Loading [MathJax]/jax/element/mml/optable/MathOperators.js
Research article Special Issues

Dynamics of a general model of host-symbiont interaction

  • We consider a model of host-symbiont interactions, in which symbionts can only live in association with their host and are transmitted both vertically from associated hosts to their offspring and horizontally from associated hosts to nearby unassociated hosts. The effect of the symbiont is modelled by a change in the birth rate of associated hosts. We analyze the two-dimensional dynamics in the resulting four-dimensional parameter space, and determine the qualitative behaviour for all parameter values. We find that for all but one choice of parameter values, solutions in the feasible region, apart from a 0- or 1-dimensional set of initial conditions, tend either to a unique equilibrium, or to one of two distinct equilibria. Moreover, the bistable case occurs only when the symbiont is a mutualist whose horizontal spread rate through the host population exceeds the positive change in the birth rate of associated hosts.

    Citation: Eric Foxall. Dynamics of a general model of host-symbiont interaction[J]. Mathematical Biosciences and Engineering, 2019, 16(4): 3047-3070. doi: 10.3934/mbe.2019151

    Related Papers:

    [1] Conrad Ratchford, Jin Wang . Multi-scale modeling of cholera dynamics in a spatially heterogeneous environment. Mathematical Biosciences and Engineering, 2020, 17(2): 948-974. doi: 10.3934/mbe.2020051
    [2] Jessica L. Hite, André M. de Roos . Pathogens stabilize or destabilize depending on host stage structure. Mathematical Biosciences and Engineering, 2023, 20(12): 20378-20404. doi: 10.3934/mbe.2023901
    [3] Bruno Buonomo, Marianna Cerasuolo . The effect of time delay in plant--pathogen interactions with host demography. Mathematical Biosciences and Engineering, 2015, 12(3): 473-490. doi: 10.3934/mbe.2015.12.473
    [4] Kale Davies, Suzanne Lenhart, Judy Day, Alun L. Lloyd, Cristina Lanzas . Extensions of mean-field approximations for environmentally-transmitted pathogen networks. Mathematical Biosciences and Engineering, 2023, 20(2): 1637-1673. doi: 10.3934/mbe.2023075
    [5] Rocio Caja Rivera, Shakir Bilal, Edwin Michael . The relation between host competence and vector-feeding preference in a multi-host model: Chagas and Cutaneous Leishmaniasis. Mathematical Biosciences and Engineering, 2020, 17(5): 5561-5583. doi: 10.3934/mbe.2020299
    [6] Beryl Musundi . An immuno-epidemiological model linking between-host and within-host dynamics of cholera. Mathematical Biosciences and Engineering, 2023, 20(9): 16015-16032. doi: 10.3934/mbe.2023714
    [7] Holly Gaff . Preliminary analysis of an agent-based model for a tick-borne disease. Mathematical Biosciences and Engineering, 2011, 8(2): 463-473. doi: 10.3934/mbe.2011.8.463
    [8] Sukhitha W. Vidurupola, Linda J. S. Allen . Basic stochastic models for viral infection within a host. Mathematical Biosciences and Engineering, 2012, 9(4): 915-935. doi: 10.3934/mbe.2012.9.915
    [9] Tzy-Wei Hwang, Yang Kuang . Host Extinction Dynamics in a Simple Parasite-Host Interaction Model. Mathematical Biosciences and Engineering, 2005, 2(4): 743-751. doi: 10.3934/mbe.2005.2.743
    [10] 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
  • We consider a model of host-symbiont interactions, in which symbionts can only live in association with their host and are transmitted both vertically from associated hosts to their offspring and horizontally from associated hosts to nearby unassociated hosts. The effect of the symbiont is modelled by a change in the birth rate of associated hosts. We analyze the two-dimensional dynamics in the resulting four-dimensional parameter space, and determine the qualitative behaviour for all parameter values. We find that for all but one choice of parameter values, solutions in the feasible region, apart from a 0- or 1-dimensional set of initial conditions, tend either to a unique equilibrium, or to one of two distinct equilibria. Moreover, the bistable case occurs only when the symbiont is a mutualist whose horizontal spread rate through the host population exceeds the positive change in the birth rate of associated hosts.


    Symbiosis is an important aspect of ecology [1]. In fact, an entire journal, titled Symbiosis, is devoted to the study of symbiotic interactions at all levels ranging from the molecular to the organismic. The precise meaning assigned to the term symbiosis may vary depending on context; a fairly general definition, given in [2], is a close physical association between different species, regardless of whether that association is harmful, beneficial or neutral to any of the species involved.

    In many cases of interest, the symbiosis is between a host, and one or more types of symbiont living within the host, often in large numbers. Two examples, both relating to humans, are infectious diseases [3,4] and gut microbe interactions [5]. A harmful symbiont is a pathogen, while a beneficial symbiont is a mutualist. Previous mathematical studies have focused on studying either the pathogen case [6] or the mutualist case [7,8], however it would be useful to understand, from a unified perspective, how the nature of the host-symbiont relationship affects the range of observed interactions.

    To this end, we study a simple two-variable model of host-symbiont interactions with four parameters (described in the next section) that includes both horizontal and vertical transmission of the symbiont, and naturally incorporates both the pathogen and mutualist cases, as well as the intermediate neutral (neither harmful nor beneficial) case, modelled by the symbiont's effect on the host birth rate. As well, we assume recovery is possible; that is, it is possible for the symbiont within a host to die without killing the host. For simplicity, we assume perfect vertical transmission of the symbiont, and distinguish only presence or absence of the symbiont within the host, tracking the population density of both unassociated hosts (those without the symbiont) and associated hosts. A generalization to imperfect vertical transmission is also easily implemented, though we have not considered it here.

    A special case of our model appears in [9], in which the authors discuss how it is possible, with a mixture of horizontal and vertical transmission, for a pathogen to successfully invade and persist within a host population, and they also give precise conditions for the pathogen to reach full occupation. In their model, only the pathogen case is considered, and recovery from the symbiont does not occur. In addition, the concern is with existence of certain equilibria and so a full description of the dynamics is not given. However, they also consider the case of imperfect transmission which is not covered here. From a different perspective, a model of mutualism [7], which is similar though not identical to this model, shows a qualitatively similar nullcline structure in some cases that leads to bistability; interestingly, this similarity is only observed after a change of variables is made in our model, from associated and unassociated host density to total host density and proportion of hosts that are associated.

    For our model we are able to obtain a complete description of the dynamics for all values of the four parameters. In all but a single case, every trajectory tends to one of at most three distinct equilibria, ruling out sustained oscillations and reducing the study of the model to the characterization of equilibria. In turn, the number and location of equilibria is closely related to two factors:

    ● the effect of the symbiont upon the host (pathogen, mutualist or neutral), and

    ● which of the two is larger: the difference in the birth rate of associated vs. unassociated hosts (which is positive only for mutualists), and the horizontal rate of spread of the symbiont through the host population.

    The second quantity is shown to be relevant only after making the above-mentioned change of variables. One way to view it is by how the symbiont allocates resources: is it focused on increasing the host's rate of birth and thus vertical transmission, or is it focused on spreading immediately to existing hosts via horizontal transmission?

    An important finding for this model is that bistability (between host-symbiont coexistence and either host extinction or host survival only) can only occur when the symbiont is a mutualist that is more focused on horizontal transmission than on improving host birth rate (i.e. when the second quantity above is positive), and that in the absence of bistability, all solutions with a positive density of both host and symbiont tend to a unique equilibrium. Moreover, in the absence of bistability the set of qualitative behaviours displayed by the system is insensitive as to whether the symbiont is a pathogen or a mutualist, except in the difference in host density between the host-without-symbiont equilibrium vs. the host-symbiont coexistence equilibrium, when both exist.

    The original motivation for making a detailed study of this model was to obtain complementary results for a study of the corresponding stochastic spatial model [10]. When working on that paper, a preliminary analysis of the differential equations yielded partial but artifically restrictive results. The reason for this restriction was revealed only after making the change of variables mentioned above, from associated vs. unassociated host density to total host density vs. proportion of hosts that are associated, which made evident the occurrence of bistability, which previously we had thought was not possible. Thus one conclusion of this analysis is that the (biological) perspective of host-symbiont interactions as being a "stacked" system, in which hosts have their own birth-death dynamics, while symbionts within the host population have a similar birth-death dynamics taking place over the host population, is also a useful perspective when analyzing the dynamical system. It is worth noting that the stacked perspective is also used in [11] and indeed in our paper [10], where it assists in the analysis of the stochastic spatial model by suggesting comparisons with simpler models.

    For simplicity, we ignore environmental transmission of the symbiont, and distinguish only presence or absence of the symbiont within the host. Unassociated hosts reproduce at some rate, and all hosts die at some rate which is normalized to 1. Associated hosts can infect unassociated hosts, and can also recover and become unassociated. Moreover, associated hosts reproduce at some possibly different rate from associated hosts. We assume, again for simplicity, perfect vertical transmission of the symbiont, so that offspring of associated parents are always associated. The symbiont is identified as pathogen, mutualist, or neutral by its effect on the host's birth rate.

    To derive the differential equations that we study, we begin with a set of N sites each of which can be either empty, occupied by an unassociated host, or occupied by an associated host. Each unassociated host attempts to give birth to an unassociated host onto a site chosen uniformly at random at rate λ10, being successful if that site is empty. Similarly, each associated host attempts to give birth to an associated host at rate λ20. Each host dies at rate 1, while each associated host becomes an unassociated host at rate δ. Each associated host attempts to transmit the symbiont to a randomly chosen site at rate λ21, being successful if the recipient is an unassociated host. Let U(N)0(t),U(N)1(t),U(N)2(t) denote the number of empty sites, unassociated hosts and associated hosts, respectively. Since 2i=0U(N)i(t)=N, we find that (U(N)1(t),U(N)2(t)) is a Markov chain with the following transitions.

    (U1,U2)(U1,U2)+{(1,0)at rateλ10U0U1/N(1,0)at rateU1(0,1)at rateλ20U0U2/N(0,1)at rateU2(1,1)at rateλ21U1U2/N(1,1)at rateδU2

    If we rescale to proportions u(N)i(t)=U(N)i(t)/N, then a result of Kurtz [12] implies that as N, on any bounded time interval, sample paths of (u(N)1(t),u(N)2(t)) converge in probability to solutions of the following pair of equations, which we take as our object of study:

    u1=λ10u0u1u1+δu2λ21u1u2u2=λ20u0u2u2δu2+λ21u1u2. (2.1)

    We note of course that the asymptotic behaviour of these equations is, in general, not the same as the above stochastic process. That said, for such models the connections often run deeper than just finite-time convergence; see for example the work of [13].

    As it turns out, it is more productive to study the total proportion of hosts: x1=u1+u2 and the proportion of hosts that are associated: x2=u2/x1. Letting λa=λ20λ10 and λb=λa+λ21, after a bit of algebra we obtain the system

    x1=G1(x1,x2)=x1((λ10+λax2)(1x1)1)x2=G2(x1,x2)=x2((λa+λbx1)(1x2)δ). (2.2)

    The coordinate change ux is singular at x1=0 but note the inverse u1=x1(1x2), u2=x1x2 is smooth and onto. Thus, every solution of (2.1) is the image of a solution of (2.2), so to understand (2.1) it is enough to study (2.2). In terms of the original description, the host population survives if lim inftx1(t)>0 and the symbiont survives if lim inftx2(t)>0. The symbiont takes over if limtx2(t)=1. It should be clear that takeover is only possible if δ=0, and even then only if further conditions are satisfied.

    From the form of (2.2) it should be clear that (x1,x2) is a more natural choice of variables than (u1,u2), since the right-hand sides in (2.2) both factor nicely into density dependent form, with a good deal of symmetry between the two equations. This form is the previously mentioned "stacked" perspective, wherein the total host density, and the prevalence of the symbiont within the host population, are thought to be the best descriptors of the population demographics. The sign of λa separates the pathogen, neutral and mutualist cases by λa<0,=0,>0 respectively. The sign of λb determines whether or not a higher host density is beneficial to the spread of the symbiont, measured by the proportion of hosts that are associated – since it depends on λ21, we say the symbiont is weakly infectious if λb<0, neutrally infectious if λb=0 and strongly infectious if λb>0. Distinguishing various cases according to the sign of λa and λb is crucial to making a complete analysis.

    We first identify the regions of interest. Define the feasible region Λ=[0,1]2 and let

    Λ+={(0,1]×(0,1)={(x1,x2)Λ:x1>0, 0<x2<1}if δ=0,(0,1]2={(x1,x2)Λ:x1,x2>0}if δ>0.

    In addition, let

    p0=(0,0),p1=(a1,0)=(11/λ10,0),p2=(a2,1)=(11/λ20,1) andp3=(0,a3)=(0,1δ/λa).

    Except for some corner cases, these are the only candidates for equilibria on the boundary of Λ.

    Define λM=max(λ10,λ20) which is an upper bound on the host birth rate. Also define the conditions

    (AinvU):λ20(1a1)+λ21a1>1+δ and(UinvA):λ10(1a2)λ21a2>1.

    (AinvU) stands for "associated invades unassociated", and is relevant iff λ10>1, in which case it corresponds to parameter values for which a small introduction of associated hosts in a stable population of unassociated hosts leads to an increase in the proportion of associated hosts. Equivalently, G2(p1+ϵe2)>0 for small ϵ>0, where e2=(0,1). (UinvA) is analogous, and is relevant iff λ20>1 and δ=0.

    The following is the main result of this article. Since the classification is somewhat lengthy we begin with a brief summary. There are four main cases: (E), (UH), (AH), (C), and two exceptional cases: (RS) and (B). In every case except for redundant symbiont (RS), every trajectory in Λ+, the interior of the feasible region, tends to one of at most three equilibria, and tends to a unique equilibrium when bistability (B) does not occur. When (RS) and (B) do not occur, the system settles into one of the following four states:

    ● extinction of both host and symbiont (E),

    ● survival of host and extinction of symbiont (UH),

    ● survival of host and complete adoption of the symbiont (AH), which is only possible if δ=0, and

    ● survival of host, with stable coexistence of associated and unassociated hosts (C).

    Case (RS) is defined precisely as λ10=λ20>1, δ=λ21=0 and would otherwise correspond to (UH). Case (B) is defined later, under section heading Subcase 5b, in terms of intersections of the nullclines. It falls strictly within the range of parameter values satisfying λa>0 and λb>0, that is a strongly infectious, mutualist symbiont, and would otherwise correspond to either (E) or (UH).

    Theorem 2.1. The following six cases include all parameter values.

    We begin with two exceptional cases.

    1. Redundant symbiont (RS). Suppose λ10=λ20>1 and δ=λa=λ21=0, which would otherwise correspond to (UH). For all xΛ+, limtϕ1(t,x)=a1 and tϕ2(t,x) is constant.

    2. Bistability (B). This occurs for a non-empty set of parameter values satisfying min(λa,λb)>0, δ>0, that would otherwise correspond to (E) if λ101, or to (UH) if λ10>1. Moreover, the set of values (δ,λ21) that gives (B) is non-empty iff λ20>max(1,λ210).

    There are three subcases: onset (OB), marginal (MB) and non-marginal (NB) bistability. (NB) is an open, contractible set of parameter values and (OB), (MB) are piecewise smooth and on the boundary of (NB). In each subcase there is an interior equilibrium ˉxΛ+ satisfying max(0,a1)<ˉx1<a2.

    i) Onset of bistability (OB). limtϕ(t,x)=ˉx for xΛ+.

    ii) Marginal bistability (MB). There is a set UΛ+, open in Λ+ and not containing ˉx, such that

    limtϕ(t,x)=(max(0,a1),0)forxU,andlimtϕ(t,x)=ˉxforxΛ+U.

    iii) Non-marginal bistability (NB). There are ˉxU2, ˉyΛ(U1U2), with max(0,a1)<ˉy1<ˉx1<a2 and max(0,a3)<ˉy2<ˉx2<1, and disjoint sets U1,U2Λ+ both open in Λ+, with Λ+(U1U2) the stable manifold of ˉy, such that

    limtϕ(t,x)=(max(0,a1),0)forxU1,limtϕ(t,x)=ˉxforxU2, andlimtϕ(t,x)=ˉyforxΛ+(U1U2).
    Figure 1.  Diagram indicating for which values of λ10,λ20 there exist δ,λ21 such that (B) (on the left) or (E), (C), (UH) and (AH) (on the right) can occur.

    Suppose (RS) and (B) do not hold. Then, there exists ˉxΛ such that limtϕ(t,x)=ˉx for all xΛ+.

    Assuming (RS) and (B) do not hold, four cases are possible.

    1. Extinction. (E) ˉx=(0,max(0,a3)), if λ101 and λ201+δ.

    2. Survival and coexistence of associated and unassociated host. (C)

    max(0,min(a1,a2))<ˉx1<max(a1,a2)andmax(0,a3)<ˉx2<1,if

    (a) δ>0 and either

    i. λ101 and λ20>1+δ, or

    ii. λ10>1 and (AinvU) holds, or

    (b) δ=0 and either

    i. λ101, λ20>1 and (UinvA) holds,

    ii. λ10>1, λ201 and (AinvU) holds, or

    iii. min(λ10,λ20)>1, (AinvU) holds and (UinvA) holds.

    3. Survival of unassociated host only. (UH) ˉx=p1 if λ10>1 and (AinvU) does not hold.

    4. Survival of associated host only. (AH) ˉx=p2 if δ=0, λ20>1 and (UinvA) does not hold.

    It is worth noting the behaviour under (OB) is the same as in (C). The reason why we do not include it in (C) is because it occurs for parameter values that would otherwise satisfy either (E) or (UH).

    In Figure 2 we show streamlines together with nullclines in a few different cases.

    Figure 2.  Streamlines in the (x1,x2) plane for λ10=0.5,λ20=4,λ21=12 and δ=2,4 and 6 from left to right, corresponding to coexistence, bistability subcase (NB) and extinction, respectively. x1 nullcline in blue, x2 nullcline in red.

    Theorem 2.1 gives a complete picture of the dynamics, with the possible objections that

    ● the exact parameter ranges for (OB), (MB), and (NB) are not specified, and

    ● a precise formula for interior equilibria is not given.

    In the case of (MB) and interior equilibria this is because they are determined by intersection conditions between curved nullclines on the interior of Λ. Despite this, in Subcase 5b we give a fairly nice formula for (OB), which leads to the iff condition λ20>max(1,λ210) for the possibility of bistability given in Theorem 2.1. We also give a satisfying characterization of the values of (δ,λ21) such that (NB) occurs.

    Before moving onto the proof we unpack somewhat the result of Theorem 2.1, tying it to the original motivation. We begin with (B) since it is the most unusual case. Ignoring (OB) and (MB) since they are boundary cases, when (NB) occurs there are two stable equilibria; the first gives either extinction of the host (if λ101) or survival without symbiont (if λ10>1), while the second is a coexistence equilibrium at a higher host density. The case λ101 is obligate mutualism, as for example with termites and their protozoa [14], since neither the host nor the symbiont can survive without the other; this has been previously observed mathematically in, for example, [15]. The case λ10>1 is facultative mutualism, whereby the host is able to survive without the symbiont, but still benefits from its introduction. We note that, in order for (B) to occur, not only must the symbiont be a mutualist, but because of the condition λ20>max(1,λ210), the benefit it imparts must be sufficiently high, and quite significant if λ10 is fairly large. On the other hand, since λb=λ21+λa, its tendency λ21 to spread through the host population must also exceed the benefit λa=λ20λ10 that it imparts to the host birth rate. Thus bistability can only occur when we have a very helpful, yet selfish mutualist in the sense that the significant benefit it provides is still exceeded by its rate of spread through the host population. The latter trait can also be viewed as a strategy for survival which is more focused on horizontal transmission than on vertical transmission.

    When we have either a pathogen (λa<0), or a mutualist that does more to aid host survival than to spread itself through the host population (λb<0), (B) does not occur. In this case there is a single equilibrium attracting all solutions in the interior of the feasible region, which is determined by the location of boundary equilibria (survival of unassociated hosts or associated hosts, in isolation) together with the invasion conditions (AinvU) or (UinvA), and whether or not the recovery rate is zero. In particular, it is straightforward to determine whether we have host extinction (E), survival of unassociated host only (UH), coexistence (C), or survival of associated host only (AH), so we leave it to the reader to inspect each case. This description, of course, bears the exception of the trivial case (RS) that essentially corresponds to two equally viable species that don't interact except in their competition for space.

    The remainder of the paper is organized as follows. In Section 3 we recall some basic theory of plane systems of differential equations. In Section 4 we take care of the dynamics on ΛΛ+, which consists of invariant lines on the boundary of the feasible region. In Section 5 we study the dynamics on the interior of Λ, in several steps. We first study special parameter values where (2.2) separates into a simpler form. We then determine boundary equilibria and their stability. Next we study the nullclines, and move on to study the itinerary of trajectories on regions delineated by the nullclines. We then give a sufficient condition to rule out periodic orbits, homoclinic orbits and heteroclinic cycles, using Dulac's criterion. At this point, we have the tools to study the dynamics in the remaining cases where (2.2) does not assume a simple form, and we complete the proof.

    We first recall some basic theory that can be found in [16], specialized to the present context. Let FC1(R2,R2). Corresponding to F there exists a C1 function ϕ:SR2 called the flow, defined on an open SR×R2 containing {0}×R2 and satisfying

    ϕ(0,x)=xandtϕ(t,x)=F(ϕ(t,x))  for  (t,x)S.

    In particular, for fixed x, tϕ(t,x) solves the initial value problem

    y=F(y),y(0)=x

    in an open time interval around 0. Letting τ+(x)=sup{t:(t,x)S} and τ(x)=inf{t:(t,x)S}, if we require that for every xR2,

    1. τ(x)= or limtτ(x)ϕ(t,x)= and that

    2. τ+(x)= or limtτ+(x)ϕ(t,x)=,

    then ϕ exists and is the unique function with the stated properties. In particular, for every xR2 the above initial value problem has a unique solution forward and backward in time, to either ±t= or until it diverges. Given xR2, the corresponding trajectory Γ is defined as

    Γ(x)={ϕ(t,x):τ(x)<t<τ+(x)},

    and the sets {Γ(x):xR2} form a partition of R2 into trajectories. We are also interested in the positive semi-trajectory

    Γ+(x)={ϕ(t,x):0t<τ+(x)}.

    We say that a set ER2 is invariant if xE implies Γ(x)E, and forward invariant if xE implies Γ+(x)E. If E is both bounded and forward invariant, the above implies τ+(x)= for every xE. In this case we are interested in the omega-limit set of points xE, defined by

    ω(x)={y:y=limnϕ(tn,x) for some sequence (tn) with limntn=}

    and contained in the closure of E. Omega-limit sets are known to be closed, connected, invariant, and made up of trajectories, which can be of a few types, including the following. An equilibrium point is a point x such that F(x)=0. A periodic orbit is the positive semi-trajectory of a point x satisfying ϕ(T,x)=x for some T>0, and a separatrix cycle is the union of finitely many equilibrium points p1,,pm for some m1 and trajectories Γ(xi),i=1,,m satisfying

    limtϕ(t,xi)=pifori=1,,m,limtϕ(t,xi)=pi+1fori=1,,m1  andlimtϕ(t,xm)=p1. (3.1)

    A separatrix cycle with m=1 is a homoclinic orbit, and with m>1 is a heteroclinic cycle. It should be clear from the definition that if ω(x) is a single point, then limtϕ(t,x)=ω(x).

    We return to (2.2) and begin by studying the dynamics on invariant boundary lines.

    Dynamics on ΛΛ+. Let L1={x2=0}, L2={x1=0} and L3={x2=1}, so that

    Λ+={Λ(L1L2)ifδ>0Λ(L1L2L3)ifδ=0.

    The lines L1,L2 are invariant, as is L3 if δ=0. On L1 the equilibria are p0 and p1. Since G1=x1(λ10(1x1)1) if x2=0, the segment {0<x11, x2=0} is forward invariant and for x in that segment,

    limtϕ(t,x)=(max(0,a1),0).

    On L2 the equilibria are p0 and p3. Since G2=x2(λa(1x2)δ) if x1=0, the segment {x1=0, 0<x21} is forward invariant and for x in that segment,

    limtϕ(t,x)=(0,max(0,a3)).

    On L3, G2=δx2, so if δ=0 that line is invariant and the equilibria are p3 and p2. Since G1=x1(λ20(1x1)1) if x2=1, the segment {0<x11, x2=1} is forward invariant and for x in that segment,

    limtϕ(t,x)=(max(0,a2),1).

    Forward invariance of Λ,Λ+. Note that on the line {x1=1}, G1=1<0, and if δ>0 then G2<0 on {x1>0, x2=1}. Therefore Λ is forward invariant. This is because every side of [0,1]2 is either invariant, or else is such that for x on that side, ϕ(t,x)(0,1)2 for small enough t>0, which means that trajectories cannot exit [0,1]2 along any of its four sides. The same is true if we remove any number of invariant lines from Λ. In particular, Λ+ is also forward invariant.

    Next we determine ω(x) for xΛ+, proving Theorem 2.1. By forward invariance ω(x)Cl(Λ+)=Λ, where Cl denotes the closure. The analysis is broken up into five broad cases. In each situation we point out which case in Theorem 2.1 ((RS), (OB), (MB), (NB), (E), (C), (UH), (AH)) applies.

    Case 1 – low host birth rate: λM=max{λ10,λ20}1 (E). If xΛ then G1x1(λM(1x1)1).

    So, λM1 implies G1x21 and limtϕ1(t,x)=0.

    Case 2 – neutral symbiont: λM>1, λa=0. In this case λ10=λ20>1 and λb=λ21, so

    G1=x1(λ10(1x1)1)andG2=x2(λ21x1(1x2)δ).

    Since λ10,λ20>1, a1,a2>0. Moreover if xΛ then limtϕ1(t,x)=a1. Since λ10=λ20, we have λ20(1a1)=λ10(1a2)=1, so (AinvU)λ21a1>δ and (UinvA)λ21a2>0λ21>0.

    If δ=λ21=0 (RS) then G2=0, so tϕ2(t,x) is constant.

    If δ=0 and λ21>0 (AH) then G2>0 for x1>0 and 0<x2<1, so limtϕ2(t,x)=1 for xΛ+.

    If δ>0 and λ21=0 (UH) then G2<0 for x2>0, so limtϕ2(t,x)=0 for xΛ+.

    If δ,λ21>0, then using the limiting value of ϕ1 we find that

    limtϕ2(t,x)=ˉx2=max(0,1δλ21a1) for xΛ+.

    If λ21a1δ (UH) then ˉx2=0 and if λ21a1>δ (C) then 0<ˉx2<1.

    Case 3 – neutrally infectious symbiont: λM>1, λa0, λb=0. In this case λ10λ20 and λ10=λ20λ21. This forces λ20>λ10 so that λa>0 and λ21>0. Multiplying on both sides by λ10,

    (AinvU) λ20+λ21(λ101)>(1+δ)λ10λ21λ10>δλ10

    which is λ10>0 and λ21>δ. Multiplying on both sides by λ20,

    (UinvA) λ10λ21(λ201)>λ20λ21λ20>0,

    so (UinvA) does not hold. We find

    G1=x1((λ10+λax2)(1x1)1)andG2=x2(λa(1x2)δ),

    so

    limtϕ2(t,x)=ˉx2=max(0,1δ/λa) for xΛ+.

    Then,

    ifδ=0thenˉx2=1,(AH)if0<δ<λathen0<ˉx2<1(C) or (E)and ifδλathenˉx2=0(UH) or (E).

    Plugging ˉx2 into the equation for G1 we find

    limtϕ1(t,x)=max(0,11/(λ10+λaˉx2)) for xΛ+.

    In particular, ϕ10 exactly on the condition λ10+λaˉx21, which

    ifδ=0isλ201 thus empty,if0<δ<λaisλ201+δ(E)and ifδλaisλ101(E).

    This concludes the easy cases. To treat cases with λa,λb0, more effort is required. We will start by studying equilibria on the boundary of Λ+, then nullclines on Λ+, then itineraries on Λ+. Then, we will be ready to consider the remaining cases.

    Equilibria on L. Let L=L1L2 if δ>0 and L=L1L2L3 if δ=0, so that Λ+=ΛL. Since λa,λb0, p0,p1,p2,p3 are the possible equilibria (eq) on L. Letting a1=11/λ10, a2=11/λ20, a3=1δ/λa,

    1. p0=(0,0)L1L2 is always an eq in Λ.

    2. p1=(a1,0)L1 is an eq in Λ{p0} if λ10>1.

    3. p3=(0,a3)L2 is an eq in Λ{p0} if δ<λa, and if δ=0 then p3=(0,1)L2L3.

    4. p2=(a2,1)L3 is an eq in Λ{p3} if δ=0 and λ20>1.

    To assess stability we compute the Jacobian matrix J(x)=(jGi(x))ij.

    J(x)=((λ10+λax2)(12x1)1λax1(1x1)λbx2(1x2)(λa+λbx1)(12x2)δ).

    If xL then at least one of x1,x2 is equal to 0 or 1, and J(x) is triangular, so its eigenvalues are equal to its diagonal entries. Straightforward computation shows that

    1. J(p0) has eigenvalues λ101 and λaδ,

    2. J(p1) has eigenvalues 1λ10 and J22(p1),

    3. J(p2) has eigenvalues 1λ20 and J22(p2) if δ=0, and

    4. J(p3) has eigenvalues λ20(1+δ) and δλa.

    Since λb=λa+λ21, λa=λ20λ10, λ10(1a1)=1 and 1a1=1/λ10,

    J22(p1)=λa+λba1δ=λa(1a1)+λ21a1δ=λ20(1a1)+λ21a1(1+δ)=λ20/λ10+λ21(11/λ10)(1+δ),

    and if δ=0 then since λa=λ10λ20 and λ20(1a2)=1,

    J22(p2)=λaλba2=λa(1a2)λ21a2=λ10(1a2)λ21a21=λ10/λ20λ21(11/λ20)1.

    Using these calculations and the fact that G2(x+ϵe2)=ϵG2(x)+o(ϵ),

    ⅰ) J22(p1)>0 (AinvU) G2>0 just above p1.

    ⅱ) J22(p2)>0 (UinvA) G2<0 just below p2.

    Recall the following three types of equilibria p:

    ⅰ) Attracting: for some ϵ>0, |xp|<ϵ implies limtϕ(t,x)=p.

    ⅱ) Saddle point: the stable manifold and unstable manifold

    Ws(p)={x:limtϕ(t,x)=p} and Wu(p)={x:limtϕ(t,x)=p}

    are smooth curves each consisting of p and a pair of trajectories, and the tangent lines to Ws,Wu at p are eigenvectors of J(p) with eigenvalues whose real part is negative, respectively positive.

    ⅲ) Repelling: for some ϵ>0, |xp|<ϵ implies limtϕ(t,x)=p.

    Given equilibrium p, let μ1,μ2 be the eigenvalues of J(p), and recall that [hartman-grobman]

    ⅰ) p is attracting if (μ1), (μ2)<0,

    ⅱ) p is a saddle point if (μ1)<0<(μ2), and

    ⅲ) p is repelling if 0<(μ1), (μ2).

    If one or both of (μi)=0 then p could be of any type, or none.

    Nullclines.

    {G1=0}={x1=0}γ1whereγ1={(λ10+λax2)(1x1)=1}, and{G2=0}={x2=0}γ2whereγ2={(λa+λbx1)(1x2)=δ}.

    γ1 is a hyperbola with vertical asymptote x1=1, so instead let γ1 be the branch with x1<1.

    If δ=0 then γ2={x2=1}{x1=λa/λb} so instead let γ2 be the vertical line {x1=λa/λb}.

    If δ>0 then γ2 is a hyperbola with horizontal asymptote x2=1, so instead let γ2 be the branch with x2<1. Both γ1 and γ2 intersect each of the invariant lines at most once. The intersection points, and conditions for existence, are

    γ1{x2=0}=p1 if λ10>0,

    γ1{x1=0}=r1=(0,b1) with b1=(1λ10)/λa,

    γ1{x2=1}=p2 if λ20>0,

    γ2{x2=0}=r2=(b2,0) with b2=(δλa)/λb,

    γ2{x1=0}=p3, and

    γ2{x2=1}=r3=(b2,1) if δ=0.

    By considering what happens when x2=0 and noting that {x1=0},γ1 are the only null sets for G1 when 0x11, it follows that G1>0 to the left of γ1 in Λ+. On the other hand, since x1G2=λbx2(1x2), if λb<0 then G2>0 to the left of γ2, while if λb>0 then G2>0 to the right of γ2, in Λ+.

    If λ10>1 then b2=a1 marks the boundary between (AinvU) and its complement.

    Similarly, if λ20>1 and δ=0 then b2=a2 marks the boundary between (UinvA) and its complement.

    Also, it's easy to check that b1=a3 is equivalent to λ20=1+δ.

    It's useful to think of γ1 as the graph of the function

    x2=g1(x1), x1<1whereg1(x1)=λ1a(11x1λ10),

    and when δ>0 to think of γ2 as the graph of the function

    x2=g2(x1), x2<1whereg2(x1)=1δλa+λbx1.

    If δ=0 then γ1γ2 has cardinality 0 or 1 as γ1 has at most one intersection point with any vertical line. If δ>0 then since the sign of the pairs (g1(x1),g1(x1)) and (g2(x1),g2(x1)) do not depend on x1 and never coincide (to see this, break up according to (sgn(λa),sgn(λb)), noting that λa+λb=λ210 implies (sgn(λa),sgn(λb))(1,1)), γ1γ2 has cardinality 0,1 or 2.

    Itineraries. Let G=(G1,G2) and for a pair (a,b) let sgn(a,b)=(sgn(a),sgn(b)){1,0,1}2. For each sign vector σ{1,0,1}2 let

    R(σ)={xΛ+:sgn(G(x))=σ}.

    Since all signs are represented, {R(σ):σ{1,0,1}2} is a partition of Λ+. Note that one or more of these sets could be empty.

    ● The set R(0,0)=γ1γ2Λ+ consists of equilibrium points and has cardinality 0,1 or 2.

    ● The sets R(0,1)R(0,1)=γ1Λ+γ2 and R(1,0)R(1,0)=γ2Λ+γ1, so we refer to the union of all four as Rγ. Since neither hyperbola has either a horizontal or vertical tangent at any point, while G is horizontal or vertical on each one (or if δ=0, γ2 is vertical while G is horizontal on γ2), it follows that on Rγ, G is transverse to the tangent space of Rγ, so for any xΛ+, the set of times {t:ϕ(t,x)Rγ} is a discrete set.

    ● The sets R(σ) for σ{1,1}2 are open subsets of (0,1]2 that, based on the number of intersection points of γ1 and γ2, each have at most 2 connected components.

    If x(0,1]2R(0,0) then since R(0,0) is invariant and {t:ϕ(t,x)Rγ} is discrete, there is a (possibly finite) symbolic sequence I(x)=(σ1,σ2,) called the itinerary of x, with each σi{1,1}2, and an increasing sequence of times t(x)=(t1,t2,), with t0=0, and with ti= iff I(x) has length i, with the property that

    ϕ(t,x)R(σi) for t(ti1,ti).

    Define a directed graph called the sign graph with vertices {1,1}2 by including each directed edge (σ,σ) iff there is x such that (Ii(x),Ii+1(x))=(σ,σ) for some i1. The sign graph is cyclic if there is a sequence of edges (σ1,σ2),,(σk,σ1), and acyclic if not. Appearance of (σ,σ) in the sign graph is also represented by R(σ)R(σ). R(σ) is absorbing if none of the edges (σ,) appear on the sign graph. An equilibrium point pCl(R(σ)) is a sink for R(σ) if for some xR(σ), I(x)=σ and limtϕ(t,x)=p. If p is a sink for R(σ) then it is a local maximum of xσx on R(σ). The following lets us seal the fate of trajectories, when the sign graph is acyclic. The proof is given in the Appendix.

    Lemma 5.1. If I(x) is finite and its last entry is σ then ω(x) is a sink for R(σ).

    In particular, if the sign graph is acyclic then for all xΛ+, ω(x) is a sink.

    Periodic orbits and cycles. Lemma 1 makes our life easier when the sign graph is acyclic. When the sign graph is cyclic, we will use the following proposition to simplify the analysis.

    Proposition 5.1. For each xΛ+, ω(x) is either

    1. an equilibrium point or

    2. contains a separatrix cycle intersecting an equilibrium in L.

    Proof. There are three steps.

    1. Show there are no periodic orbits in Λ and no separatrix cycles in Λ+.

    2. Show that a separatrix cycle intersecting L must intersect an equilibrium in L.

    3. Invoke a generalized form of the Poincaré-Bendixson theorem.

    Step 1. We begin with

    Lemma 5.2 (Dulac's criterion). - Suppose we have the planar system

    u=F(u)whereuR2andF:R2R2 isaC1function.

    Let RR2 be a simply connected region in the plane and B:R2R2 be a C1 function (called a Dulac function) such that the divergence of BF is non-zero and has constant sign almost everywhere on R. Then, the system u=F(u) has no periodic orbit and no separatrix cycle on R.

    Since most proofs of Dulac's criterion only rule out periodic orbits, we include a proof in the appendix.

    No periodic orbit or separatrix cycle in Λ+. Since the restriction of x(x1(1x2),x1x2) to {0<x11, 0<x2<1} maps bijectively onto {u1,u2>0, u1+u21}, periodic orbits and separatrix cycles are mapped to the same. So, to rule them out on the former set, we find a Dulac function for (2.1) on the latter set. The function B(u)=1/(u1u2) works, since

    (BF)=u1(λ10u0/u21/u2λ21+δ/u1)+ u2(λ20u0/u11/u1+λ21δ/u1)=λ10/u2δ/u21λ20/u1

    is strictly negative when u1,u2>0 (at least one of λ10,λ20 is positive).

    If δ=0 the set {0<x11, 0<x2<1} is equal to Λ+.

    If δ>0, Λ+ includes also the segment {0<x11, x2=1}. Since G2<0 on that segment, any trajectory that intersects it also intersects Λc, so is not contained in Λ.

    No periodic orbit in L. If a periodic orbit C intersects L=ΛΛ+, it intersects an invariant line on L. Since C is a single trajectory, it is a subset of that invariant line. Since C is the continuous injective image of a circle and L is either a pair of orthogonal lines, or two parallel lines and an orthogonal line, this cannot occur.

    Step 2. A separatrix cycle is the continuous injective image of a circle. If it intersects L, it intersects an invariant line on L, so it must contain an equilibrium point on that line. If it did not, it would consist of a single trajectory on that line.

    Step 3. This step is a trivial application of the following result.

    Theorem 5.1 (Generalized Poincaré-Bendixson Theorem). - Let ER2 be an open set and FC1(E,R2). Suppose the system x=F(x) has a positive semi-trajectory Γ+(x) contained in a compact set KE with the property that F(x)=0 for at most finitely many xK. Then, ω(x) is either

    1. an equilibrium point,

    2. a periodic orbit, or

    3. contains a separatrix cycle.

    Proof. This is given by Theorem 2 in Section 3.7 of [17], and the comments that follow.

    This concludes the proof of Proposition 5.1.

    Trajectories. We are now ready to study dynamics in the remaining cases. We break up our analysis according to the sign of λa,λb. The interested reader will note that Case 3 above can also be covered here using the same methods, although it does not give an explicit formula for the interior equilibrium, when there is one.

    Case 4 – pathogen: λM>1, λa<0, λb0. In this case, λ10>max(1,λ20) and λb=λa+λ21>0, so a1=11/λ10>0, b2=(δλa)/λb>0 and G2>0 to the right of γ2. g1 is decreasing and concave, and if δ>0 then g2 is increasing and concave, while if δ=0 then γ2={x1=b2} – in either case γ2 lies to the right of {x1<b2} so G2<0 on this set. Since G1>0 near p0 it attracts no points in Λ+, and since p3=(0,a3) with a3=1δ/λa1, if p3Λ then δ=0, p3=(0,1) and G2<0 in a Λ+-neighbourhood of p3, which means p3 attracts no points in Λ+.

    If b2a1 then (AinvU) does not hold, which gives (UH). R(1,1)R(1,1)R(1,1) appear from right to left. If p2Λ then since λ20<λ10, a2<a1b2 so G2<0 in a Λ+-neighbourhood of p2. Since the sign graph is acyclic, every solution tends to a sink, and the only candidate is p1.

    If δ>0 and b2<a1 (AinvU) which gives (a.ii) of (C), or δ=0 and a2<b2<a1 which gives (AinvU) and (UinvA) and thus (b.iii) of (C) if a2>0 or (b.ii) of (C) if a20, then (C) holds. γ1γ2 is a unique point p4Λ+ and p1 is a saddle, as is p2 if δ=0. The sign graph is cyclic, so we use Proposition 1 to conclude that solutions tend to p4 provided no separatrix cycles on Λ contain an equilibrium on L – let C denote a possible candidate. Since p1 is a saddle, Ws(p1)={0<x1, x2=0} so p1C implies p0C. Since p0 is a saddle with Ws(p0)={x1=0, x2<a3=1δ/λa}, p0C implies p3C. If δ>0, p3Λ so C cannot exist, otherwise δ=0 and p0C implies p3=(0,1)C. If λ20<1 then Ws(p3)={a2<x1, x2=1} and C cannot exist. If λ20=1 then since γ2 does not intersect p3, still G2<0 in a Λ+-neighbourhood of p3, so Ws(p3) is disjoint from Λ+ which means Ws(p3)={0x1, x2=1} and C cannot exist. If λ20>1 then p3 is repelling, and C cannot exist. If δ>0 then p2Λ, and if δ=0 and p2Λ then p2 is a saddle with Ws(p2)={0<x1, x2=1}, so p2C implies p3C. Since p3 is repelling, C cannot exist.

    If δ=0 and b2a2<a1 then (AinvU) holds but (UinvA) does not, which gives (AH). R(1,1)R(1,1)R(1,1) appear from left to right, and p1 is a saddle. Since the sign graph is acyclic, solutions tend to a sink, and the only candidate is p2.

    Case 5 – mutualist: λM>1, λa>0, λb0. In this case λ20>max(1,λ10) so a2>0, and g1 is increasing and convex. Notice that g1 intersects the set {x1=0,0x2<1}{0x1<a2, x2=0} at exactly one point, which is p1 if λ10>1, p0 if λ10=1 and r1=(0,b1)=(0,(1λ10)/λa) if λ10<1.

    Subcase 5a – weakly infectious mutualist: λb<0. If δ>0 then g2 is decreasing and concave, and recall γ2{x2=0} contains only the single point (b2,0) with b2=(δλa)/λb, and γ2{x1=0} contains only the single point p3=(0,a3) with a3=1δ/λa. γ1γ2Λ+ contains exactly one point if either 0<a1<b2 or a10 and a3>b1, both of which correspond to (C), and otherwise is empty. Notice that

    a3>b1(λaδ)/λa>(1λ10)/λaλaδ>1λ10λ20>1+δ.

    If δ=0 then b2=λa/λb>0 and p3=(0,1) is a saddle, so attracts no points in Λ+. If a1<b2<a2, which corresponds to (C), then γ1γ2Λ+ contains exactly one point, otherwise γ1γ2Λ+ is empty.

    If δ>0 and γ1γ2Λ+ is empty then either a1>0 and a1b2 (UH), or a10 and a3b1 which implies λ101 and λ201+δ (E). There are at most three sign regions in Λ (two if b20): R(1,1)R(1,1)R(1,1), appearing in that order from left to right. In case of (UH), p0 is a saddle and p1Λ, and in case of (E), p0 is the only equilibrium in Λ.

    If γ1γ2Λ+ is not empty then it is a unique point p4Λ+. If a10 then p0 is a saddle, otherwise p0 is repelling and p1 is a saddle, and if δ=0 then p2,p3 are saddles. A similar argument as before shows there is no periodic orbit or separatrix cycle in Λ. Thus, all solutions tend to p4.

    If δ=0 and 0<b2a1<a2 (UH), R(1,1)R(1,1)R(1,1) appear from left to right. Since a1>0, p0 is a saddle and p1Λ, while p2,p3 are saddles, so solutions tend to p1.

    If δ=0 and a1<a2b2 (AH), then R(1,1)R(1,1)R(1,1) appear from right to left, p0 is repelling, and p1,p3 are saddles, so solutions tend to p2.

    Subcase 5b – strongly infectious mutualist: λb>0. In this case (B) can occur, so we take a moment to define precisely the three conditions in terms of the nullclines:

    1. (OB) occurs when g1,g2 intersect at two points in Λ, one in Λ+ and the other on the subset {x1=0,0x21}{0x11,x2=0} of the boundary.

    2. (MB) occurs when g1,g2 intersect tangentially at one point in Λ+.

    3. (NB) occurs when g1,g2 intersect at two points, both in Λ+.

    As we go through the various cases we will point out when one of the above might occur; in particular, we'll see that they all occur for values that would otherwise correspond to (E) or (UH). We'll also describe the dynamics in all possible cases. We'll then show that each of (OB), (NB), (MB) does occur, make some effort to characterize them, and show (NB) is simply connected.

    If δ=0 then G2=x2((λa+λbx1)(1x2)>0 on Λ+, so (AH) holds. b1=λa/λb<0 and R(1,1)R(1,1) appear from right to left. Because of the sign of G2, p0 (and p1 if a1>0) attract no points, and p3=(0,1) is a saddle. Since the sign graph is acyclic, solutions tend to a sink, and the only candidate is p2.

    For the rest we assume δ>0. g2 is increasing and concave, so γ2 intersects the set {x1=0,0x2<1}{x10,x2=0} at exactly one point, which is p3=(0,a3)=(0,1δ/λa) if δ<λa, p0=(0,0) if δ=λa and r2=(b2,0)=((δλa)/λb,0) if δ>λa.

    If either a1>max(b2,0), or a10 and a3>max(b1,0), then γ1γ2Λ+ is a unique point p4Λ+. A similar argument as before shows that all solutions tend to p4. The first option gives λ10>1 and (AinvU) so corresponds to (C). For the second option, note that a10λ101 and

    a3>b1λaδ>1λ10λ20>1+δ,

    so it also corresponds to (C).

    If b2=a1>0 (UH), then γ1,γ2 intersect at p1. If g1(a1)g2(a1), then γ1γ2Λ+ is empty, and the sign regions R(1,1),R(1,1),R(1,1) appear from left to right with R(1,1),R(1,1)R(1,1), so the sign graph is acyclic. Since b2>0 and γ2 is increasing, a3<0 so p3Λ+ and G1>0 near p0 so solutions tend to p1. If g1(a1)<g2(a1), which as we'll see corresponds to (OB), then γ1γ2Λ+ is a single point p4. All four sign regions now appear, with R(1,1),R(1,1)R(1,1),R(1,1). To show all solutions tend to p4 it remains to show that p1 is not a sink for any sign region. p1 is in the closure of R(1,1),R(1,1) and R(1,1). Since G2>0 on R(1,1) p1 cannot be a sink, and p1 is a local minimum of xx(1,1) on R(1,1), whereas a sink is a local maximum. The only remaining candidate is R(1,1). G2 has a double root at p1, which can be seen from either the transverse intersection of two x2-nullclines ({x2=0} and γ2) or by writing G2 as a function of ˜x=xp1. Writing G in these coordinates,

    G1=(˜x1+a1)((λ10+λa˜x2)(1(˜x1+a1))1)=(˜x1+a1)(λ10+λa˜x2)(λ110˜x1)1)=(˜x1+a1)(λ10˜x1+λa˜x2(λ110˜x1))

    and

    G2=˜x2(λa+λb(˜x1+(δλa)/λb)(1˜x2)δ)=λb˜x1˜x2(1˜x2)δ˜x22

    If a trajectory in R(1,1) tends to p1 then ˜x10 and ˜x20+. If ˜x1<0<˜x2 are small in magnitude then by 1st-order approximation G1c1(˜x1+˜x2)>0 for some c1>0, and by 2nd-order approximation, G2c2˜x1˜x2c3˜x22<0 with c3,c4>0. Writing a solution to (2.2) as ˜x2=h(˜x1),

    h=ddt˜x2ddt˜x1=G2G1c2˜x1˜x2c3˜x22c1(˜x1+x2)=c2c1˜x2˜x1˜x1+˜x2c3c1˜x2˜x2˜x1+˜x2c2+c3c1˜x2,

    noting that 0<˜x1,x2˜x1+x2 on the last step. So, given initial value ˜x1(0),˜x2(0), the corresponding solution has h(˜x1)˜x1(0)ec˜x1 with c=(c2+c3)/c1, so intersects {˜x1=0} at a positive value of ˜x2. Thus no trajectories tend to p1.

    If a3=b10 (E), then γ1,γ2 intersect at p3. If g1(a3)g2(a3), then as above, γ1γ2Λ+ is empty and R(1,1),R(1,1)R(1,1). Since a10, if a3=0 then p0 is the only equilibrium, and if a3>0 then G2>0 near p0 and p1Λ+ so solutions tend to p3. If g1(a3)<g2(a3), then γ1γ2Λ+ is a single point p4, and R(1,1),R(1,1)R(1,1),R(1,1). A similar argument as above shows that p3 attracts no points in Λ+ – this time, G1 has a double root, so write a solution as ˜x1=h(˜x2) and perform a similar estimate. This situation corresponds to (OB).

    If 0<a1<b2 (UH) then γ1γ2Λ+ has cardinality 0,1 or 2, with a saddle node bifurcation when the cardinality is 1.

    1. If the cardinality is 0, then R(1,1),R(1,1)R(1,1) and solutions tend to p1, similar to the case g1(a1)g2(a1) above.

    2. If the cardinality is 1, then we have (MB). To see this, note that p1 is attracting, and there is a unique interior equilibrium p4. Three sign regions R(1,1),R(1,1),R(1,1) appear, with R(1,1),R(1,1)R(1,1) and R(1,1) disconnected by p4 into two components, the lower of which with sink p1 and the upper with sink p4. Since p1 is locally stable it attracts an open neighbourhood Sp1 of points in Λ+. Since the basin of attraction of p1 is t<0ϕ(t,S) it is open in Λ+.

    3. If the cardinality is 2, we have (NB). To see this, note that p1 is again attracting, and there are two interior equilibria p4,p5, letting p5 denote the upper one. All sign regions appear, with R(1,1),R(1,1)R(1,1),R(1,1) and R(1,1) split in two as before. p1 is the unique sink for the lower part of R(1,1), and p5 the unique sink for R(1,1) and the upper part of R(1,1). Since γ1,γ2 intersect transversally at p4 and p5 and are the only nullclines through those points, the Jacobian is non-singular so both equilibria are hyperbolic. Since p5 attracts more than a single curve of points it is attracting, and since p4 fails to attract nearby points in R(1,1) and R(1,1) it is not attracting. As argued above, both p1 and p5 have open basins of attraction in Λ+. Since Λ+{p4} is open in Λ+ and connected, it cannot be the combined basins of attraction of p1 and p5 since that would imply a disconnection. The only option is that some points other than p4 tend to p4, so p4 is a saddle and its basin of attraction is its stable manifold, which is a smooth curve.

    If a10 and 0<a3<b1 (E) then γ1γ2Λ+ has cardinality 0, 1 or 2 as before. Similar observations show that (E), (MB), (NB) occur, respectively.

    Properties of (B). We now show that each of (OB), (MB) and (NB) occurs for some parameter values, that (MB) and (OB) lie on the boundary of (NB), and that (NB) is simply connected. For this it is helpful to view the parameter space in terms of the four independent variables (λ10,λa,δ,λb), subject to the general constraints λ10,δ0, λaλ10 and λbλa. Here, of course, we may assume further that δ,λa,λb>0, since (B) has been ruled out in other cases. We'll let ˉx and ˉy,ˉx denote γ1γ2Λ when its cardinality is, respectively, 1 or 2, with ˉy<ˉx in the elementwise order.

    We recall g1,g2 and their derivatives:

    g1(x1)=1λa(11x1λ10),g2(x1)=1δλa+λbx1,g1(x1)=1λa1(1x1)2andg2(x1)=δλb(λa+λbx1)2.

    Existence. Our approach is to first characterize (OB), then show a small change in δ leads to (NB), then show a large enough change in δ gives (MB). We treat separately the cases a10 and a1>0, beginning with a10. To have (OB) we need g1(0)=g2(0), i.e., b1=a3, and g1(0)<g2(0). Earlier we showed that a3=b1λ20=1+δ, that we write as δ=λa+λ101, which requires λa+λ10>1, i.e., λ20>1, since δ>0 by assumption. If g1(0)<g2(0) then λ1a<δλb/λ2a, so λa<δλb, that we write as λb>λa/δ. Thus if λ101 and λa are fixed, (OB) occurs iff λa+λ10>1, λb(λa/(λa+λ101),) and δ=λa+λ101, which is a non-empty set of parameters. Take some such choice of parameters, and note that g1(0)g2(0)=0 and g1(0)g2(0)<0. Since g2 decreases with δ while g1 is unchanged, a small increase in δ then gives (NB). If δ is increased enough, then g2(1)<g1(0) and γ1γ2Λ+ is empty, so there is a unique intermediate value of δ that gives (MB). This is depicted in Figure 3.

    Figure 3.  Transition from (OB) to (NB) to (MB) with λ10=0.5, λa=1.5, λb=5 and δ=1,1.1,1.25 from left to right. g1 in blue, g2 in red.

    If a1>0, to have (OB) we need a1=b2>0 and g1(a1)<g2(b2). The former gives a1=(δλa)/λb, that we write as δ=λba1+λa noting a1 depends only on λ10, while the latter gives λ210/λa<δλb/δ2, or δ<λaλb/λ210. Plugging in δ and putting λb terms to one side gives the condition λb(λa/λ210a1)>λa. Given λ10 and λa, λb can be chosen large enough to satisfy this inequality iff

    λa/λ210>a1λa>a1λ210=λ210λ10λ20>λ210.

    Thus for a1>0, (OB) occurs iff λa>λ210λ10, λb(λa/(λa/λ210a1),) and δ=λba1+λa. Again, this is a non-empty set of parameters. As before, we can vary δ to obtain (NB) and (MB).

    A consequence of the above existence arguments is that they specify for which values of λ10 and λa bistability can be achieved by appropriate choice of δ,λb, namely, for λ20>max(1,λ210). They also show that for any choice of λ10,λa,λb, the set of values of δ that gives (B) is a bounded set, with (OB) at the lower endpoint, (MB) at the upper endpoint and (NB) in the interior. A slice of the phase portrait is shown in Figure 4.

    Figure 4.  Typical (δ,λb) slice of the phase portrait when λ101 (left) and λ10>1 (right) and bistability is possible. Left: λ10=0.5 and λa=1.5, and right: λ10=1.25 and λa=10.

    Geometry. It's clear that (NB) is an open set in parameter space. Define the functions ˜λa=λ10+max(1,λ210) and

    ˜λb={λa/(λa+λ101)if  λ101,λa/(λa/λ210(11/λ10))if  λ10>1.

    Note that both functions are continuous. From the above discussion we obtain functions δ(λ10,λa,λb) and δ+(λ10,λa,λb) with domain λ100, λa>˜λa and λb>˜λb such that (OB), (NB) and (MB) occur iff (λ10,λa,λb) are in the domain and δ=δ, δ<δ<δ+ and δ=δ+, respectively. The function δ is given above explicitly, piecewise:

    δ(λ10,λa,λb)={λa+λ101if  λ101λa+λb(11/λ10)if  λ10>1,

    and one easily verifies its continuity. The function δ+ is not given explicitly, but we can show it is smooth, using the implicit function theorem. (MB) corresponds to a unique ˉx1(0,1) such that g1(ˉx1)g2(ˉx1)=g1(ˉx1)g2(ˉx1)=0. We have

    (x1,δ)(g1g2,g1g2)=(g1g2δ(g1g2)g1g2δ(g1g2))

    where the indicates x1. At (MB), g1(ˉx1)g2(ˉx1)=0, moreover g1g2>0 since g1g2 is convex. Also, δ(g1g2)=1/(λa+λbx1)2>0 since λa>0 by assumption. Thus the above matrix is invertible. Since (g1g2,g1g2) is a smooth function of parameters, and of x1, when λa>0 and x1(0,1), by the implicit function theorem, (ˉx1,δ+) is a smooth function of (λ10,λa,λb), on the domain of δ+.

    We now show (NB) is a contractible set, via a sequence of homotopies that gives a deformation retraction. Define ˜δ=(δ+δ+)/2, which is continuous, and note that (NB) occurs when δ=˜δ. Notice that ˜λa,˜λb and ˜δ are functions of λ10, λ10 and λa, and λ10,λa and λb respectively. Define the sequence of homotopies as follows, at each step letting ˜λa,˜λb and ˜δ take the previous values in the list as their input:

    (ⅰ) H1=(λ10,λa,λb,(1t)δ+t˜δ),

    (ⅱ) H2=(λ10,λa,(1t)λb+t(˜λb+1),˜δ),

    (ⅲ) H3=(λ10,(1t)λa+t(˜λa+1),˜λb+1,˜δ),

    (ⅳ) H4=((1t)λ10+t,˜λa+1,˜λb+1,˜δ).

    Note that t[0,1] is the deformation parameter. In words,

    (ⅰ) H1 acts only on δ, centering its value to ˜δ,

    (ⅱ) H2 acts on λb and δ, centering λb to ˜λb+1 while keeping δ=˜δ,

    (ⅲ) H3 centers λa to ˜λa+1 while keeping λb=˜λb+1 and δ=˜δ and

    (ⅳ) H4 centers λ10 to 1 while keeping other parameters centered.

    We have already checked the continuity of component functions, and it's clear they keep points in (NB). Applying the Hi in sequence deforms (NB) to the single point with λ10=1 and the corresponding values of ˜λa,˜λb and ˜δ.

    In this article we consider a simple differential equation model of host-symbiont dynamics, that includes both vertical and horizontal transmission of the host, as well as recovery (i.e. transition from associated [with symbiont] to unassociated [without symbiont] host). The model naturally incorporates the pathogen, neutral, and mutualist cases through the choice of parameter values, providing a unified setting for the study of all three cases.

    Our analysis indicates that the most natural way to view the process is through the total host density (both associated and unassociated) and the proportion of hosts that are associated. Moreover, the two main factors governing the dynamics are (ⅰ) the effect of the symbiont on the host (i.e. whether it is pathogen, neutral or mutualist) and (ⅱ) the difference between the additional birth rate of associated vs. unassociated hosts (which is positive only for mutualists) and the rate of horizontal transmission of the symbiont.

    We obtain a complete picture of the model's dynamic phases, as follows. Except for the very special case when the symbiont is completely redundant (neutral, no recovery and no horizontal transmission), populations that initially contain both host and symbiont, and not located at an unstable equilibrium, always settle down to one of at most three equilibria. The limiting equilibrium is unique when the symbiont is either a pathogen, is neutral, or is a mutualist but focuses more on increasing the host birth rate than on horizontal transmission. Bistability (exactly two locally attracting equilibria) can occur when the symbiont is a sufficiently beneficial mutualist that focuses more on horizontal transmission than on increasing the host birth rate. There are three limiting equilibria only in a boundary case at the onset of bistability. When the limiting equilibrium is unique, survival and coexistence are determined by straightforward invasibility conditions, i.e., whether one type can invade the extinction equilibrium (for survival) or the stable equilibrium containing only the other type (for coexistence).

    Interestingly, the dynamical behaviour, and in particular the number of stable equilibria, hinges less on whether the symbiont is pathogen or mutualist, and more on whether the symbiont is highly beneficial (increases host birth rate), versus highly infectious (has a high rate of horizontal transmission). Moreover, we have shown that, at least when including only the simple mechanisms considered in this model, stable periodic behaviour is not possible.

    All authors declare that there is no conflicts of interest in this paper.

    Proof of Lemma 5.1. For a definite sign vector σ{1,1}2 and ϵ>0 let

    R(σ,ϵ)={xΛ+:sgn(G(x))=σ and G(x)>ϵ},

    noting that R(σ)=ϵ>0R(σ,ϵ) and σG(x)ϵ for xR(σ,ϵ). If xR(σ,ϵ) for definite σ and letting

    τ(x,ϵ)=inf{t:ϕ(t,x)R(σ,ϵ)},

    ϕ(t,x)R(σ,ϵ) for t<τ(x,ϵ), so using the second fact above,

    tσϕ(t,x)=σG(ϕ(t,x))ϵ.

    Since for y,zΛ+, |σyσz|2, it follows that τ(x)2/ϵ. Thus, letting E(σ,ϵ)=Cl(R(σ)R(σ,ϵ)) and

    E(σ)=ϵ>0E(σ,ϵ)

    where Cl denotes the closure, E(σ) consists of equilibrium points in Cl(R(σ)). Letting τ(x)=inf{t:ϕ(t,x)R(σ)}, the only way that τ(x)= is if for any ϵ>0, eventually ϕ(t,x)E(σ,ϵ), which means that ω(x)E(σ). Since ω(x) is connected and (for the cases that remain) the set of equilibrium points in Λ is finite, the first claim follows. The second claim follows since acyclic implies itineraries are finite.

    Proof of Lemma 5.2. Suppose Γ is a periodic orbit or a separatrix cycle, both of which are the image of a circle in R2, under a piecewise C1 function. Since trajectories do not cross, the image is injective, so let D be the interior of Γ, whose existence is given by the Jordan curve theorem. By Green's theorem,

    where F = (F_1, F_2) . First suppose \Gamma is a periodic orbit with solution curve u(t) satisfying u(T) = u(0) . Ignoring the orientation, since it only changes the sign, the latter integral can be parametrized with dx = \dot u_1(s)ds, \ dy = \dot u_2(s)ds as

    \int_0^T -B(u(s))(F_2(u(s))\dot u_1(s) + F_1(u(s))\dot u_2(s))ds.

    However, F_2(u(s)) = \dot u_2(s) and F_1(u(s)) = \dot u_1(s) so the integrand is identically zero, thus so is the integral, and this contradicts the assumption on \nabla \cdot (BF) . Suppose instead that \Gamma is a separatrix cycle and let u^{(1)}(t), \dots, u^{(m)}(t) denote the corresponding heteroclinic solutions. Since solutions are arranged tip-to-tail as described by (3.1), it follows that \{\dot u^{(i)}(t):t \in \mathbb{R}, i = 1, \dots, m\} orients the tangent space for \Gamma , so up to its sign, we can parametrize the above contour integral as

    \sum\limits_{i = 1}^m \int_{-\infty}^\infty -B(u(s))(F_2(u(s))\dot u_1(s) + F_1(u(s))\dot u_2(s))ds

    and the same observation shows the integral is zero, which is a contradiction.



    [1] N. Moran, The ubiquitous and varied role of infection in the lives of animals and plants, Amer. Nat. 160 (2002), S1–S8.
    [2] T. Fenchel, Ecology of Protozoa. Springer-Verlag, 1987.
    [3] C. C. Khor and M. L. Hibberd, Host-pathogen interactions revealed by human genome-wide surveys, Trends Genet. 28 (2012), 233–243.
    [4] D. Harvell, Ecology and evolution of host-pathogen interactions in nature, Amer. Nat. 164 (2004), S1–S5.
    [5] M. Li, B.Wang, M. Zhang, et al., Symbiotic gut microbes modulate human metabolic phenotypes, PNAS 105 (2008) 2117–2122.
    [6] R. M. Anderson and R. M. May, Infectious Diseases of Humans: Dynamics and Control. Oxford University Press, 1991.
    [7] C. L. Wolin and L. R. Lawlor, Models of facultative mutualism: density effects, Am. Nat., 124 (1984), 843–862.
    [8] D. H. Wright, A simple, stable model of mutualism incoporating handling time, Am. Nat., 134 (1989), 664–667.
    [9] M. Lipsitch, M. A. Nowak, D. Ebert, et al., The population dynamics of vertically and horizontally transmitted parasites, P. Roy. Soc. B-Biol. Sci., 260 (1995).
    [10] E. Foxall and N. Lanchier, Generalized stacked contact process with variable host fitness, arXiv:1511.01184 (2016).
    [11] S. J. Court, R. A. Blythe and R. J. Allen, Parasites on parasites: coupled fluctuations in stacked contact processes, EPL 101 (2013).
    [12] T. G. Kurtz, Strong approximation theorems for density dependent Markov chains, Stoch. Proc. Appl. 6 (1978), 223–240.
    [13] M. I. Freidlin and A. D. Wentzell, Random Perturbations of Dynamical Systems Third Edition, Springer-Verlag, 2012.
    [14] G. O. Poinar, Description of an early Cretaceous termite (Isoptera: Kalotermitidae) and its associated intestinal protozoa, with comments on their co-evolution, Parasite. Vector., 2 (2009).
    [15] C. L.Wolin, The population dynamics of mutualistic systems. In: D. H. Boucher (ed.) The Biology of Mutualism, Oxford University Press, New York, (1985), 248–269.
    [16] H. Amann, Ordinary Differential Equations: an Introduction to Nonlinear Analysis. de Gruyter Studies in Mathematics, 13, 1990.
    [17] L. Perko, Differential Equations and Dynamical Systems. Third edition. Texts in Applied Mathematics, 7. Springer-Verlag, New York, 2001. xiv+553 pp.
  • This article has been cited by:

    1. Eric Foxall, Nicolas Lanchier, Generalized stacked contact process with variable host fitness, 2020, 57, 0021-9002, 97, 10.1017/jpr.2019.79
    2. Eric Wajnberg, Fernando L. Cônsoli, Dynamics of Insects and Their Facultative Defensive Endosymbiotic Bacteria: A Simulation Model, 2024, 14, 2045-7758, 10.1002/ece3.70676
  • Reader Comments
  • © 2019 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(4217) PDF downloads(506) Cited by(2)

Figures and Tables

Figures(4)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog