963
Views
6
CrossRef citations to date
0
Altmetric
Articles

Global stability with selection in integro-differential Lotka-Volterra systems modelling trait-structured populations

ORCID Icon &
Pages 872-893 | Received 03 May 2017, Accepted 17 Aug 2018, Published online: 24 Oct 2018

ABSTRACT

We analyse the asymptotic behaviour of integro-differential equations modelling N populations in interaction, all structured by different traits. Interactions are modelled by non-local terms involving linear combinations of the total number of individuals in each population. These models have already been shown to be suitable for the modelling of drug resistance in cancer, and they generalize the usual Lotka-Volterra ordinary differential equations. Our aim is to give conditions under which there is persistence of all species. Through the analysis of a Lyapunov function, our first main result gives a simple and general condition on the matrix of interactions, together with a convergence rate. The second main result establishes another type of condition in the specific case of mutualistic interactions. When either of these conditions is met, we describe which traits are asymptotically selected.

1. Introduction

1.1. Biological motivations

We are interested in the evolution of N populations of individuals, each of which is structured by a continuous phenotypic trait, also called trait. In each species the phenotype models some continuous biological characteristics (such as the size of the individual, the concentration of a protein inside it, etc). We shall consider both interactions inside a given population and between the populations and we assume that mutations can be neglected. Mathematical modelling and analysis of such ecological scenarios is one purpose of the field of adaptive dynamics, a branch of mathematical biology which aims at describing evolution among a population of individuals, see [Citation15,Citation29,Citation32] for an introduction to deterministic models.

The basis for our model stems from the logistic ODE dN/dt=(rdN)N where r is the net growth rate, dN the logistic death rate due to competition for nutrients and for space by direct or indirect inhibition of proliferation between individuals. Its natural extension to a density n(t,x) of individuals of phenotype x (say in [0,1]) is a non-local logistic model (1) tn(t,x)=r(x)d(x)ρ(t)n(t,x),(1) with ρ(t):=01n(t,x)dx the total number of individuals.

Following [Citation37], one might consider that these two terms combine both Darwinism selection through the intrinsic growth rate r, and Lamarckism induction through the logistic death rate since it depends on the environment. Note that these models can be derived from stochastic models at the individual level [Citation7,Citation16,Citation21], and more generally measure-valued functions n can be considered [Citation6,Citation22]. The asymptotic behaviour of the previous model (Equation1) and variants is analysed in [Citation20,Citation26,Citation32], and one important property among others is that solution typically tend to concentrate on a few phenotypes, a convergence to Dirac masses in mathematical terms. These models are thus successful at representing the survival of only a few phenotypes, which we will refer to as selected.

To account for more complex interactions, one may want to consider a more general non-local logistic term than d(x)ρ(t)=01d(x)n(t,y)dy, in the form 01K(x,y)n(t,y)dy. The behaviour of such equations strongly depends on how localized the kernel is, and therefore so do the mathematical techniques to analyse them. Indeed, with an added diffusion term, a special case of this situation is the non-local Fisher-KPP equation. When the kernel is localized (small as soon as |xy| is large), then the solutions typically remain bounded independently of the mutation rate [Citation23]: selection is no longer a feature. This property highlights how differently the solutions behave depending on the kernel, and that some choices are not appropriate for the ecological situation we are concerned with.

The non-locality d(x)ρ(t) actually implies that interaction of an individual of phenotype x with other individuals of phenotype y has the same strength d(x) regardless of y, because individuals do not only necessarily compete with those which have close phenotypes. As such, our choice can serve as a case study to understand the effect of a blind competition across individuals, essentially mediated by the total density.

In [Citation20,Citation26], the biological motivation to use this type of models comes from drug resistance in cancer: the phenotype represents the level of resistance to a given drug and the authors argue that this might be a better approach than a discrete description of the phenotype taking only a finite number of values. Indeed, it can be correlated to some continuous biological characteristics, such as the intracellular concentration of a detoxication molecule, the activity of detoxifying enzymes in metabolizing the administered drug, or drug efflux transporters eliminating the drug [Citation9].

This model is further developed in [Citation26] to incorporate the healthy cell population. Neglecting mutations, it becomes a system of two integro-differential equations of the form n1t(t,x)=r1(x)d1(x)a11ρ1(t)+a22ρ2(t)n1(t,x),n2t(t,x)=r2(x)d2(x)a22ρ2(t)+a21ρ1(t)n2(t,x), with a11>0, a22>0, ρ1, ρ2 the total number of individuals in the cancer cell and healthy cell populations. The interspecific competition (between the two populations) is taken to be competitive, that is a12>0, a21>0, but below the intraspecific competition because each cell population is considered to belong to a different ecological niche: (2) a12<a11,a21<a22.(2) In the context of a system arises the central question of persistence (whether asymptotically both species remain), complementing that of identifying which phenotypes are selected. With the additional difficulty of control terms to represent chemotherapeutic drugs, the asymptotic properties of this model are elucidated in [Citation34], and assumption (Equation2) happens to be crucial.

In the aforementioned work, carrying out the asymptotic analysis is instrumental for the application, which is the optimal control problem of minimizing the number of cancer cells through chemotherapy. Simulations suggest that the optimal strategy is to first let the cancer cell density concentrate on a sensitive phenotype, before using the maximum tolerated doses. The convergence to Dirac masses is at the cornerstone of the theoretical analysis of why this strategy happens to be optimal. It is also in sharp contradiction with the classical approach in the clinic, which relies on constant infusion of high doses.

These integro-differential models have therefore already proved their efficiency at helping understanding phenotypic heterogeneity in cancer. The mathematical results available for N=1 and N=2 for competitive interactions naturally call for generalizations on systems of interacting species with such non-local logical terms based on the total number of individuals. For instance, to study resistance in cancer, one may think also of different cancer subpopulations interacting with healthy cells and between them, each one of them being endowed with a specific drug resistance phenotype in a tumour 'bet hedging' strategy [Citation4]. These generalizations, in turn, might help both unravel general principles about the underlying ecological processes, and develop new mathematical techniques to analyse them.

1.2. The model

We consider N populations structured by respective phenotypes xXi, where Xi is some compact subset of Rpi, with piN, for i=1,,N. Although they model distinct quantities, we abusively denote all variables x to improve readability.

The model writes (3) tni(t,x)=ri(x)+di(x)j=1Naijρj(t)ni(t,x),i=1,,N,(3) where, for i=1,,N, ri and di>0 are functions in L(Xi), ρi(t):=Xini(t,x)dx is the total number of individuals in species i, and aijR are fixed (interaction) coefficients.

The initial conditions are (4) ni(0,)=ni0i=1,,N(4) where each initial density ni0 is taken to be a non-negative function in L1(Xi). From now on, we will call these equations integro-differential Lotka-Volterra equations.

The matrix A:=(aij)1i,jN, called matrix of interactions, describes the interactions between the populations: if aij>0, the species j acts positively on the species i, and negatively if aij<0. We will also say that the interaction between species i and j for ij is:

  • mutualistic if aij>0 and aji>0,

  • competitive if aij<0 and aji<0,

  • predator-prey like if aijaji<0.

Finally, we will say that the equations are competitive (resp., mutualistic) if aij<0 (resp., aij>0) for all ij.

Another interpretation of the equations is to see them as coupled logistic equations of the form (5) tni(t,x)=(ri(x)di(x)Ii(t))ni(t,x),i=1,,N.(5) In other words, the species i reacts to its environment through the non-local variable Ii defined for i=1,,N by (6) Ii:=j=1Naijρj.(6) The terms ri(x) and di(x)Ii respectively stand for the net proliferation rate and logistic death rate of individuals in species i, of phenotype x.

We will also use the notation Ri(x,ρ1,,ρN):=ri(x)+di(x)j=1Naijρj, with which the equations rewrite: (7) tni(t,x)=Rix,ρ1(t),,ρN(t)ni(t,x),i=1,,N.(7) These models generalize Lotka-Volterra ordinary differential equation (ODE) models [Citation1]: if the functions ri, di are all constant (say equal to some ri, and di=1), then after integration with respect to xXi, the equations boil down to (8) ddtρi(t)=ri+j=1Naijρj(t)ρi(t),i=1,,N,(8) which we will from now on refer to as classical N-dimensional Lotka-Volterra equations. Thus, another advantage of a logistic term directly defined by ρ is that it makes our model more tractable with respect to the corresponding already well understood ODE models. Conversely, the integro-differential model can be seen as a perturbation of the ODE one where the individuals among a given population are allowed to have different proliferation and death rates depending on their phenotype x.

Our goal is to understand the asymptotic behaviour of the solutions to these equations, both in terms of convergence at the level of the total number of individuals ρi, and in terms of concentration at the level of the densities ni. The first problem is usual in population dynamics while the second is specific to adaptive dynamics and consists of determining which traits asymptotically survive, taking over the whole population. These are then called Evolutionary Attractors, and the fact that it is the generic situation has been coined exclusion principle. Mathematically, this corresponds to a given density ni converging to a sum of Dirac masses. For one Dirac mass only, concentration writes, for some x0Xi: (9) ni(t,)ρi(t)δx00(9) as t+, in the weak sense of measures.

The more precise aim of this paper is to study the global asymptotic stability (GAS) of what we will call coexistence steady-states, namely of possible ρ with positive components (all species asymptotically survive) such that ρ converges to ρ, because we will see how it determines on which phenotypes the densities concentrate. When it is possible, we will investigate the speed at which convergence and concentration occur. An interesting question within the scope of this paper is also to see if a result of that type is sharp, i.e. to compare the assumptions needed to obtain global asymptotic stability in our generalized setting to those known for classical Lotka-Volterra equations.

At this stage, we did not make any restrictive assumptions on the matrix A. However, it will be clear from the results recalled below in the ODE case and the ones presented in Section 2, that answers to the previous questions are available when interspecific interactions are low compared to the intraspecific ones (reminiscent of (Equation2)). Thus, we are covering the ecological scenario of each species i having its own niche, but inside which competition (if aii<0) is blind because of the term aiiρi.

Notations. In what follows, R>0N will stand for the positive orthant in RN, the set of vectors whose components are all positive, and we will write x>y when xyR>0N. We will also use the usual ordering on the set of symmetric matrices: for A a real symmetric matrices, we denote A0 (resp., A>0) when A is positive semidefinite (resp., positive definite). Finally, M1(X) will denote the set of Radon measures supported in X.

1.3. State of the art

Classical Lotka-Volterra equations. The ODE system (Equation8) has been extensively studied, dating back to the pioneering works of Lotka and Volterra for two populations of preys and predators [Citation28,Citation36]. Since then, many contributions to the analysis of steady-states and their stability have been made, and we refer to [Citation31] for an introduction and to [Citation1] for a review.

Regarding the global asymptotic stability of coexistence steady-states, a very well-known result due to Goh [Citation19] states a simple and very general condition on the matrix A=(aij)1i,jN which ensures convergence to the (unique) coexistence steady-state:

Theorem 1.1

[Citation19]

Assume that the equation Aρ+r=0 (where rRN and ρRN are the vectors (ri)1iN and (ρi)1iN) has a solution ρ in R>0N. If there exists a diagonal matrix D>0 such that ATD+DA<0, then ρ is GAS in R>0N (and hence is the unique coexistence steady-state) for system (Equation8).

A result also worth stating is that the mere existence of a unique coexistence steady-state is not enough for it to be GAS. Other steady-states on the boundary of R>0N can attract trajectories even in dimension N=2. Another possibility is the occurrence of chaotic behaviour even in low dimension as evidenced in [Citation35] for N=3. Finally, we mention the more recent work [Citation13], where the authors tackle the problem of GAS for some type of Lotka-Volterra ODEs with mutations. In particular, they obtain GAS of the coexistence steady-state in the case where the logistic variables Ii, i=1,,N all coincide, that is, when they are equal to some variable I:=j=1Najρj(t). In such a case, it is proved that the convergence to the equilibrium is exponential. The result of GAS is also extended to perturbations of this specific case.

Integro-differential Lotka-Volterra equations. The first question for such equations is the existence of a solution for all positive times. This obviously does not hold true in full generality since the ODE y=y2 is a particular case. Let us first state an existence and uniqueness theorem.

Theorem 1.2

Assume that for a given n0i=1NL1(Xi), n00, there exists 0<ρsup such that we have an a priori upper bound ρ(t)ρsup for the functions ρi whenever they are defined. Then the Cauchy problem (Equation3) and (Equation4) has a unique solution n=(ni)1iN, n0, in C([0,+),i=1NL1(Xi)).

The proof is a straightforward generalization of that given in [Citation32, Theorem 2.4] for N=1, relying on the Banach Fixed Point Theorem.

In the case of a single equation, the asymptotic behaviour is well understood. For N=1, assuming a11<0 to avoid blow-up, the equation is simply tn1(t,x)=r1(x)d1(x)ρ1(t)n1(t,x)where, without loss of generality, we have set a11=1. The first result is that ρ1 converges.

Theorem 1.3

Assume some regularity on X1, r1, d1, and r1>0, Then, for any positive continuous initial condition n10, ρ1 the function tρ1(t) is well defined on [0,+) and converges to ρ1M:=maxxX1(r1(x)/d1(x)) as t+.

This, in turn, completely determines where n1 concentrates.

Corollary 1.1

Under the previous hypotheses, n1(t), viewed as a Radon measure on X1, concentrates on the set xX1, r1(x)d1(x)ρ1M=0 as t+. If this set is reduced to some x1, we obtain in particular n1(t,)ρ1Mδx1 as t+ in M1(X1) equipped with its usual weak star topology.

A proof of this result can be found in [Citation34] in the special case X1=[0,1], and it relies on proving that ρ1 is a bounded variation (BV) function on [0,+). Let us stress that when the set on which n1 concentrates is not reduced to a singleton, the steady-state (at the level of n1) is not unique. For example, if the set is made of two points, the repartition of the limiting density on these two points depends on the initial condition, see for example [Citation12]. This is why for this equation and the general equations considered here, there is no hope in proving general GAS results directly at the level of the densities ni.

For a general logistic term (XK(x,y)n(t,y)dy)n(t,x) and a single equation, the asymptotic behaviour is also analysed in detail in both [Citation14] and [Citation24]. In the latter, under some suitable assumptions on the kernel K, a Lyapunov functional is used to prove that some measure is GAS, in a very specific sense depending on K. Similar results can be found in [Citation8], where their counterpart for competitive classical Lotka-Volterra equations are also discussed.

In the case of integro-differential systems, however, much less is known about the asymptotic behaviour. A Lyapunov functional inspired by [Citation24] has been used successfully in [Citation34] to prove GAS for a competitive system of two populations which writes exactly as (Equation3). We also mention [Citation5] where an integro-differential system of two populations is analysed, and whose form does not fit in our framework. A particular triangular coupling structure allows the authors to perform an asymptotic analysis.

The paper is organized as follows. In Section 2, we explain how coexistence steady-states can be identified, allowing us to state rigorously what we mean by GAS for system (Equation3). We explain why, under the hypothesis of GAS, only some phenotypes are generically selected, and how to compute them. Then, we present the two main results about GAS for such equations. Section 3 is devoted to the proof of the first result, which applies for any type of interactions and relies on analysing a suitably designed Lyapunov functional. In the specific case of mutualistic interactions, our second main result gives alternative conditions sufficient for GAS. It is presented in Section 4. In Section 5, we conclude with several comments and open questions.

2. Possible coexistence steady-states and main results

For the rest of the article, we will work with the following assumptions: (10) ri,di,ni0C(Xi),ni0>0 for i=1,,N.(10) This will simplify statements, but we will be more specific below as to which data our results generalize.

2.1. Analysis of coexistence steady-states

In the context of this system of integro-differential equations, the expression ‘GAS in R>0N’ must be defined. By that, we mean that there exists ρ>0 such that, whatever the positive continuous initial conditions ni0 are, ρi converges to ρi for all i.

First, let us explain how to compute the possible steady-states at the level of ρ, i.e. possible limits ρ>0 for positive continuous initial conditions. We will work with the following topological assumption on the sets Xi: (11) xXi,η>0,λpiB(x,η)Xi>0,(11) where λpi stands for the Lebesgue measure on Rpi and B(x,η) for the open ball of centre x and radius η.

Assume that each ρi converges to some ρi>0, in which case the exponential behaviour of ni(t,x) is asymptotically governed by ri(x)+di(x)j=1Naijρj, the sign of which we can analyse as follows.

  • If this quantity is positive for some x0, let us prove that ni(t,x) blows up in its neighbourhood, leading to the explosion of ρi. If ri(x0)+di(x0)j=1Naijρj>0, there exists η>0 such that by continuity ri(x)+di(x)j=1Naijρj>0 on (B(x0,η)Xi), and then λpi(B(x0,η)Xi)>0 whether x0int(Xi) or also if x0Xi thanks to (Equation11). For ε>0 small enough and t large enough (say tt0) such that ri(x0)+di(x0)j=1Naijρj>ε, we can write: ρi(t)B(x0,η)Xini(t,x)dxB(x0,η)Xini(t0,x)et0tRix,ρ1(s),,ρN(s)dsdxλpiB(x0,η)XiinfB(x0,η)Xini(t0,x)eε(tt0), with the right-hand side blowing up as t+, which cannot be compatible with the convergence of ρi.

  • If ri+dij=1Naijρj is negative globally on Xi, this clearly implies the extinction of species i, which is also incompatible with the convergence of ρi to a positive limit.

Remark 2.1

It is possible to relax the regularity on both the sets Xi and the data ri and di by working only with points which are both Lebesgue points of ri/di and of Lebesgue density 1 for Xi, see [Citation17]. If the functions ri/di are in L1(Xi), one can indeed check that ri+dij=1Naijρj0 a.e. on Xi.

The previous results motivate the following definition: Ii:=maxxXiri(x)di(x),i=1,,N. With this definition, a steady-state ρ>0 exists if and only if the following assumption holds: (12) the equation Aρ+I=0 has a solution ρ in R>0N,(12) which we assume from now on.

The previous computations also show that ni vanishes where ri(x)di(x)Ii<0, which implies the following result:

Proposition 2.1

Assume that assumption (Equation12) holds, and that ρ converges to the coexistence steady-state ρ. Then, ni(t), viewed as a Radon measure, concentrates on the set Ki:={xXi, ri(x)di(x)Ii=0} as t+, for all i=1,,N.

If, for some i, Ki is reduced to some x, we obtain in particular ni(t,)ρiδxi as t+ in M1(Xi).

Densities ni of total mass ρi and concentrated on Ki are called Evolutionary Stable Distributions (ESD) in [Citation24].

Remark 2.2

In the hypothesis of global existence and convergence of ρ towards ρ, the previous reasoning actually shows that the concentration is ensured as soon as ni0L1(Xi) is bounded by below by a positive constant on a neighbourhood of one of the points of Ki. For more general hypotheses ensuring concentration, we refer to [Citation24].

Remark 2.3

If all the sets Ki are reduces to some singletons xi, then the dynamics of ρ are asymptotically governed by classical Lotka-Volterra equations concentrated in (x1,,xN), namely ddtρi(t)ri(xi)+di(xi)j=1Naijρj(t)ρi(t),i=1,,N, as t goes to +. For a precise statement, see [Citation34].

2.2. Analysis of coexistence steady-states

Our first approach to prove GAS is to further generalize the approach of [Citation34] in dimension N. The main idea is to mix a Lyapunov functional which is inspired by the one designed in [Citation24] and the Lyapunov functional used for classical Lotka-Volterra equations [Citation19], which is the key tool to obtain Theorem 1.1. With some mild regularity assumptions on the data, we obtain the following result:

Theorem 2.1

Assume (Equation12) and that there exists a diagonal matrix D>0 such that DA is symmetric and DA<0. Then the solution to the Cauchy problem (Equation3) and (Equation4) is globally defined. Furthermore, the solution ρ to Aρ+I=0 is GAS (and hence, unique).

We emphasize that there is no assumption on the type of interactions, i.e. on the sign of the coefficients of A. However, a consequence of this result is that A must be such that aiiajj>aijaji for all i, j. For this result to apply, interactions must therefore be stronger inside species than between them.

We also remark that our hypothesis is exactly the one exhibited in [Citation8] for competitive classical Lotka-Volterra equations. The analysis of the Lyapunov functional allows to determine a speed at which convergence to ρ and concentration on a given set of phenotypes occur. In dimension 2, we also analyse more deeply the link between this condition and the one for classical Lotka-Volterra equations, which in most interesting cases happen to be equivalent.

Our second main result focuses on the special case of mutualistic interactions, and an informal statement of the theorem is the following.

Theorem 2.2

Assume (Equation12), that for i=1,,N, ri>0 and that for some explicit 0<ci<Ci, the matrix Aˆ defined by aˆii:=ciaii and aˆij=Ciaij for ij is Hurwitz. Then the solution to the Cauchy problem (Equation3) and (Equation4) is globally defined. Furthermore, the solution ρ to Aρ+I=0 is GAS.

Again, this applies to the case of interspecific interactions being higher that intraspecific ones, because a Hurwitz matrix is a matrix such that all its eigenvalues have negative real part and it has to do with diagonally dominant matrices (see Section 4).

Because of the hypothesis of mutualism, the system is cooperative, and sub and supersolution techniques can succeed. More precisely, it is possible to prove that all functions ρi are BV on [0,+) and this implies their convergence.

3. General interactions

3.1. Proof of the main theorem

In this section, we need slightly more regularity for the data, namely that the functions are Lipschitz continuous: (13) for i=1,,N,ri,diC0,1(Xi).(13) We can now restate the first theorem:

Theorem 3.1

Assume (Equation12) and (Equation13). Assume that there exists a diagonal matrix D>0 such that DA is symmetric and DA<0. Then the solution to the Cauchy problem (Equation3) and (Equation4) is globally defined.

Furthermore, the solution ρ to Aρ+I=0 is GAS with (14) ρ(t)ρ=Oln(t)t1/2.(14) Concentration of a given ni occurs at speed O(ln(t)/t), in the following sense: (15) Ximi(x)Rix,ρ1,,ρNni(t,x)dx=Oln(t)t.(15) In particular, if Ki is reduced to a singleton xi, then (16) ε>0,XiBxi,εni(t,x)dx=Oln(t)t.(16)

Proof.

First step: definition of the Lyapunov functional.

From (Equation12), Evolutionary Stable Densities exist and we can pick one of them: for i=1,,N, we choose any measure ni in M1(Xi) satisfying ni(Xi)=ρi, which is furthermore concentrated on Ki, i.e. (17) supp (ni)Ki.(17) We abusively write integration of functions g against measures μ as Xg(x)μ(x)dx. We also set mi:=1/di and define N functions Vi by Vi(t):=Ximi(x)ni(x)ln1ni(t,x)+ni(t,x)ni(x)dx. In what follows, we consider the following Lyapunov functional: V(t):=i=1NλiVi(t) where the positive constants λi are to be chosen later. The diagonal matrix of diagonal entries λ1,,λN is denoted by D.

Second step: computation and sign of the derivative.

For any i, we compute dVidt=Ximi(x)ni(x)tni(t,x)ni(t,x)+tni(t,x)dx=Ximi(x)Rix,ρ1,,ρNni(t,x)ni(x)dx=Ximi(x)Rix,ρ1,,ρNRix,ρ1,,ρNni(t,x)ni(x)dx+Ximi(x)Rix,ρ1,,ρNni(t,x)ni(x)dx. The definition of mi implies that the first term simplifies as follows Ximi(x)di(x)Aρρini(t,x)ni(x)dx=Aρρiρiρi. For the second term, the choice (Equation17) leads to Bi(t):=Ximi(x)Rix,ρ1,,ρNni(t,x)dx. The functions Bi are all non-positive by definition of ρ.

Defining the vector u:=ρρ, we arrive at: dVdt=i=1NλiAρρiρiρi+i=1NλiBi=i=1Nλi(Au)iui+i=1NλiBi. Thus, we end up with the expression (18) dVdt=uT(DA)u+i=1NλiBi.(18) Since the antisymmetric part of DA does not play any role, this can also be expressed dVdt=12uT(ATD+DA)u+i=1NλiBi. By assumption, M:=ATD+DA<0, from which we deduce dV/dt0. Furthermore, the convergence of the term uTMu to 0 is equivalent to that of ρ to ρ. However, we do not have the usual property V0 for Lyapunov functions, so that we cannot yet conclude.

Third step: estimates on dV/dt.

Let G:=12uTMu+2i=1NλiBi. We are going to show that G is non-decreasing.

We denote by u,v the canonical scalar product of two vectors u and v in RN. ddtuT(DA)u=ddt(DA)u,u=(DA)dudt,u+(DA)u,dudt. For i=1,,N, the derivative of Bi is given by dBidt=Ximi(x)Rix,ρ1,,ρNRix,ρ1,,ρNni(t,x)dx=Ximi(x)Ri2x,ρ1,,ρNni(t,x)dx+Ximi(x)Rix,ρ1,,ρNRix,ρ1,,ρNRix,ρ1,,ρNni(t,x)dx[Aρρ]XiRix,ρ1,,ρNni(t,x)dx=(Au)idudti leading to the bound ddti=1NλiBii=1Nλi(Au)idudti=(DA)u,dudt. Put together, these estimates yield: dGdt(DA)dudt,u+(DA)u,dudt2(DA)u,dudt=(DA)dudt,u(DA)u,dudt. The last expression is equal to 0 if DA is symmetric, in which case G is non-decreasing as claimed.

The assumptions that DA is symmetric and that ATD+DA<0 are equivalent to the assumption made for the theorem: DA is supposed to be a symmetric negative definite matrix.

As a consequence of the monotonicity of G, we get uT(DA)uG(0) for all t. The left-hand side is the square of some norm on RN, which means that ρ has to be bounded: these a priori bounds ensure the global definition of the solution to (Equation3) and (Equation4).

Fourth step: a lower estimate for V .

To estimate V from below, we need a uniform (in x) upper bound on the densities ni. Because of the regularity assumption (Equation13), there exists C>0 such that: i=1,,N,(x,y)Xi2,Riy,ρ1,,ρNRix,ρ1,,ρNC|xy|. The constant C can be chosen to be independent of t since the functions ρi are bounded.

This implies for a given i Xini(t,y)dy=Xini0(y)exp0tRiy,ρ1,,ρNdsdyXini0(y)ni0(x)ni0(x)exp0tRix,ρ1,,ρNexpCt|xy|dyni(t,x)ni0(x)XiexpCt|xy|dy. Computing the integral, we write, thanks to the boundedness of ρi ni0 and ni0 and (C has changed and is independent of t and x): for t large enough, ni(t,x)Ct. The bound on V follows immediately: (19) V(t)Cln(t)+1.(19)

Fifth step: convergence.

G bounds dV/dt from above: dV/dt12G. Thus V(t)V(0)120tG(s)ds12tG(t) thanks to the third step.

We can now write G(t)C((ln(t)+1)/t): G=O(ln(t)/t), consequently, each non-positive term it is composed of also vanishes like O(ln(t)/t) as t+.

In other words, 12uTMu and each Bi onverge to 0 as as well O(ln(t)/t). This is nothing but the two first statements (Equation14) and (Equation15).

For the last statement, we fix i and ε>0. We denote hi:=miRi(,ρ1,,ρN), which is non-negative on Xi, and by assumption vanishes at xi only. We choose a>0 small enough such that a1XiB(xi,ε)h on Xi. This enables us to write XiBxi,εni(t,x)dx1aXimi(x)Rix,ρ1,,ρNni(t,x)dx=Oln(t)t.

Remark 3.1

The first interesting fact is that the convergence rate of G to 0, in O(ln(t)/t), is almost optimal in many cases. Indeed, if the sets Ki are reduced to singletons, there cannot exist any α>1 such that this sum vanishes like O(1/tα). This comes from the fact that if it were true, dV/dt would be integrable on the half-line, which would imply the convergence of V . This is not possible since each Vi has to go to as t goes to +.

This might seem contradictory with the exponential convergence rates obtained in [Citation13] for some classical Lotka-Volterra equations, but the Lyapunov functional gives us information on the speed of both phenomena in the sense defined above (through the function G) and it does not say whether one of the two is faster.

3.2. Sharpness in dimension 2

It is clear that if we can find D>0 diagonal such that DA is symmetric and DA<0, then ATD+DA<0. The condition that DA should be symmetric is constraining, especially if N3 in which case it imposes some polynomial equalities on the coefficients of the matrix A. In dimension 2, however, the result is sharp in various contexts, as stated in the following proposition.

Proposition 3.1

Assume N=2, a11<0, a22<0 and a12a21>0. Then the following conditions are equivalent.

  1. there exists D>0 diagonal such that DA is symmetric and DA<0;

  2. there exists D>0 diagonal such that ATD+DA<0;

  3. det(A)>0.

Proof.

(i) implies (ii) as noticed before.

Now, let us assume (ii) and compute M:=ATD+DA=2λ1a11λ1a12+λ2a21λ1a12+λ2a212λ2a22, which has positive determinant, i.e. det(M)=4λ1λ2a11a22(λ1a12+λ2a21)2>0. If det(A)0, det(M)4λ1λ2a12a21(λ1a12+λ2a21)2=(λ1a12λ2a21)20, a contradiction.

Now, if (iii) holds, we take λ1:=1/|a12| and λ1:=1/|a21| for which DA=a11|a12|sgn (a12)sgn (a21)a22|a21| is clearly symmetric negative definite, whence (i).

4. Cooperative case

4.1. Some facts about Hurwitz matrices

We now focus on the cooperative case, i.e. on the case where all off-diagonal elements of A are non-negative. We will also assume that the diagonal elements are negative, since otherwise blow-up clearly occurs: there is intra-spectific competition inside any given species. We will say that such a matrix is cooperative.

In this case, we can hope for stronger results at the level of the integro-differential system because we can use sub and super-solution techniques. For our purpose, the following result on ODEs is sufficient.

Lemma 4.1

For T>0 (possibly T=+), let f:[0,T)×RNRN be a continous function on [0,T)×RN, locally Lipschitz in xRN uniformly in t[0,T). Denoting f(t,x):=(fi(t,x1,,xN))1iN, further assume that for all i=1,,N, fi is non-decreasing with xj for all ji.

Assume that we have a solution z on [0,T) of the following Cauchy problem: (20) dzdt=f(t,z)z(0)=z0,(20) where z0RN, and a function y subsolution to the previous Cauchy problem, i.e. (21) dydtf(t,y)y(0)z0.(21) Then y(t)z(t) on [0,T).

When the matrix A is cooperative, it is possible to give an equivalent condition to the one required in Theorem 1.1 for GAS in classical Lotka-Volterra equations. Let us explain how, starting with the three following lemmas, the two first of which can be found in [Citation1].

Lemma 4.2

Let A be a cooperative matrix. Then it is Hurwitz if and only if it is negatively diagonally dominant, i.e. if and only if there exists a vector v>0 such that aiivi+jiaijvj<0 for i=1,,N.

This first lemma will be useful in its own right in this section. A consequence is that

Lemma 4.3

If A is cooperative and r>0, Aρ+r=0 has a unique solution in R>0N if and only if A is Hurwitz.

Finally, it comes from the theory of M-matrices (see [Citation33] for a review) that

Lemma 4.4

Let A be cooperative. Then A is Hurwitz if and only if there exists D>0 diagonal such that ATD+DA<0.

An important consequence of these three lemmas is the following rephrasing of Theorem 1.1 for classical Lotka-Volterra equations.

Proposition 4.1

Assume that A is cooperative, r>0 and that the equations (Equation8) have a unique steady-state in R>0N. Then the equations are globally defined and this steady-state is GAS.

Thus, in the context of cooperation between the species, the requirement that A is Hurwitz is somehow optimal to obtain a GAS coexistence steady-state, since it is already required to have its mere existence, a fact mentioned in [Citation18]. We will mainly work with this characterization (rather than the equivalent one given by Lemma 4.4 which we used for a general interaction matrix A) because the next results will lead us to modify the matrix A: analysing whether it is still Hurwitz or not is easier than checking this equivalent condition.

4.2. A priori bounds

For the remaining part of this section, we make the assumption that ri is positive on Xi for i=1,,N, and we define the lower and upper bounds 0<dimdi(x)diM, 0<rimri(x)riM.

Theorem 4.1

Assume that the matrix A~ defined by a~ii:=dimaii and a~ij:=diMaij is Hurwitz. Then the solutions to (Equation3) are globally defined and bounded.

Proof.

First remark that since A~ is Hurwitz, then so is A from Lemma 4.2.

We integrate the equations with respect to x and bound them (through ri(x)riM) ddtρi(t)riM+j=1Naijρj(t)Xidi(x)ni(t,x)dxi=1,,N. Since the diagonal elements are negative, the off-diagonal non-negative, we obtain ddtρi(t)riM+aiidimρi+jiNaijdiMρj(t)ρi(t),i=1,,N. Thus, the vector (ρ1,,ρN) is a subsolution of the previous system which is nothing but classical Lotka-Volterra equations with interaction matrix A~. Thanks to (4.1), the solutions to this system are bounded. Thus, so are those of the integro-differential one.

Remark 4.1

Note that the assumption that A~ is Hurwitz reduces to A being Hurwitz in the case of constant coefficients. Thus, this result for boundedness is sharp, since it is exactly the one required for convergence to the coexistence steady-state when the equations at hand are classical Lotka-Volterra equations.

Using Theorem 4.1, we can thus define ρM>0 as the GAS steady-state for the system obtained in the previous proof, that is to the equations ddtui=riM+aiidimui+jiNaijdiMuj(t)ui(t),i=1,,N. In other words, ρM:=(A~)1rM where rM is the vector (riM)1iN. This means that we can write (22) lim supt+ρiρiMi=1,,N.(22) In a similar fashion to the previous proposition, bounding ρi away from 0: ddtρirim+aiidiMρi+jiNaijdimρj(t)ρi(t),i=1,,N, leading to (23) lim inft+ρiρimi=1,,N(23) where ρm:=(B)1rm>0 with B, a Hurwitz matrix defined by bii:=diMaii, bij:=dimaij for ij and rm:=(rim)1iN.

4.3. GAS in the mutualistic case

We can now state the main result:

Theorem 4.2

Assume ri>0 for all i=1,,N, and that the matrix Aˆ defined by aˆii:=dimρimaii and aˆij:=diMρiMaij for ij, is Hurwitz.

Then A~, A and B are also Hurwitz. Furthermore, ρ:=A1I lies in R>0N and it is GAS.

Proof.

The fact that A~, A and B are also Hurwitz is a direct consequence of Lemma 4.2.

A~ being Hurwitz, the solutions are globally defined with ρ bounded thanks to Theorem 4.1.

A being Hurwitz, it is invertible and ρ:=A1I is in R>0N thanks to Lemma 4.3.

Let us now prove that it is GAS. The idea is to prove that each ρi is BV on [0,+). Identifying the limit is straightforward, thanks to the reasoning made in Section 2.

For any i, we define qi:=dρi/dt and write Ri=Ri(x,ρ1,,ρN) for readability. Since qi=dρi/dt=XiRini, we obtain dqidt=XiRi2ni+Xij=1NRiρjqjnij=1NaijXidi(x)ni(t,x)dxqj. Let bi(t):=Xidi(x)ni(t,x)dx. The idea is that ρi is ‘mostly’ increasing, so we are interested in the negative part of qi, denoted by (qi). For this quantity we have the (a.e.) bound d(qi)dtbij=1Naijqj1qi<0 On the one hand, biaiiqi1qi<0=biaii(qi). On the other hand, for ij, biaijqj1qi<0biaij(qj). Combining these two, we get d(qi)dtbiA(q)i. We fix ε>0 small enough and t large enough (say tt0) such that Aˆ+εJ is Hurwitz (where J is the matrix composed of ones only) and such that, for each (i,j), bi(t)aijaˆij+ε. The first requirement is easily derived from Lemma 4.2 since Aˆ+εJ is clearly cooperative and negatively diagonally dominant for ε>0 small enough. The second requirement comes from the lower and upper bounds for the functions ρi as stated in (Equation22) and (Equation23).

The resulting inequality is d(qi)dtAˆ+εJ(q)i, so that ((q1),,(qN)) is a subsolution of the system with same initial conditions at t0, given by dydt=Aˆ+εJy. The solutions to this system go exponentially to 0 since Aˆ+εJ is Hurwitz.

For any i, we have thus proved that (qi) goes to 0 exponentially. Together with the fact that ρi is bounded from above, we conclude that it is BV on [0,+). Indeed it holds true that a function u which is both bounded from above and such that uL1([0,+)) is BV on [0,+), see [Citation32, Lemma 6.7]. Therefore, ρ converges (to ρ).

5. Conclusion

We have analysed the global asymptotic stability properties for integro-differential systems of N species structured by traits x belonging to different trait spaces Xi. The coupling comes from a non-local logistic term, which is a linear combination of the total number of individuals ρi in each species. Theses systems generalize the usual Lotka-Volterra ODEs for which many stability analyses are available in the literature. Our main focus has been on the asymptotic behaviour of the functions ρi(t) as t+, especially towards equilibrium ρ with positive components, i.e. of persistence of all species. In Section 2, we explained how identifying it essentially determines the asymptotic behaviour of the underlying density ni, namely the phenotypes on which the measures ni(t,) concentrate in large time.

In Section 3, an adequate Lyapunov functional allowed us to state a general result relying solely on an assumption on the matrix A, regardless of the type of interactions. For N=2, this is essentially a sharp result, but becomes more restrictive for N3. This tool also provided us with convergence rates to equilibrium. In Section 4, we presented another strategy based on a BV bound which yielded a second result of global asymptotic stability, this time for mutualistic equations.

The result of Theorem 4.2 is partly less general than the one given in Theorem 3.1 because it requires a sign on the coefficients of A. However, the set of matrices which satisfy the hypothesis given in the last theorem is an open subset of the set of real matrices RN×N in any dimension. This in sharp contrast with the hypothesis of Theorem 3.1, which, as already mentioned, imposes some polynomial equalities on the coefficients of A as soon as N3. In other words, for a small perturbation of a cooperative matrix for which GAS holds, GAS still holds. In particular, if one has weakly (but mutualistically) coupled equations, GAS holds, whereas Theorem 3.1 does not cover the case of any weakly coupled equations for general interactions, except for N=2.

In both cases, the assumptions fall within the class of matrices which cannot have off-diagonal coefficients which are too high compared to the diagonal ones. The present results thus apply to cases where interactions among individuals of a same species are not only blind because of the term a11ρ1, but also stronger than the interactions between species. In other words, each one of them has its own ecological niche inside which interactions are independent of how given phenotypes x and y are away from another.

Let us remark that the BV method would apply to more general functions Ri(x,ρ1,,ρN), as long as they are increasing in the variables ρj, ji. However, the Lyapunov functional used in Theorem 3.1 seems to be dependent on the linear coupling chosen here and it is an open problem to generalize our results for other settings. Another open question is about finding whether there are matrices A for which the underlying classical Lotka-Volterra equations converge to the coexistence steady state (for example such that there exists D>0 with ATD+DA<0), but for which there is no GAS for the integro-differential system. Numerically at least, we could not build any such case.

We finally mention a natural extension of these integro-differential systems, which has drawn much attention in adaptive dynamics: that is when one adds a mutation term M[n], and a typical single equation then writes (24) tn(t,x)=r(x)XK(x,y)n(t,y)dyn(t,x)+M[n](t,x).(24) This is usually done either through a second order elliptic operator like M[n](t,x)=βΔn(t,x), with Neumann boundary conditions, or using an integral operator allowing for long-range mutations like M[n](t,x)=Xm(x,y)n(t,y)n(t,x)dy. Finally, let us mention that in some more recent works an advection term is also considered [Citation10,Citation11]. The idea is to model stress-induced adaptation: individuals actively adapt to their environment and this can be thought of as an appropriate modelling of Lamarckism induction.

Most of the studies on model (Equation24) have aimed at understanding how small mutations affect the dynamics of the surviving phenotype (when one expects a single Dirac mass) with a vanishing mutation term Mε, after a proper rescaling of time [Citation2,Citation27,Citation30]. Without smallness assumptions on the mutations, non-trivial steady-states have been investigated in detail in [Citation3,Citation12,Citation25], and results of GAS have essentially been obtained for K(x,y)=k(y). It would be interesting to investigate the asymptotic stability properties of steady states for more general kernels. To the best of our knowledge, general asymptotic results are indeed still elusive, even in the case explored in our paper, namely when K(x,y)=d(x) for d non constant. It is not clear whether the techniques developed in the present paper can be extended to that setting nor to systems of this form.

Disclosure statement

No potential conflict of interest was reported by the authors.

References

  • S. Baigent, Lotka-Volterra Dynamics: an introduction, preprint (2010).
  • G. Barles, S. Mirrahimi, and B. Perthame, Concentration in Lotka-Volterra parabolic or integral equations: A general convergence result, Methods Appl. Anal. 16(3) (2009), pp. 321–340.
  • O. Bonnefon, J. Coville, and G. Legendre, Concentration phenomenon in some non-local equation, preprint (2015). Available at arXiv:1510.01971.
  • B. Brutovsky and D. Horvath, Structure of intratumor heterogeneity: Is cancer hedging its bets? (2013). arxiv, page 1307.0607.
  • J.-E. Busse, P. Gwiazda, A. Marciniak-Czochra, Mass concentration in a nonlocal model of clonal selection. J. Math. Biol. 73 (2016), pp. 1–33. doi: 10.1007/s00285-016-0979-3
  • J.A. Cañizo, J.A. Carrillo, and S. Cuadrado, Measure solutions for some models in population dynamics, Acta Appl. Math. 123(1) (2013), pp. 141–156. doi: 10.1007/s10440-012-9758-3
  • N. Champagnat, R. Ferrière, and S. Méléard, From individual stochastic processes to macroscopic models in adaptive evolution, Stoch. Models 24(S1) (2008), pp. 2–44. doi: 10.1080/15326340802437710
  • N. Champagnat, P.-E. Jabin, and G. Raoul, Convergence to equilibrium in competitive Lotka-Volterra and chemostat systems, Comptes Rendus Math. 348(23) (2010), pp. 1267–1272. doi: 10.1016/j.crma.2010.11.001
  • R.H. Chisholm, T. Lorenzi, and J. Clairambault, Cell population heterogeneity and evolution towards drug resistance in cancer: Biological and mathematical assessment, theoretical treatment optimization, Biochimic. Biophys. Acta (BBA) - General Subjects 1860(11) (Nov 2016), pp. 2627–2645. doi: 10.1016/j.bbagen.2016.06.009
  • R.H. Chisholm, T. Lorenzi, and A. Lorz, Effects of an advection term in nonlocal lotka-volterra equations, Commun. Math. Sci. 14 (2016), pp. 1181–8. doi: 10.4310/CMS.2016.v14.n4.a16
  • R.H. Chisholm, T. Lorenzi, A. Lorz, A.K. Larsen, L.N. de Almeida, A. Escargueil, and J. Clairambault, Emergence of drug tolerance in cancer cell populations: An evolutionary outcome of selection, nongenetic instability, and stress-induced adaptation, Cancer Res. 75(6) (2015), pp. 930–939. doi: 10.1158/0008-5472.CAN-14-2103
  • J. Coville, Convergence to equilibrium for positive solutions of some mutation-selection model, preprint (2013). Available at arXiv:1308.6471.
  • J. Coville and F. Fabre, Convergence to the equilibrium in a Lotka-Volterra ODE competition system with mutations, preprint (2013). Available at arXiv:1301.6237.
  • L. Desvillettes, P.E. Jabin, S. Mischler, and G. Raoul, On selection dynamics for continuous structured populations, Commun. Math. Sci. 6(3) (2008), pp. 729–747. doi: 10.4310/CMS.2008.v6.n3.a10
  • O. Diekmann, A beginner's guide to adaptive dynamics, Banach Center Publications 63 (2004), pp. 47–86.
  • O. Diekmann, M. Gyllenberg, and J. Metz, Steady-state analysis of structured population models, Theor. Popul. Biol. 63(4) (2003), pp. 309–338. doi: 10.1016/S0040-5809(02)00058-8
  • L.C. Evans, R.F. Gariepy, Measure Theory and Fine Properties of Functions, CRC press, Boca Raton, 2015.
  • B. Goh, Stability in models of mutualism. Am. Nat. 113 (1979), pp. 261–275. doi: 10.1086/283384
  • B.S. Goh, Global stability in many-species systems. Am. Nat. 111 (1977), pp. 135–143. doi: 10.1086/283144
  • J. Greene, O. Lavi, M.M. Gottesman, and D. Levy, The impact of cell density and mutations in a model of multidrug resistance in solid tumors, Bull. Math. Biol. 76(3) (2014), pp. 627–653. doi: 10.1007/s11538-014-9936-8
  • J.M. Greene, D. Levy, K.L. Fung, P.S. Souza, M.M. Gottesman, and O. Lavi, Modeling intrinsic heterogeneity and growth of cancer cells, J. Theor. Biol. 367 (2015), pp. 262–277. doi: 10.1016/j.jtbi.2014.11.017
  • M. Gyllenberg and G. Meszéna, On the impossibility of coexistence of infinitely many strategies, J. Math. Biol. 50(2) (2005), pp. 133–160. doi: 10.1007/s00285-004-0283-5
  • F. Hamel and L. Ryzhik, On the nonlocal fisher–kpp equation: Steady states, spreading speed and global bounds, Nonlinearity 27(11) (2014), pp. 2735. doi: 10.1088/0951-7715/27/11/2735
  • P.-E. Jabin and G. Raoul, On selection dynamics for competitive interactions, J. Math. Biol. 63(3) (2011), pp. 493–517. doi: 10.1007/s00285-010-0370-8
  • H. Leman, S. Méléard, and S. Mirrahimi, Influence of a spatial structure on the long time behavior of a competitive Lotka-Volterra type system, Discrete Continuous Dyn. Syst. Ser. B. J. Bridging Math. Sci. 20(2) (2015), pp. 469–493. doi: 10.3934/dcdsb.2015.20.469
  • A. Lorz, T. Lorenzi, M.E. Hochberg, J. Clairambault, and B. Perthame, Populational adaptive evolution, chemotherapeutic resistance and multiple anti-cancer therapies, ESAIM: Math. Model. Numer. Anal. 47(2) (2013), pp. 377–399. doi: 10.1051/m2an/2012031
  • A. Lorz, S. Mirrahimi, and B. Perthame, Dirac mass dynamics in multidimensional nonlocal parabolic equations, Commun. Part. Diff. Equ. 36(6) (2011), pp. 1071–1098. doi: 10.1080/03605302.2010.538784
  • A.J. Lotka, Elements of physical biology, (1924). Reprinted 1956 as elements of mathematical biology.
  • J.A. Metz, O. Diekmann, The Dynamics of Physiologically Structured Populations, Vol. 68, Springer, Berlin, 2014.
  • S. Mirrahimi and B. Perthame, Asymptotic analysis of a selection model with space, J. Math. Pures Appl. 104(6) (2015), pp. 1108–1118. doi: 10.1016/j.matpur.2015.07.006
  • J.D. Murray, Mathematical Biology I: An Introduction, Vol. 17 of Interdisciplinary Applied Mathematics, 2002.
  • B. Perthame, Transport Equations in Biology, Springer Science & Business Media, Basel, 2006.
  • R.J. Plemmons, M-matrix characterizations. I-nonsingular M-matrices, Linear Algebra Appl. 18(2) (1977), pp. 175–188. doi: 10.1016/0024-3795(77)90073-8
  • C. Pouchol, J. Clairambault, A. Lorz, E. Trélat, Asymptotic analysis and optimal control of an integro-differential system modelling healthy and cancer cells exposed to chemotherapy. J. Math. Pures Appl. 116 (2017), pp. 268–308. doi: 10.1016/j.matpur.2017.10.007
  • R. Vance, Predation and resource partitioning in one predator – two prey model communities. Am. Nat. 112 (1978), pp. 797–813. doi: 10.1086/283324
  • V. Volterra and M. Brelot, Leçons sur la théorie mathématique de la lutte pour la vie, Vol. 1, Gauthier-Villars, Paris, 1931.
  • J.X. Zhou, A.O. Pisco, H. Qian, and S. Huang, Nonequilibrium population dynamics of phenotype conversion of cancer cells, PloS one 9(12) (2014), pp. e110714. doi: 10.1371/journal.pone.0110714