,I=ˇf(S,κI)γI, (6.2)

    with I(0)=0u0(a)da and S(0)=S0. Theorem 2.9 now follows from [1, Thm.5.8]. In turn, any solution S, I of (6.2) can be written in the form of (6.1) with

    I0(t)=I(0)eγt=eγt0u0(s)ds,u0(s)=γI(0)eγs. (6.3)

    If α>0, by Example 2.8, we obtain that the susceptible hosts persist uniformly weakly in model (6.2). Actually, the persistence is uniform by [26, Exp.2.13].

    We establish the existence of solutions to the model formulations in Section 2.3 and Section 2.4. The formulation in 2.3 is equivalent to the one in 2.2. Existence of solutions to the model formulations in Section 2.3 is proved under the assumption that the incidence function f:R+×L1+[0,b)R+ is Lipschitz continuous on bounded subsets of its domain. The proof involves a fixed point argument within a fixed point argument.

    Existence of solutions to the model formulations in Section 2.2 is proved in the case that f does not necessarily satisfy a Lipschitz condition but is of the special form

    f(S,w)=˘f(S,J),J=b0κ(a)w(a)da,κL+[0,b),˘f:R2+R+ continuous. (7.1)

    The proof for more infectious impacts than just J is similar.

    We rewrite (2.7) in integrated form,

    S(t)=[S0+t0S(s)g(S(s))dst0B(s)ds+t0R(s)ds]+,R(t)=t0α(a)B(ta)P(a)da+b0α(r+t)u0(r)P(r+t)P(r)dr. (7.2)

    Here, []+ is the positive part of a real number rR, [r]+=max{r,0}, and is used because we want S(t)0 whereas the right hand side may be negative. Further, we use the notation x(t)(a)=u(t,a), t,aR+, x(t)X+=L1+[0,b), for a more rigorous mathematical treatment,

    B(t)=f(S(t),x(t)),x(t)L1+[0,b),x(t)(a)={B(ta)P(a),0a<tu0(at)P(a)P(at),0t<a}a[0,b). (7.3)

    Notice that x(0)=u0 in L1+[0,b)=X+. Our approach consists in considering this system as a fixed-point problem for S and B and apply the contraction mapping theorem (Banach's fixed point theorem). The subsystem (7.3) seems to make it difficult to apply the existence results for functional equations in [74, Sec.12.2]. We will deal with subsystem (7.3) as a fixed point problem within a fixed point problem (Section 7.1.1).

    Then we will show that solutions to (7.2) and (7.3) are solutions to (2.7).

    Our first step consists in eliminating R for the equation for S in (7.2). We change the order of integration,

    t0R(s)ds=t0(s0α(sr)P(sr)B(r)dr)ds+t0(b0α(r+s)u0(r)P(r+s)P(r)dr)ds=t0(trα(sr)P(sr)ds)B(r)dr+b0(u0(r)/P(r))(t0α(r+s)P(r+s)ds)dr.

    After a substitution,

    t0R(s)ds=t0A(tr)B(r)dr+v0(t) (7.4)

    with

    A(t)=t0α(s)P(s)ds,v0(t)=b0u0(r)˜A(t,r)dr,˜A(t,r)=(1/P(r))t+rrα(s)P(s)ds,0r,t.

    Since α(s)γ(s)=P(s)/P(s),

    A(t)1P(t),˜A(t,r)1. (7.5)

    We substitute (7.4) into (7.2),

    S(t)=[S0+t0S(s)g(S(s))dst0(1A(ts))B(s)ds+v0(t)]+ (7.6)

    where B satisfies (7.3).

    In the next two sections, we will find solutions S and B to (7.6) and (7.3) by a fixed point argument. We will first assume that the incidence function f:R+×X+R+ is Lipschitz continuous rather than just Lipschitz continuous on bounded subsets (Remark 4.2). The argument will then be extended to incidence functions that are Lipschitz continuous on bounded subsets in Section 7.1.3 by using retractions, and solutions to the original system (2.7) will be established.

    Let u0L1+[0,b)=X+ be fixed. For given S, we treat (7.3) as a fixed point problem for B and then study how B depends on S.

    Let T(0,). Consider the vector space C[0,T] of real-valued continuous functions on [0,T] and let C+[0,T] be the cone on non-negative continuous functions.

    Lemma 7.1. For given BC+[0,T], x(t), t[0,T], in (7.3) is an element in L1+[0,b) and

    x(t)1t0B(s)ds+u01,0t<. (7.7)

    Further, x:[0,T]L1+[0,b)=X+ in (7.3) is a continuous function of t[0,T] with values in L1+[0,b).

    Proof. If b(0,), we extend u0L1+[0,b) from [0,b) to [0,) by u0(a)=0 for t[b,). The extended u0 is an element of L1+[0,). We also define u0P(a)=0 for a[b,). To show the continuity (cf. [2]), we let t[0,T] and (tn) and (sn) be sequences in [0,T] with tn,snt and sntn for all nN and show that x(tn)x(sn)10 as n. Indeed, for each nN,

    x(tn)x(sn)1sn0|B(tna)B(sna)|P(a)da+tnsnB(tna)P(a)da+tnsnu0(asn)P(asn)P(a)da+tn|u0(atn)P(atn)u0(asn)P(asn)|P(a)da.

    Since P is decreasing and takes values in [0,1], after some changes of variables,

    x(tn)x(sn)1sn0|B(tnsn+a)B(a)|da+tnsn0B(a)da+tnsn0u0(a)da+tn|u0P(atn)u0P(asn)|P(a)da.

    As n and tn,snt, all these integrals converge to 0 except possibly the last one. By the triangle inequality, the last integral satisfies

    tn|u0(atn)u0(asn)|P(a)da+btnbu0(atn)|1P(atn)1P(asn)|P(a)da,

    with tnb=min{tn,b}. After some changes of variables, this expression can be estimated by

    0|u0(a)u0(a+tnsn)|da+[btn]+0u0(a)|1P(a)1P(a+tnsn)|P(a+tn)da.

    The first integral tends to 0 because u0L1(R+) [76, X.1], and the second by the a.e. dominated convergence theorem. Since P is decreasing and bounded, P is continuous a.e. and the integrands in the second integral tend to 0 for a.e. a0. Further, the integrands are dominated by the finitely integrable function 2u0.

    In order to obtain a strict contraction for the fixed point problem (7.3), we consider exponentially weighted supremum norms on C[0,T]. Cf. [77, Sec.3.3] [46, Sec.9.1.2]. For λ(0,) (the "weight" of the norm), let

    Bλ=supt[0,T]eλt|B(t)|,BC[0,T]. (7.8)

    For λ=0, this is the supremum norm. One readily checks that λ is a norm on C[0,T]. Since eλTB0BλB0, λ is equivalent to the supremum norm, and C[0,T] is a Banach space under λ and C+[0,T] is a complete metric space under the metric induced by λ.

    We also assume that f:R+×X+R+ is Lipschitz continuous (Remark 4.2). In Section 7.1.3, we will extend the existence proof to cover incidence functions that are only Lipschitz continuous on bounded sets.

    Lemma 7.2. Let f:R+×X+R+ be Lipschitz continuous (Remark 4.2). Let T(0,) and SC+[0,T] and u0X+ be given. Then the equation (7.3) has a unique solution B=β(S)C+[0,T] which is a Lipschitz continuous function of S. Further, the function x:[0,T]L1+[0,b) defined by (7.3) is continuous.

    Further, if Λ>0 is a Lipschitz constant for f and λ>Λ, then

    β(S)β(˜S)λΛλλΛS˜Sλ,S,˜SC+[0,T]. (7.9)

    Proof. Let SC+[0,T] be given. Let FS:C+[0,T]C+[0,T] be defined by

    FS(B)(t)=f(S(t),x(t)) (7.10)

    where

    x(t)(a)={B(ta)P(a),0a<t,u0(at)P(a)P(at),0t<a.}0a<b. (7.11)

    Then x:[0,T]L1+[0,b) is continuous [2] (Lemma 7.1) and FS maps C+[0,T] into itself by Lemma 7.1. FS will turn out to be a strict contraction and the solution B of (7.3) its fixed point. Let B,˜BC[0,T] and ˜xL1+[0,b) be given by (7.11) with ˜B replacing B. By (7.11),

    x(t)˜x(t)1t0|B(ta)˜B(ta)|P(a)da. (7.12)

    By (7.10),

    |FS(B)(t)FS(˜B)(t)|=|f(S(t),x(t))f(S(t),˜x(t))|.

    Let Λ>0 be a Lipschitz constant for f; for t[0,T],

    |FS(B)(t)FS(˜B)(t)|Λx(t)˜x(t)1Λt0|B(ta)˜B(ta)|P(a)daΛt0|B(ta)˜B(ta)|da. (7.13)

    Here, we have used that P(a)1 for all aR+. As mentioned above, the norm λ defined in (7.8) is equivalent to the supremum norm on C[0,T] and thus makes C+[0,T] a complete metric space. By (7.13),

    FS(B)FS(˜B)λ=f(S,x)f(S,˜x)λΛλB˜Bλ. (7.14)

    So, for λ>Λ, FS becomes a strict contraction with respect to λ and has a fixed point

    B=β(S)=f(S,x)C+[0,T], (7.15)

    which solves (7.3).

    To show (7.9), let S,˜SC+[0,T] and B=β(S) and ˜B=β(˜S). For all t[0,T],

    |B(t)˜B(t)|=|f(S(t,x(t))f(˜S(t),˜x(t))|Λ|S(t)˜S(t)|+Λx(t)˜x(t)1,

    where ˜x is given by (7.11) with ˜B replacing B. By (7.11),

    |B(t)˜B(t)|Λ|S(t)˜S(t)|+Λt0|B(ta)˜B(ta)|da.

    By (7.8),

    B˜BλΛS˜Sλ+(Λ/λ)B˜Bλ.

    We solve this inequality for B˜Bλ and obtain (7.9).

    Let T>0. On [0,T], using β:C+[0,T]C+[0,T] from Lemma 7.2, (7.6) becomes

    S(t)=[S0+t0S(s)g(S(s))dst0(1A(ts))β(S)(s)ds+v0(t)]+=:˜F(S)(t),t[0,T]. (7.16)

    Assume that ssg(s) is Lipschitz continuous from R+ to R+ with a Lipschitz constant ˜Λ. By Lemma 7.2, for S,˜SC+[0,T],

    ˜F(S)˜F(˜S)λ˜ΛλS˜Sλ+λ1β(S)β(˜S)λ(˜Λλ+ΛλΛ)S˜Sλ.

    By choosing λ>0 large enough, ˜F becomes a strict contraction on C+[0,T] under λ and has a unique fixed point S by Banach's fixed point theorem, which is a unique solution of (7.16) on [0,T]. Since our arguments did not depend on T(0,), we find a unique solution of (7.16) on [0,T] for every T(0,).

    Notice that a solution of (7.16) on [0,T] is also a solution of (7.16) on [0,˜T] for any ˜T(0,T). Since the solutions are unique, the restriction of the solution on [0,T] to the subinterval [0,˜T], ˜T<T, is the unique solution on [0,˜T].

    Then we also have a unique solution S,B of (7.16) on [0,) and thus of (7.6) and of (7.2) with (7.3).

    Let us summarize what we have established so far.

    Proposition 7.3. Assume that f:R+×X+R+ is Lipschitz continuous, there exists some Λ(0,) such that

    |f(S,w)f(˜S,˜w)|Λ(|S˜S|+w˜w1),S,˜SR+,w,˜wX+. (7.17)

    Assume that ssg(s) is Lipschitz continuous from R+ to R+.

    Then, for any S0(0,) and u0X+, there exists a unique solution S,B:R+R+ of (7.2) and (7.3).

    Unfortunately, many incidence functions are not Lipschitz continuous, but only Lipschitz continuous on bounded subsets of R+×X+ (Remark 4.2).

    Theorem 7.4. Assume that g is Lipschitz continuous on every bounded subset of R+ and that f is Lipschitz continuous on every bounded subset of R+×X+ (Remark 4.2). Further, assume that f satisfies the following additional assumption: For any c1>1 there is some c2>0 such that

    f(S,w)c2S,0S,w1c1. (7.18)

    Then, for any S0(0,) and u0X+, there exists a unique solution S,B:R+R+ of (2.7) with S(t)>0 for all t0.

    Remark 7.5. The assumption involving (7.18) can be replaced by the assumptions of Lemma 3.2.

    Proof. Guided by the a priori estimates in Section 3.1, we proceed as follows.

    Let S0>0 and u0X+. Set

    c=max{H(0),K+1,1νmax0SKS(g(S)+ν)},H(0)=S0+u01. (7.19)

    We define the retraction ξc:R+×X+R+×X+ by

    ξc(S,w)={(S,x),|S|+w1c,cS+w1(S,w),|S|+w1c. (7.20)

    By the proof of [78, Lem.19.8]. ξc is Lipschitz continuous with a Lipschitz constant 2. Further, ξc maps R+×X+ into a bounded subset of itself.

    We define fc=fξc on R+×X+ with c in (7.19). Since f is Lipschitz continuous on bounded subsets of R+×X+, fc is Lipschitz continuous on R+×X+ and fc(S,w)=f(S,w) if SR+, wX+ and S+w1c.

    Further, we define gc:R+R by

    gc(S)=max{g(S),g(c)},SR+. (7.21)

    Since g is decreasing, gc(S)=g(S) for S[0,c] and gc(S)=g(c) for Sc. Then gc also satisfies the Assumptions 2.4, but in addition SSgc(S) is Lipschitz continuous on R+ with cΛc+|g(c)| being a Lipschitz constant for SSgc on R+ where Λc is a Lipschitz constant for g on [0,c].

    By Proposition 7.3, (7.2) and (7.3) with f replaced by fc and g replaced by gc, has a unique solution S,B on [0,) with S(t)0 and B(t)0 for all t>0.

    Suppose that S(t)>0 does not hold for all t>0. Then there exists some T>0 such that

    S(t)>0,t[0,T), and S(T)=0. (7.22)

    By (7.2),

    S(t)=S0+t0S(s)gc(S(s))dst0B(s)ds+t0R(s)ds,t[0,T].

    Then S is absolutely continuous on [0,T] and

    S(t)=S(t)gc(S(t))B(t)+R(t),B(t)=fc(S(t),x(t)) for a.a. t[0,T].

    Our a priori estimates in Section 3 also hold for the modified equation on R+ because they were independent of the incidence function and

    1νmax0SKS(g(S)+ν)

    does not change if g is replaced by gc because cK+1 by (7.19). This implies that S(t)+x(t)1c for t[0,T]. So S and B are solutions of the original system (2.7) with g and f on [0,T]. Then S(T)>0 by Lemma 3.1. This is a contradiction to (7.22) and so S(t)>0 for all t>0.

    (Alternatively, one argues with Remark 7.5 and Lemma 3.2.)

    By repeating the previous argument on [0,) instead of [0,T], S and u(t,)=x(t) satisfy the a priori estimates in Section 3.1 on R+ and are solutions of the original system (2.7) with g on f on R+ and S(t)>0 and B(t)0 for all t0.

    With the contraction mapping theorem no longer at our disposal, we consider the Volterra equations formulation of our system (Section 2.4), which is a special case of the Lotka formulation (Section 2.3). For ease of presentation, we only consider one infectious impact, but the case of finitely many infectious impacts can be done analogously. Let

    f(S,w)=˘f(S,J),J=b0κ(a)w(a)da,κL+[0,b),˘f:R2+R+ continuous. (7.23)

    Since we want to cover ˇf(S,J)=SqJp with q,p>0, p+q>1 and p<1 or q<1, we do not assume that ˘f is Lipschitz continuous on bounded subsets of R2+.

    We could try to use some appropriate version of Schauder's or Tychonov's fixed point theorem [79, Thm.8.8, 10.1] or to check the assumptions of the existence results for functional equations in [74, Sec.12.2]. Instead, we take advantage of the work we did for Lipschitz continuous incidence functions and use an approximation procedure and Arzela-Ascoli type arguments [70, Sec.8.3]. Notice that we do not assume that αL+[0,b). See Section 2.6. This will lead to some technical complications.

    Choose a sequence ˘fn:R2+R+, nN, that are Lipschitz continuous and ˘fn˘f as n uniformly on bounded subsets of R2+. This is possible by the following result.

    Let M be a metric space with metric d. For a function ˘f:MR+, define the sequence (˘fn) of functions ˘fn:MR+ by

    ˘fn(x)=inf{˘f(y)+nd(x,y);yM},xM,nN. (7.24)

    [80, Thm.3.13] [81].

    Proposition 7.6. (˘fn) is an increasing sequence of bounded Lipschitz continuous functions (with n being a Lipschitz constant of ˘fn) such that inf˘f˘fn˘f for all nN. The following convergence results hold.

    (a) If ˘f is lower semi-continuous on M, ˘fn˘f pointwise on M.

    (b) If ˘fC+(M), ˘fn˘f uniformly on compact subsets of M.

    Part (b) follows from part (a) by Dini's lemma. [82, p.225].

    We choose c(0,) as in (7.19) and set gc as in (7.21). Notice that the functions fn:R+×X+R+ given by

    fn(S,w)=˘fn(S,J),J=b0κ(a)w(a)da,nN,SR+,wX+,

    are Lipschitz continuous on R+×X+. By Proposition 7.3, in the Volterra equations formulation (Section 2.4), there are solutions Sn,Bn:R+R+ such that

    Sn(t)=[S0+t0Sn(s)gc(S(s))dst0Bn(s)ds+t0Rn(s)ds]+,Rn(t)=t0α(a)Bn(ta)P(a)da+R0(t),nN,R0(t)=b0α(r+t)u0(r)P(r+t)P(r)dr, (7.25)
    Bn(t)=˘fn(S(t),Jn(t)),Jn(t)=t0κ(a)Bn(ta)P(a)da+J0(t),nN,J0(t)=b0α(r+t)u0(r)P(r+t)P(r)dr. (7.26)

    Theorem 7.7. Let g be Lipschitz continuous on bounded sets. Assume that there is some η0(0,1) such that

    α(a)η0κ(a),a[0,b). (7.27)

    Further, assume that there are c,δ>0 such that

    ˘f(S,J)cS+η0J,S,J0,S+Jδ. (7.28)

    Then, for each S0>0 and u0X+, there exists solution to the model equations on R+ with S(t)>0 and J(t)0 for all t0.

    Proof. Since ˘fn˘f, the approximations ˇfn in inherit the inequality (7.28). Arguing as in the proof of Theorem 7.4, using Lemma 3.2, we have

    0<Sn(t)=S0+t0Sn(s)gc(S(s))dst0Bn(s)ds+t0Rn(s)ds,tR+. (7.29)

    These solutions satisfy the a priori estimates in Section 3.1, which do not depend on the incidence functions. So {Sn(t);tR+nN} is a bounded subset of R+ and {xn(t);nN,tR+} is a bounded subset of L1+[0,b).

    Since ˘fn˘f and ˘f is bounded on bounded subsets of R2+, {Bn(t);tR+,nN} is a bounded subset of R+.

    By (7.29), {Sn:R+R+;nN} is equicontinuous. Also, {Sn(t)N,tR+} has compact closure in R+.

    By [70, Thm.8.5], after choosing subsequences, we get SnS, S:R+R+, as n, uniformly on bounded subsets of Rn.

    Let rR+ and 0rst. After a substitution,

    Rn(t)Rn(r)(R0(t)R0(r))=t0α(ta)P(ta)Bn(a)dar0α(ra)P(ra)Bn(a)da.

    By the triangle inequality,

    |Rn(t)R)(t)(Rn(r)R0(r))|=|t0α(ta)P(ta)Bn(a)dar0α(ra)P(ra)Bn(a)da|˜cr0|α(ta)P(ta)α(ra)P(ra)|da+˜ctrα(ta)P(ta)da,

    with some ˜c(0,) that does not depend on n. Since 0α(a)P(a)da1 (2.18), the last expression goes to zero as tr0 [76, X.1] and does not depend on n.

    So {RnR0;nN} is equicontinuous. Also, {Rn(t)R0(t);nN,t0} are bounded subsets of R.

    After choosing subsequences, using [70, Thm.8.5], RnR as n uniformly on bounded subsets of R+ for some measurable R on R+. Similarly, JnJ as n uniformly on bounded subsets of R+ for some measurable J on R+.

    By (7.29), {˜Sn;nN} with ˜Sn(t)=Sn(t)t0Rn(s)ds is equicontinuous and {˜Sn(t);nN,t0} is bounded. After choosing subsequences again, using [70, Thm.8.5] once more, ˜Sn˜S as n uniformly on bounded subsets of R+ with a continuous function ˜S. So Sn(t)˜S(t)+t0R(s)ds as n uniformly for t in bounded subsets of R+.

    Since the solutions satisfy the a priori estimates of Section 3.1 and ˘fn˘f uniformly on bounded subsets of R2+, Bn=˘fn(Sn,Jn)˘f(S,J)=:B as n uniformly on bounded subsets of R. Notice that the equations are also satisfied with g replacing gc.

    Now we can take limits as n in (7.25) and (7.26) and obtain

    S(t)=S0+t0S(s)g(S(s))dst0B(s)ds+t0R(s)ds,R(t)=t0α(a)B(ta)P(a)da+b0α(r+t)u0(r)P(r+t)P(r)dr.B(t)=˘f(S(t),J(t)),J(t)=t0κ(a)B(ta)P(a)da+b0κ(r+t)u0(r)P(r+t)P(r)dr.

    This is the Volterra equations formulation of our model system (Section 2.4).

    In this paper, we discuss the fundamental question of whether infectious diseases can drive their host population into extinction.

    Previous work [1,26], using planar SIS type ODEs, shows that this is never the case (in a deterministic setting) if the incidence function is of the generalized density-dependent type (unless the population is subject to an Allee effect or there is a disease reservoir) but that it can happen for generalized frequency-dependent incidence functions.

    The choice of general incidence functions is motivated by the fact that density-dependent incidence and frequency-dependent incidence fit the infection experiments for salamander larvae and ranavirus badly.

    The results are largely the same whether or not infected individuals can become susceptible again. An exception is the power incidence f(S,J)=SqIp with p,q>0, p+q>1 and q<1.

    If infected individuals do not become susceptible again, there is initial-value dependent extinction of the host (Theorem 2.9), but if some infected individuals become susceptible again, the host persists uniformly weakly.

    This paper discusses whether the previous results still hold if, more realistically and in the spirit of Kermack and McKendrick [3,31,32], infected individuals do not leave the infected stage at a constant per capita rate (exponentially distributed length of the infected stage) but at a per capita rate that depends on infection-age (time since the moment of infection) and if the infectivity of an infected individual also depends on its infection-age.

    The findings are that the previous results essentially extend to generalizations of density-dependent incidence functions. Generalizations of frequency-dependent incidence functions were not considered because there are too many different ways to combine frequency-dependence and infection-age.

    For certain power incidences, it is sufficient for uniform weak host persistence that, at any infection-age a, if some infected individuals are infectious at age a there are also some infected individuals that leave the infected class at age a and become susceptible again (Example 2.8). This condition seems to be difficult to verify biologically, though it may hold for some diseases. The ODE results show that, mathematically, this assumption cannot be dropped completely (Theorem 2.9). The question is to what degree it is necessary. We could not answer this question analytically; perhaps numerical simulations of some well-chosen special cases can give some insight in the future.

    We also mention that, for the host population dynamics, it is assumed (as in [2,26]) that infected individuals do not give birth and do not compete for resources. This somewhat extreme assumption has been made for mathematical convenience; it allows us to consider the problem in terms of the number of susceptibles and the disease incidence.

    In this paper, the survival of the susceptible host population is established in the form of uniform weak persistence and not in the form of the more desirable uniform persistence (see Section 1.6).

    The weaker form could be proved because a convenient tool for doing this is available, the Laplace transform. Routes from uniform weak to uniform persistence are described in [46, Chap.4]. Direct routes to uniform persistence can be found in [50, Sec.1.3].

    However one proceeds in proving uniform persistence, it is useful to show that the solutions induce a semiflow on R+×X+. This would be facilitated by the uniqueness of solutions, e.g., by the Lipschitz continuity assumptions made in Section 7.2. One can use the model formulation in Section 2.3 and proceed as in [46, Sec.9.4] for the special case of density-dependent incidence. Alternatively, one can interpret the model formulation in Section 2.2 as a Cauchy problem with a non-densely defined operator and use the theory of integrated semigroups [83,84]. Further, something like a compact attractor is helpful [46,49,50]. This should be possible to obtain if the model can be formulated as Volterra integral equations (Section 2.4) proceeding similarly as in the special case of a density-dependent incidence in [46, 9.5].

    We conjecture that the uniform persistence of susceptible hosts holds in Theorem 2.7 if g and ˘f are Lipschitz continuous on bounded subsets of R+ and R2+, respectively. We do not know whether the uniform persistence of susceptible hosts holds in Example 2.8, because the incidence function is not Lipschitz continuous in the neighborhood of (0,0) and it is difficult to show that the solutions induce a semiflow because they may not be uniquely determined by the initial data.

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

    The authors declare no conflict of interest.

    The authors thank Odo Diekmann and two anonymous referees for their constructive comments.

    Lemma 8.1. Let f:[0,b)IR be absolutely continuous and g:IR be continuously differentiable.

    Then gf is absolutely continuous and (gf)(x)=g(f(x))f(x) for almost all x[a,b).

    Proof. Let ˜b[0,b). For nN with 1/n<b˜b and x[0,˜b], define

    fn(x)=nx+(1/n)xf(y)dy=10f(x+(z/n))dz.

    Then fn is continuously differentiable and

    fn(x)=n(f(x+(1/n))f(x))=nx+(1/n)xf(y)dy=10f(x+(z/n))dy.

    Since f is finitely integrable.

    fnf,n,a.e.

    By the chain rule, gfn is continuously differentiable and

    (gfn)(x)=g(fn(x))fn(x).

    Further, fn(x)f(x) as n. So

    g(f(x))g(f(0))=limng(fn(x))g(fn(0))=limnx0g(fn(y))fn(y)dy.

    As for the last expression,

    |x0g(fn(y))fn(y)dyx0g(f(y))f(y)dy|x0|g(fn(y))g(f(y))||fn(y)|dy+x0|g(f(y))||f(y)fn(y)|dy.

    The first expression tends to 0 because fnf as n uniformly on [0,x] and g is uniformly continuous on bounded closed intervals and thus g(fn(y))g(f(y)) uniformly on [0,x]. As for the second expression,

    x0|g(f(y))||f(y)fn(y)|dycx0|10(f(y)f(y+(z/n))dz|dyc10(x0|f(y)f(y+(z/n))|dy)dz.

    For each x,

    x0|f(y)f(y+(z/n))|dy0,n,

    because f is finitely integrable on every subinterval of [0,b). By the dominated convergence theorem, it remains to show that these expressions that depend on n can be dominated by an integrable function. Indeed, for all z[0,1] and all nN with x+(1/n)<b.

    0x0|f(y)f(y+(z/n))|dyx0|f(y)|dy+x0|f(y+(z/n)|dyx0|f(y)|dy+x+(z/n)z/n|f(y)|dy2b0|f(y)|dy.



    [1] A. P. Farrell, J. P. Collins, A. L. Greer, H. R. Thieme, Do fatal infectious diseases eradicate host species?, J. Math. Biol., 77 (2018), 2103–2164. https://doi.org/10.1007/s00285-018-1249-3 doi: 10.1007/s00285-018-1249-3
    [2] A. P. Farrell, J. P. Collins, A. L. Greer, H. R. Thieme, Times from infection to disease-induced death and their influence on final population sizes after epidemic outbreaks, Bull. Math. Biol., 80 (2018), 1937–1961. https://doi.org/10.1007/s11538-018-0446-y doi: 10.1007/s11538-018-0446-y
    [3] W. O. Kermack and A. G. McKendrick, A contribution to the mathematical theory of epidemics, in Proceedings of the Royal Society of London. Series A, Containing papers of a mathematical and physical character, 115 (1927), 700–721. https://doi.org/10.1098/rspa.1927.0118
    [4] H. R. Thieme, Epidemic and demographic interaction in the spread of potentially fatal diseases in growing populations, Math. Biosci., 111 (1992), 99–130. https://doi.org/10.1016/0025-5564(92)90081-7 doi: 10.1016/0025-5564(92)90081-7
    [5] L. Han, A. Pugliese, Epidemics in two competing species, Nonlinear Anal. Real World Appl., 10 (2009), 723–744. https://doi.org/10.1016/j.nonrwa.2007.11.005 doi: 10.1016/j.nonrwa.2007.11.005
    [6] H. W. Hethcote, W. Wang, Y. Li, Species coexistence and periodicity in host-host-pathogen models, J. Math. Biol., 51 (2005), 629–660. https://doi.org/10.1007/s00285-005-0335-5 doi: 10.1007/s00285-005-0335-5
    [7] R. D. Holt, J. Pickering, Infectious disease and species coexistence: A model of Lotka-Volterra form, Am. Natur., 126 (1985), 196–211. https://doi.org/10.1086/284409 doi: 10.1086/284409
    [8] A. Friedman, A. A. Yakubu, Host demographic Allee effect, fatal disease, and migration: Persistence or extinction, SIAM J. Appl. Math., 72 (2012), 1644–1666. https://doi.org/10.1137/120861382 doi: 10.1137/120861382
    [9] F. M. Hilker, Population collapse to extinction: The catastrophic combination of parasitism and Allee effect, J. Biol. Dynam., 4 (2010), 86–101. https://doi.org/10.1080/17513750903026429 doi: 10.1080/17513750903026429
    [10] H. R. Thieme, T. Dhirasakdanon, Z. Han, R. Trevino, Species decline and extinction: Synergy of infectious disease and Allee effect?, J. Biol. Dynam., 3 (2009), 305–323. https://doi.org/10.1080/17513750802376313 doi: 10.1080/17513750802376313
    [11] S. Busenberg, K. Cooke, H. R. Thieme, Demographic change and persistence of HIV/AIDS in a heterogeneous population, SIAM J. Appl. Math., 51 (1991), 1030–1052. https://doi.org/10.1137/0151052 doi: 10.1137/0151052
    [12] S. Busenberg, P. van den Driessche, Analysis of a disease transmission model in a population with varying size, J. Math. Biol., 28 (1990), 257–270. https://doi.org/10.1007/BF00178776 doi: 10.1007/BF00178776
    [13] L. Q. Gao, H. W. Hethcote, Disease transmission models with density-dependent demographics, J. Math. Biol., 30 (1992), 717–731. https://doi.org/10.1007/BF00173265 doi: 10.1007/BF00173265
    [14] W. M. Getz, J. Pickering, Epidemic models: Thresholds and population regulation, Am. Natur., 121 (1983), 892–898. https://doi.org/10.1086/284112 doi: 10.1086/284112
    [15] D. Greenhalgh, R. Das, An SIR epidemic model with a contact rate depending on population density, Math. Popul. Dynam. Anal. Heterog., 1 (1995), 79–101.
    [16] K. P. Hadeler, K. Dietz, M. Safan, Case fatality models for epidemics in growing populations, Math. Biosci., 281 (2016), 120–127. https://doi.org/10.1016/j.mbs.2016.09.007 doi: 10.1016/j.mbs.2016.09.007
    [17] T.-W. Hwang, Y. Kuang, Deterministic extinction effect of parasites on host populations, J. Math. Biol., 46 (2003), 17–30. https://doi.org/10.1007/s00285-002-0165-7 doi: 10.1007/s00285-002-0165-7
    [18] T.-W. Hwang, Y. Kuang, Host extinction dynamics in a simple parasite-host interaction model, Math. Biosci. Eng., 2 (2005), 743–751. https://doi.org/10.3934/mbe.2005.2.743 doi: 10.3934/mbe.2005.2.743
    [19] R. M. May, R. M. Anderson, A. R. McLean, Possible demographic consequences of HIV/AIDS epidemics: ii, assuming HIV infection does not necessarily lead to AIDS, in Mathematical Approaches to Problems in Resource Management and Epidemiology, Springer, 1989,220–248. https://doi.org/10.1007/978-3-642-46693-9_16
    [20] J. Mena-Lorca, H. W. Hethcote, Dynamic models of infectious diseases as regulators of population sizes, J. Math. Biol., 30 (1992), 693–716. https://doi.org/10.1007/BF00173264 doi: 10.1007/BF00173264
    [21] J. Zhou, H. W. Hethcote, Population size dependent incidence in models for diseases without immunity, J. Math. Biol., 32 (1994), 809–834. https://doi.org/10.1007/BF00168799 doi: 10.1007/BF00168799
    [22] O. Diekmann, H. Heesterbeek, T. Britton, Mathematical Tools for Understanding Infectious Disease Dynamics, Princeton University Press, Princeton, 2013. https://doi.org/10.23943/princeton/9780691155395.001.0001
    [23] F. De Castro, B. Bolker, Mechanisms of disease-induced extinction, Ecol. Letters, 8 (2005), 117–126. https://doi.org/10.1111/j.1461-0248.2004.00693.x doi: 10.1111/j.1461-0248.2004.00693.x
    [24] K. F. Smith, D. F. Sax, K. D. Lafferty, Evidence for the role of infectious disease in species extinction and endangerment, Conserv. Biol., 20 (2006), 1349–1357. https://doi.org/10.1111/j.1523-1739.2006.00524.x doi: 10.1111/j.1523-1739.2006.00524.x
    [25] A. L. Greer, C. J. Briggs, J. P. Collins, Testing a key assumption of host-pathogen theory: density and disease transmission, Oikos, 117 (2008), 1667–1673. https://doi.org/10.1111/j.1600-0706.2008.16783.x doi: 10.1111/j.1600-0706.2008.16783.x
    [26] B. H. Carlson, H. R. Thieme, Can infectious diseases that do not confer immunity eradicate host species?, J. Biol. Syst. (to appear). https://doi.org/10.1142/S0218339023400016
    [27] A. Taylor, W. Mann, Advanced Calculus, Sec. ed., Xerox College Publishing, Lexington, 1972.
    [28] P. E. Sartwell, The distribution of incubation periods of infectious disease, Am. J. Hyg., 51 (1950), 310–318. https://doi.org/10.1093/oxfordjournals.aje.a119397 doi: 10.1093/oxfordjournals.aje.a119397
    [29] P. E. Sartwell, The incubation period and the dynamics of infectious disease, Am. J. Epidemiol., 83 (1966), 204–216. https://doi.org/10.1093/oxfordjournals.aje.a120576 doi: 10.1093/oxfordjournals.aje.a120576
    [30] Z. Feng, Applications of Epidemiological Models to Public Health Policymaking: The Role of Heterogeneity in Model Predictions, World Scientific, Singapore, 2014. https://doi.org/10.1142/8884
    [31] W. O. Kermack, A. G. McKendrick, Contributions to the mathematical theory of epidemics. ii. The problem of endemicity, in Proceedings of the Royal Society of London. Series A, containing papers of a mathematical and physical character, 138 (1932), 55–83. https://doi.org/10.1098/rspa.1932.0171
    [32] W. O. Kermack, A. G. McKendrick, Contributions to the mathematical theory of epidemics. iii. Further studies of the problem of endemicity, in Proceedings of the Royal Society of London. Series A, Containing papers of a mathematical and physical character, 141 (1933), 94–122. https://doi.org/10.1098/rspa.1933.0106
    [33] H. R. Thieme, C. Castillo-Chavez, How may infection-age-dependent infectivity affect the dynamics of HIV/AIDS?, SIAM J. Appl. Math., 53 (1993), 1447–1479. https://doi.org/10.1137/0153068 doi: 10.1137/0153068
    [34] H. Inaba, Age-structured Population Dynamics in Demography and Epidemiology, Springer, Singapore, 2017. https://doi.org/10.1007/978-981-10-0188-8
    [35] M. Martcheva, An Introduction to Mathematical Epidemiology, Springer, New York, 2015. https://doi.org/10.1007/978-1-4899-7612-3
    [36] H. R. Thieme, J. Yang, An endemic model with variable re-infection rate and applications to influenza, Math. Biosci., 180 (2002), 207–235. https://doi.org/10.1016/S0025-5564(02)00102-5 doi: 10.1016/S0025-5564(02)00102-5
    [37] F. Hoppensteadt, Mathematical Theories of Populations: Demographics Genetics and Epidemics, Society for Industrial and Applied Mathematics, Philadelphia, 1975. https://doi.org/10.1137/1.9781611970487
    [38] S. Luckhaus, A. Stevens, A free boundary problem-in time-for the spread of Covid-19, J. Math. Biol., 86 (2023), 17. https://doi.org/10.1007/s00285-023-01881-0 doi: 10.1007/s00285-023-01881-0
    [39] S. Luckhaus, A. Stevens, Kermack and McKendrick models on a two-scale network and connections to the Boltzmann equations, in Mathematics Going Forward: Collected Mathematical Brushstrokes, Springer, 2022,417–427. https://doi.org/10.1007/978-3-031-12244-6_29
    [40] Z. Ma, J. Li, Dynamical Modeling and Analysis of Epidemics, World Scientific, Singapore, 2009.
    [41] F. Abbona, E. Venturino, An eco-epidemic model for infectious keratoconjunctivitis caused by Mycoplasma conjunctivae in domestic and wild herbivores, with possible vaccination strategies, Math. Methods Appl. Sci., 41 (2018), 2269–2280. https://doi.org/10.1002/mma.4209 doi: 10.1002/mma.4209
    [42] T. Dhiraksakdanon, A model of infectious diseases in amphibian populations with ephemeral larval habitat, Dissertation, Arizona State University, 2010.
    [43] O. Diekmann, Thresholds and traveling waves for the geographical spread of infection, J. Math. Biol., 6 (1978), 109–130. https://doi.org/10.1007/BF02450783 doi: 10.1007/BF02450783
    [44] H. R. Thieme, A model for the spatial spread of an epidemic, J. Math. Biol., 4 (1977), 337–351. https://doi.org/10.1007/BF00275082 doi: 10.1007/BF00275082
    [45] H. R. Thieme, Mathematics in Population Biology, Princeton University Press, Princeton, NJ, USA, 2003.
    [46] H. L. Smith, H. R. Thieme, Dynamical Systems and Population Persistence, American Mathematical Soc., 2011. https://doi.org/10.1090/gsm/118
    [47] A. Farrell, Prey-predator-parasite: An ecosystem model with fragile persistence, Arizona State University, 2017.
    [48] J. A. P. Heesterbeek, J. A. Metz, The saturating contact rate in marriage-and epidemic models, J. Math. Biol., 31 (1993), 529–539. https://doi.org/10.1007/BF00173891 doi: 10.1007/BF00173891
    [49] 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
    [50] X.-Q. Zhao, Dynamical Systems in Population Biology, Springer, New York, 2003.
    [51] C. K. Busenberg S, Vertically Transmitted Diseases. Models and Dynamics, Springer, Berlin Heidelberg, 1993. https://doi.org/10.1007/978-3-642-75301-5
    [52] V. Capasso, Mathematical Structures of Epidemic Systems, Springer, Berlin Heidelberg, 1993. https://doi.org/10.1007/978-3-540-70514-7
    [53] D. Breda, O. Diekmann, W. De Graaf, A. Pugliese, R. Vermiglio, On the formulation of epidemic models (an appraisal of Kermack and McKendrick), J. Biol. Dynam., 6 (2012), 103–117. https://doi.org/10.1080/17513758.2012.716454 doi: 10.1080/17513758.2012.716454
    [54] O. Diekmann, The 1927 epidemic model of Kermack and Mckendrick: A success story or a tragicomedy?, SMB Newsletter, 92, 1–4.
    [55] H. R.Thieme, Disease extinction and disease persistence in age-structured epidemic models, Nonlinear Anal. Theory Methods Appl., 47 (2001), 6181–6194. https://doi.org/10.1016/S0362-546X(01)00677-0 doi: 10.1016/S0362-546X(01)00677-0
    [56] E. Venturino, Ecoepidemiology: A more comprehensive view of population interactions, Math. Model. Nat. Phenom., 11 (2016), 49–90. https://doi.org/10.1051/mmnp/201611104 doi: 10.1051/mmnp/201611104
    [57] L. Han, Z. Ma, H. W. Hethcote, Four predator-prey models with infectious diseases, Math. Computer Model., 34 (2001), 849–858. https://doi.org/10.1016/S0895-7177(01)00104-2 doi: 10.1016/S0895-7177(01)00104-2
    [58] O. Diekmann, M. Kretzschmar, Patterns in the effects of infectious diseases on population growth, J. Math. Biol., 29 (1991), 539–570. https://doi.org/10.1007/BF00164051 doi: 10.1007/BF00164051
    [59] T. Dhirasakdanon, H. R. Thieme, Stability of the endemic coexistence equilibrium for one host and two parasites, Math. Model. Nat. Phenom., 5 (2010), 109–138. https://doi.org/10.1051/mmnp/20105606 doi: 10.1051/mmnp/20105606
    [60] L. J. S. Allen, E. J. Allen, Deterministic and stochastic SIR epidemic models with power function transmission and recovery rates, Math. Contin. Discrete Dyn. Syst., 618 (2014), 1–15. https://doi.org/10.1090/conm/618/12329 doi: 10.1090/conm/618/12329
    [61] O. Diekmann, H. Inaba, A systematic procedure for incorporating separable static hetero- geneity into compartmental epidemic models, J. Math. Biol., 86 (2023), 29. https://doi.org/10.1007/s00285-023-01865-0 doi: 10.1007/s00285-023-01865-0
    [62] A. S. Novozhilov, On the spread of epidemics in a closed heterogeneous population, Math. Biosci., 215 (2008), 177–185. https://doi.org/10.1016/j.mbs.2008.07.010 doi: 10.1016/j.mbs.2008.07.010
    [63] N. T. Bailey, The Mathematical Theory of Infectious Diseases and its Applications, Charles Griffin & Company Ltd, 5a Crendon Street, High Wycombe, 1975.
    [64] F. Brauer, C. Castillo-Chavez, Z. Feng, Mathematical Models in Epidemiology, Springer, New York, 2019. https://doi.org/10.1007/978-1-4939-9828-9
    [65] A. Ducrot, Q. Griette, Z. Liu, P. Magal, Differential Equations and Population Dynamics I, Springer, Cham, 2022. https://doi.org/10.1007/978-3-030-98136-5
    [66] X.-Z. Li, J. Yang, d M. Martcheva, Age Structured Epidemic Modeling, vol. 52 makes no sense unless the series is mentioned as well, Springer Nature, Cham, 2020.
    [67] L. Rass, J. Radcliffe, Spatial Deterministic Epidemics, American Mathematical Soc., Providence, 2003. https://doi.org/10.1090/surv/102
    [68] K. P. Hadeler, Topics in Mathematical Biology, Springer, Cham, 2017. https://doi.org/10.1007/978-3-319-65621-2
    [69] M. Iannelli, F. Milner, The Basic Approach to Age-Structured Population Dynamics. Models, Methods and Numerics, Springer, Dordrecht, 2017. https://doi.org/10.1007/978-94-024-1146-1
    [70] J. McDonald, N. Weiss, A Course in Real Analysis, Academic Press, San Diego, 1999.
    [71] A. G. McKendrick, Applications of mathematics to medical problems, Proceed. Edinburgh Math. Soc., 14 (1926), 98–130. https://doi.org/10.1017/S0013091500034428 doi: 10.1017/S0013091500034428
    [72] A. J. Lotka, Relation between birth rates and death rates, Science, 26 (1907), 21–22. https://doi.org/10.1126/science.26.653.21.b doi: 10.1126/science.26.653.21.b
    [73] E. Hewitt and K. Stromberg, Real and Abstract analysis, Springer, Berlin Heidelberg, 1956.
    [74] G. Gripenberg, S.-O. Londen, O. Staffans, Volterra Integral and Functional Equations, Cambridge University Press, Cambridge, 1990. https://doi.org/10.1017/CBO9780511662805
    [75] W. M. Hirsch, H. Hanisch, J.-P. Gabriel, Differential equation models of some parasitic infections: methods for the study of asymptotic behavior, Commun. Pure Appl. Math., 38 (1985), 733–753. https://doi.org/10.1002/cpa.3160380607 doi: 10.1002/cpa.3160380607
    [76] K. Yosida, Functional Analysis, Sec. edition, Springer, Berlin Heidelberg, 1965–1968. https://doi.org/10.1007/978-3-662-25762-3
    [77] H. R. Thieme, Asymptotic estimates of the solutions of nonlinear integral equations and asymptotic speeds for the spread of populations, J. fur die Reine und Angewandte Mathematik, 1979 (1979), 94–121. https://doi.org/10.1515/crll.1979.306.94 doi: 10.1515/crll.1979.306.94
    [78] H. Amann, Gewöhnliche Differentialgleichungen, De Gruyter, Berlin, 1983.
    [79] K. Deimling, Nonlinear Functional Analysis, Springer, Berlin Heidelberg, 1985. https://doi.org/10.1007/978-3-662-00547-7
    [80] C. D. Aliprantis, K. C. Border, Infinite Dimensional Analysis: A Hitchhiker's Guide, Springer, Berlin Heidelberg, 1999, 2006. https://doi.org/10.1007/978-3-662-03961-8
    [81] P. Gwiazda, A. Marciniak-Czochra and H. R. Thieme, Measures under the flat norm as ordered normed vector space, Positivity, 22 (2018), 105–138. https://doi.org/10.1007/s11117-017-0503-z doi: 10.1007/s11117-017-0503-z
    [82] C. Düll, P. Gwiazda, A. Marciniak-Czochra, J. Skrzeczkowski, Spaces of Measures and their Applications to Structured Population Models, Cambridge University Press, Cambridge, 2021. https://doi.org/10.1017/9781009004770
    [83] P. Magal, S. Ruan, Theory and applications of abstract semilinear Cauchy problems, Springer, Cham, 2018. https://doi.org/10.1007/978-3-030-01506-0
    [84] H. R. Thieme, Semiflows generated by Lipschitz perturbations of non-densely defined operators, Differ. Integral Equat., 3 (1990), 1035–1066. https://doi.org/10.57262/die/1379101977 doi: 10.57262/die/1379101977
  • 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(1653) PDF downloads(49) Cited by(0)

Article outline

Figures and Tables

Tables(1)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog