PreprintPDF Available
Preprints and early-stage research may not have been peer reviewed yet.

Abstract

We study the large-time behavior of an ensemble of entities obeying replicator-like stochastic dynamics with mean-field interactions as a model for a primordial ecology. We prove the propagation-of-chaos property and establish conditions for the strong persistence of the N-replicator system and the existence of invariant distributions for a class of associated McKean-Vlasov dynamics. In particular, our results show that, unlike typical models of neutral ecology, fitness equivalence does not need to be assumed but emerges as a condition for the persistence of the system. Further, neutrality is associated with a unique Dirichlet invariant probability measure. We illustrate our findings with some simple case studies, provide numerical results, and discuss our conclusions in the light of Neutral Theory in ecology.
PERSISTENCE AND NEUTRALITY IN INTERACTING REPLICATOR
DYNAMICS
LEONARDO VIDELA1, MAURICIO TEJO2, CRISTÓBAL QUIÑINAO3, PABLO A. MARQUET3,4,
AND ROLANDO REBOLLEDO5
Abstract. We study the large-time behavior of an ensemble of entities obeying replicator-
like stochastic dynamics with mean-field interactions as a model for a primordial ecology. We
prove the propagation-of-chaos property and establish conditions for the strong persistence of
the N-replicator system and the existence of invariant distributions for a class of associated
McKean-Vlasov dynamics. In particular, our results show that, unlike typical models of
neutral ecology, fitness equivalence does not need to be assumed but emerges as a condition
for the persistence of the system. Further, neutrality is associated with a unique Dirichlet
invariant probability measure. We illustrate our findings with some simple case studies,
provide numerical results, and discuss our conclusions in the light of Neutral Theory in
ecology.
Keywords: Stochastic replicator dynamics, Propagation of Chaos, Stochastic persis-
tence, McKean-Vlasov equation, Invariant distributions, Emergence of ecologies.
Mathematics Subject Classification (2010): 60H10 and 92D25.
1. Introduction
The emergence of life corresponds to the emergence of an ecological system of interacting
self-replicating entities. Several characteristics of early life have endured through time and are
shared by all forms of life today, from the near universality of the genetic code to the inter-
mediate metabolism [32, 51]. It is unlikely that these biological universalities represent only
"frozen accidents" linked to a universal common ancestry, but may emerge as a consequence of
some fundamental principles associated with information and thermodynamics that affect the
robustness and evolvability of biological systems [51, 60]. In a similar vein, it is expected that
early ecological systems faced important challenges to their persistence in fluctuating environ-
ments, where the number of co-existing entities increases forming diverse communities, and that
some fundamental principles may also be invoked to understand their persistence. In this work,
we are primarily interested in the principle of neutrality, which is mathematically expressed
in a particular invariant probability distribution associated with the process of abundances of
different types of entities (e.g. species) in an ecological system (e.g. community) where birth
and death rates are linear, and thus entities can be considered as weakly interacting (e.g., a
neutral ecology where species are equivalent in fitness, [26]) due to the operation of fitness
equalizing mechanisms associated to tradeoffs and incompatible optima. These mechanisms
render neutrality as a good approximation, even if entities weakly interact, as envisioned in
the Red Queen and in the Neutral Theory of Ecology [57, 26, 27, 28]. In particular, using the
neutrality approximation in a stochastic framework, we showed in [33] that a one-dimensional
diffusion approximation for the frequency or proportional abundance of species in a community
admits a Beta invariant distribution (the Proportional Species Abundance Distribution), the
Date: Received: date / Accepted: date.
1
arXiv:2310.02809v2 [math.PR] 13 Nov 2024
2 VIDELA, TEJO, QUIÑINAO, MARQUET, AND REBOLLEDO
same as found, for instance, for gene frequencies in partially isolated populations [63, 31], which
generalizes to a Dirichlet in the multidimensional case.
To model an early ecological system where to shed light on the importance of neutrality in
early ecologies, we use a quintessential model for replicating entities, which has been applied to
the modeling of polynucleotides in a container, the dynamics of gene frequencies, selection (Price
equation), relative densities of interacting populations (Lotka-Volterra model), the frequency of
different strategies in a population (game theoretical models), and adaptive dynamics [55, 49,
25, 43]. One of the most basic and general equations for all these dynamics is the deterministic
equation introduced by Taylor and Jonker [55] and dubbed replicator equation by Schuster and
Sigmund [55, 49].
Thus, in this article, any given entity in our primordial ecological system will be represented
by a replicator, and we will follow its proportional abundance through time. Now, since it can
be argued that in the primordial soup, many entities shared a common habitat in this article
we will extend the simple replicator dynamics to the setup of a large number of self-reproducing
entities (community), constituting in this way an interactive set. In the setting we have chosen,
the intraspecific dynamics of each singular replicator will be affected, in the form of a mean-field
effect, by the dynamics of the other replicators in the system.
Is neutrality still plausible to emerge in this larger setting? and if so, What is its role? The
sections that follow are devoted to shed light on these questions.
1.1. Preliminaries. The main motivation of Taylor and Jonker’s replicator model ([55]) was
to unify many game-theoretic models that pursue the study of evolutionary stable strategies
(ESS) in the idealized dynamics of animal conflicts as introduced by Maynard Smith and Price
[52]. Recall that Taylor and Jonker’s replicator is a dynamical system (x(t) : t0) taking
values on the (d1)-dimensional simplex:
Sd:= {xRd:xi0,
d
X
i=1
xi= 1},
obeying the ordinary differential equation:
dx
dt =x(Ax x, Ax1),(1)
for some initial condition x0Sd. Here, stands for the Hadamard product (entry-wise), Ais
a suitable d×dreal matrix termed the payoff matrix, and 1stands for the vector of dimensions
dwith all its entries equal to 1. Equation (1) is intended to model the dynamics of the
proportions (xi:i= 1, . . . , d)of individuals playing contesting strategies in (the continuous-
time approximation of) a multi-round, two-player game: if each player can choose among
strategies 1,2, . . . , d, the entry Aij is interpreted as the payoff from using strategy iagainst
an individual using strategy j. The term x, Ax1modulates the fitness variation in such
a way that the fitness of a player of type iincreases if the payoff is greater than the average
payoff of the community (see [25, 20, 23] for a thorough study of these systems in the context
of evolutionary game theory).
Many generalizations of the original replicator dynamics can be considered. For example,
we can obtain the generalized replicator dynamics:
dx
dt =x(F(x) x, F (x)1),
INTERACTING REPLICATOR DYNAMICS 3
for some suitable instantaneous fitness function F:Sd7→ Rd. In this form, replicator dynam-
ics has become a general model for the evolution of populations under frequency-dependent
selection, “flexible enough to cover a great deal of evolution models, suggesting a unifying view
of replicator selection from the primordial soup up to animal societies” ([49]). The mathemat-
ical analysis of general replicator equations has been extensive ([24], [25], [21], [61], [18], [9],
[41], [42], [10]), [8], testifying to its fundamental role in the development of evolutionary game
theory.
Stochastic versions of the replicator equation have a long tradition. From the initial efforts
of Foster and Young ([14]) and Fudenberg and Harris ([17]) in exploring the replicator equation
in a stochastic differential equation form (see also [1], [46]), the analysis of stochastic replicator
equations have become increasingly complex, going beyond the case of only two pure strategies
([29]) to replicator dynamics with Stratonovich-type perturbations ([30]), to the analysis of
stochastic persistence ([3]), variations in game dynamics with a potential function ([2]) to the
inclusion of revision protocols such as imitation ([47], [37]).
It must be said that the interaction between the main system and the noise can be interpreted
in many different ways. For example, Fudenberg and Harris’s version of the stochastic replicator
can be obtained from the following considerations regarding the instantaneous fitness (see [17]).
Let Rd
++ be the cone of strictly positive real d-tuples, let Yt= (Y(1)
t, Y (2)
t, . . . , Y (d)
t)Rd
++
be the vector of abundances at time t0of the dtypes in a focal population, and let:
Xt=Yt
Pd
i=1 Y(i)
t
Sd,
be the relative proportions at time tof these types. We prescribe a stochastic evolution of the
abundances through the equation.
dYt=Yt(AXtdt + ΣdWt),(2)
where Σ := diag(σ1, σ2, . . . , σd)is a diagonal non-singular matrix and W= (W1, . . . , Wd)is a
standard d-dimensional Wiener process defined on a probability space (Ω,F,P)and adapted
to a right-continuous and augmented filtration F:= (Ft:t0). The rationale behind this
evolution is easy to grasp: the instantaneous fitness of type iis the payoff associated with the
strategy of playing iat the density level X,plus a random perturbation. In (2) the stochastic
term is generally interpreted as environmental fluctuations due, for example, to high-frequency
variations of the abiotic or biotic factors that underlie the ecological dynamics. Depending on
the time scale of the phenomenon we are interested in, these variations can come from different
sources: temperature fluctuations, changes in resource availability, among others.
From equation (2), it is easy to show that the dynamics of the proportion obeys the SDE:
dXt=Xt(˜
AXtDXt,˜
AXtE1)dt +
d
X
i=1
Xtei Xt,Σei1)dW (i)
t,(3)
where ˜
A=AΣΣand eiis the i-th vector of the canonical basis of Rd.
Hofbauer and Imhof [22] studied equation (3), with a focus on the time-averages of the
dynamics. Among other results, the authors show that under some conditions on the payoff
matrix, the process is positive recurrent and its transition probability function admits an in-
variant Dirichlet distribution. Interestingly, this holds when the sum of the off-diagonal payoffs
and the diagonal obey a simple linear relation (this is their Corollary 3.12). This symmetry
can be regarded as equivalent to a type of neutrality, as described in the first paragraphs of this
introduction, and in fact, induces the emergence of a neutral ecology in replicator dynamics,
4 VIDELA, TEJO, QUIÑINAO, MARQUET, AND REBOLLEDO
which provides a remarkable connection to the existing neutral theory in community ecology
[26, 33]. Thus, in this contribution we will use the replicator equation in a stochastic framework
as a vehicle to understand the emergence of primordial ecologies and test for the emergence and
generality of the fitness equivalence principle outlined in [22, 33], in the context of a community
of interacting replicators, as explained next.
1.2. Our model: interacting replicators. For measurable functions F:Sd7→ Rd, write:
ΠTF(x) := x(F(x) x, F (x)1),
and extend the definition to matrix-valued functions by tensorisation, namely: if F= (F1, . . . , Fm)
where Fi:Sd7→ Rdis bounded and measurable, then ΠTFis the d×mmatrix whose j-th
column is ΠTFj(x). With this, equation (3) can be written as:
dXt= ΠTΦ(Xt)dt + ΠTΨ(Xt)dWt,(4)
for a suitable vector field Φand a suitable matrix-valued function Ψ; here Wis a standard
d-dimensional Brownian motion on a filtered probability space. This is the basic replicator
equation we will be working with in this contribution.
Now, to capture the interaction of one replicator with other of the same type, we assume that
a smooth function Υ : Sd×Sd7→ Rdis given. The notation ΠTΥ(x,y)stands for the action
of ΠTon the first variable. Let be given a natural number N, and consider the (Sd)N-valued
stochastic process defined via
dX(N;i)
t= ΠTΦ(X(N;i)
t)dt
+1
N
N
X
j=1
ΠTΥ(X(N;i)
t,X(N;j)
t)dt + ΠTΨ(X(N;i)
t)dW(i)
t,(5)
for i= 1,2, . . . , N . Here, (W(i), i = 1,2, . . . , N )is a family of independent F-adapted, d-
dimensional standard Wiener processes carrying the local noises (i.e., noise associated to each
species represented by a single replicator).
To fix ideas, consider the case d= 2,Φlinear and Ψa constant matrix, and assume, for
instance, that Υhas the form:
Υ(x,y) = δy1Ex,E= 0 1
1 0!
where δis a small parameter. We will return to this example in Section 5, and so it seems
just right to use it to illustrate the meaning of the equations above. Indeed, in this case, we
can provide an intuitive interpretation of equation (5): the payoff obtained from using the first
strategy against a player using the second one increases proportionally to the overall average
frequency of the usage of the first strategy amongst the family of replicators. Thus, at least
intuitively, in this system of interacting replicators, the first strategy becomes more resilient,
and thus one should expect higher average usage of the first strategy as compared to the
situation with no interaction. Of course, the overall effect must take account of the particular
form of the intrinsic payoff matrix (the diagonal elements could restrain a net bias to the use of
the first strategy) and the noise part (if it is large, randomness may dominate the dynamics),
but as we will see later, under appropriate assumptions, the intuition above is not far from the
actual long-term picture.
INTERACTING REPLICATOR DYNAMICS 5
Remark 1. In this article, we are considering the case of interacting replicators of just one
type, which could be akin to traditional species that are either phylogenetically and functionally
close and thus share a similar niche, trophic level, and interactions within communities, as
envisioned in the principle of niche conservatism [62]. As we will see below, this will translate
in that, for large N, the individual replicators behave as statistically identical entities. The
multi-type model (which would amount to mean-field interactions of N1replicators of type 1,N2
replicators of type 2, etc.; see for instance models of this type associated to neurosciences [13],
social sciences [7] or statistical mechanics [6]) is technically more involved and represents a
reasonable second step in our research program.
Systems analogous to equation (5) are generically referred to as mean-field interacting parti-
cle systems, and one of the main problems associated with them is the determination of condi-
tions under which, in the thermodynamic limit N , the assembly of particles (in our case,
replicators) behave as if they were independent. This is the propagation of chaos property, whose
precise content will be given later. For the time being, assume that we are interested in the
study of the system (5) on a finite time horizon and for large values of N. Write µ(N)
tfor the em-
pirical probability measure induced on the simplex by
X(N)
t:= (X(N;i)
t,X(N;2)
t,...,X(N;N)
t),
namely:
µ(N)
t(A) := 1
N
N
X
i=1
δX(N;i)
t(A).
With a similar notation, µ(N)
[0,t](A)stands for the random probability measure on C([0, t]; Sd)
given by:
µ(N)(A) = µ(N)
[0,t](A) := 1
N
N
X
i=1
δX(N;i)
[0,t]
(A).(6)
The following convention will be used in the sequel. For an arbitrary metric space Eand a
measurable function G:Sd×Sd7→ E, we will write:
G(x, µ) := ZSd
G(x,y)µ(dy),
and thus, we can regard Gas a map Sd× P(Sd)7→ E. With this convention, equation (5) can
be written as:
dX(N;i)
t= ΠTΦ(X(N;i)
t)dt + ΠTΥ(X(N;i)
t, µ(N)
t)dt(7)
+ ΠTΨ(X(N;i)
t)dW(i)
t, i = 1,2, . . . , N.
Assume that we know that for each tin a finite time horizon, the random probability measures
(µ(N)
t:N1) converge (almost surely in the weak topology of probability measures) to a
certain deterministic probability measure, µtsay. Then, at least intuitively, we should have
that for large values of N, the dynamics of a typical particle should be governed by:
dΞt= ΠTΦ(Ξt)dt + ΠTΥ(Ξt, µt)dt + ΠTΨ(Ξt)dWt,(8)
and, following a Glivenko-Cantelli-like reasoning, it should be apparent that µtis nothing but
the law of Ξt. Equation (8) is of the McKean-Vlasov type, and its solutions (whenever they
exist) induce a paradigmatic example of a non-linear Markov process. Here, the term non-linear
has nothing to do with the algebraic degree of the unknown Ξin (8) but rather with the fact
that the random dynamics depends on the very distribution of the Markov process; in turn, it
6 VIDELA, TEJO, QUIÑINAO, MARQUET, AND REBOLLEDO
translates into a non-linearity at the level of the evolution equation associated to the dynamics.
There is a rich literature on McKean-Vlasov SDE, with the connection with nonlinear PDE
being an important research topic (see the classical [35] for the motivations underlying this
kind of process).
The aim of this article is to study the link between the mean-field system (5) and the
McKean-Vlasov equation (8); to study the persistence (in a sense to be defined below) for the
mean-field system and to prove existence and uniqueness of non-trivial invariant measures for
the McKean-Vlasov equation in some simple cases. All these issues are treated with an eye
on the problem posed at the opening of the introduction, namely the emergence and role of
neutrality in large interacting sets of biological entities. The organization of the paper reflects
closely this sequence of problems. In Section 2 we present all the mathematical preliminaries
that allow us to put the models on firm grounds, and we describe our assumptions and main
results. Then, we prove the analytical properties of the processes and provide a result of
the propagation of chaos linking the N-replicator dynamics and the McKean-Vlasov system
(Section 3). Then, for the N-replicator systems we investigate conditions that guarantee the
persistence of the whole set of replicators (Section 4). For the non-linear replicator, we prove
the existence and uniqueness of invariant measures for a particular type of interaction functions
and noise (Section 5). Section 6 shows some simulations and numerical results related to the
model treated in Section 5. Finally, Section 7 discuss our results and points out open questions
regarding our research.
2. Main results
Here we present our main definitions and assumptions and summarise the main results.
2.1. Notation and definitions. Notation
Rd
+and Rd
++ stand for the cones of real d-dimensional vectors with positive and strictly
positive entries, respectively. A generic element of Rd
+or Rd
++ will be denoted by
bold letters x,y,z, etc., and when we need to stress its components, we will write
x= (x1, x2, . . . , xd).
For a metric space (E , ρ)the notation B(E)stands for both, its Borel σ-field and the
algebra of real-valued Borel measurable functions. Also, we write P(E)for the set of
probability measures on (E, B(E)).
If Ais a subset of a topological space we write int (A),A, and ∂A to denote its interior,
closure, and boundary, respectively. In particular, int(Sd) = {xSd:Qd
i=1 xi= 0}
and Sd={xSd:Qd
i=1 xi= 0}.
If A, B are subsets of a topological space, the notation ABmeans that Ais compactly
contained in B:AB, and Ais compact.
For a function f:EE, with (E , d)and (E, d)two metric spaces, the notation
Lip(f)stands for the best Lipschitz constant of f, namely: Lip(f)Kif and only if
d(f(x), f (y)) Kd(x, y)for every x, y E.
If Eis a (regular enough) subset of Rd, we denote by C(E)the space of continuous
real-valued functions on E. Analogously, Ck(E)stands for the k-times continuously
differentiable real-valued functions on E.
INTERACTING REPLICATOR DYNAMICS 7
If Eis a compact metric space and A:E7→ RdRmis a continuous map, we define
the (uniform) Frobenius norm through:
|||A||| := sup
xEqTr{A(x)A(x)}.
For a Banach space Ewe denote with ⟨·,·⟩ the duality product of Ewith E. In
particular, if E=Rd, the notation x,ystands for the usual inner product of vector
of Rd.
For f Bb(E)and µaσ-finite measure on E, we write µf to denote the integral
REf(x)µ(dx). Regarding the point above, we also write µ, f with the same meaning.
For probability measures µ, ν on a measurable space (E , E), the notation dTV (µ, ν)
stands for total variation distance between µand ν, namely:
dT V (µ, ν) = sup
A∈E
|µ(A)ν(A)|.
If E=Rd, then Wasp(µ, ν )stands for the Wasserstein distance with respect to the
underlying p-th vector norm, namely:
Wasp(µ, ν ) = inf
πCouplings(µ,ν)ZRd×Rd
xypπ(dx, dy)1/p
,
whenever the above quantity exists. Here, Couplings(µ, ν )is the family of probability
measure on the product space Rd×Rdthat have first and second marginals µand ν,
respectively.
2.2. The stochastic replicator. The structure of the SDEs (and later, McKean-Vlasov type
equations) that we are going to consider take their general form from a geometric consideration.
Recall that the Shahshahani metric on Rd
++ is given by the metric tensor:
(9) si,j (x) = x1
xi
δi,j ,(xE, i, j = 1, . . . , d).
(see [50] for further details). The gradient of a C1-function U:Rd\ {0} Ris transformed by
the Shahshahani metric into S(x)1U(x), where S1(x)is the inverse of the diagonal matrix
S(x)whose entries are si,j, that is
S(x)1U(x) =
x1
x1
∂x1U(x)
.
.
.
xd
x1
∂xdU(x)
=1
x1
x U(x),
where denotes Hadamard product. In particular, for xSdone gets S(x)1U(x) =
x U(x). This gradient projected onto the tangent bundle at x(hereafter, Tx=T) gives:
(10) x U(x) x U(x),1x=x(U(x) x,U(x)1).
Thus, at least if Ain equation (1) is a symmetric payoff matrix, the original deterministic
replicator has the form of a gradient-type dynamics on the simplex.
Notice, however, that the right-hand side of equation (10) makes sense if we replace Uby
any Rd-valued function F. From now on, write:
ΠTF(x) := x(F(x) F(x),x1),
and extend in an obvious way the definition for matrix-valued functions G:Rd7→ RdRm.
8 VIDELA, TEJO, QUIÑINAO, MARQUET, AND REBOLLEDO
Let Φ : Sd7→ Rdand Ψ : Sd7→ RdRdbe given. We assume that, at least, both
Φand Ψare C1(Sd)functions. This allows us to affirm the existence of a uniquely defined
matrix-valued function AΨsuch that ΠTΨ(x) = xAΨ(x)for every xSd. Consider the
d-dimensional SDE given in (4),with suitable initial conditions on Sd. We note first that this
equation defines a Sd-valued process that can be obtained via the projection on the simplex of
a simple model of population dynamics (see the Appendix for the details of this derivation),
and thus the paths remain in Sdby definition. Note also that by our assumptions, both drift
and diffusion components are Lipschitz on Sd, and thus it is well known that there exists
a unique (pathwise) solution to the SDE (4) which defines a well-posed Sd-valued Markov
process. Let (P,Ψ)
t:t0) be the semigroup associated to the Markov process (Xt:t0)
associated to the characteristics ,Ψ). As usual, we consider the semigroup acting from the
left on the space of bounded Borel function and acting from the right on Borel measures on
Sd. Let Bbe a Borel set of Sd. It is a routine fact that (P,Ψ)
t:t0) is a Feller semigroup.
By compactness, it admits invariant measures [11, Theorem 12.39]. Recall that a probability
measure µ P(B)is invariant for the semigroup (equivalently, invariant for the process) if
µP ,Ψ)
t=µfor every t0, and in this case we write µ Pinv,Ψ; B). Let Perg,Ψ; B)be
the subset of Pinv,Ψ; B)of extremal invariant probability measures supported on B(that
is, ergodic probability measures that put no mass on Sd\B).
It is well known that for well-behaved Markov processes (for example, for processes whose
semigroups enjoy the strong Feller property), different ergodic measures have disjoint supports.
Thus, in order to obtain the uniqueness of an ergodic probability measure supported on the
interior of the simplex, it seems reasonable to impose conditions on the noise in such a way
that no submanifold of dimension less than d1is kept invariant by the dynamics. So, we will
impose the following non-degeneracy of the noise.
Assumptions 1.
spanTΨ(x)·,j, j = 1, . . . , m) = Tx,(11)
for all xint (Sd).
Following [3], to assess the stochastic persistence of the process X(and later, the stochastic
persistence of interacting replicators), we are going to impose conditions on the natural deter-
ministic process associated with it (namely, the process with no diffusion term). Define the
invasion growth rate:
hi(x) = Φ(x)i x,Φ(x),(12)
and set h:= (h1, h2, . . . , hd). Observe that the i-th invasion rate is just the exponent asso-
ciated to the i-th population in the deterministic dynamics. Now, if µ Pinv,0; Sd)is an
invariant probability measure for the deterministic replicator under which type iis absent, then
µcan be regarded as an equilibrium distribution for a given strict population with strategies
or types indexed by I {1,2, . . . , d}with i /I. Consider the following condition.
For every µ Perg ,0; Sd):
max
i=1,...,d µhi>0.
The biological interpretation of this condition is just the requirement that any strict popula-
tion of types in ecological equilibrium (for the deterministic dynamics) allows for the invasion
INTERACTING REPLICATOR DYNAMICS 9
of at least one of the absent types. It turns out that the above condition is equivalent to the
following assumption (see e.g. [48]).
Assumptions 2. There exists p= (p1, p2, . . . , pd)a vector of strictly positive components
such that:
inf
µ∈Perg ,0;Sd)µp,h=: ρ > 0.
The next theorem is just Theorem 1 in [3], which we state here for future comparison.
Theorem 2. Assume that Φis C2(Sd), and Assumptions 1 and 2 hold. Then, for every r > 0,
there exists ε=ε(r)>0such that for every Ψof class C3(Sd)such that |||AΨ||| < ε the
following holds:
(1) Pinv(int(Sd)) is a singleton, i.e., there exists a unique invariant probability measure µ
supported on the interior of the simplex for the system (4). Moreover µis absolutely-
continuous w.r.t. Lebesgue measure, and the function x7→ dist(x, Sd)ris µ-integrable.
(2) There exist positive constants C, α such that:
dT V (δxPt, µ)Ceαt
dist (x, Sd)r,xint(Sd).
In this article, we extend the previous result to the setup of a family of replicators interacting
in a mean-field regime. The next subsection provides the details.
2.3. Mean-field interacting replicators and McKean-Vlasov evolutionary games. As
a first step, we extend the definition of our parameters to account for the interaction among
replicators. Let be given a bounded measurable map Υ : Sd×Sd7→ Rd. Observe that:
ΠTΥ(x, µ) = µΠTΥ(x,·) = ΠTµΥ(x,·).
We assume the following regularity.
Assumptions 3. Υis a C1map.
In what follows, sometimes we will need to refer to the joint effect of both the endogenous
and the interaction functions. To that aim, we define:
b:= ΠTΦ+ΠTΥ,
which is continuous as a map from Sd× P(Sd), where the relevant topology in P(Sd)is the
weak topology of probability measures. With this notation, equation (7) becomes
(13) dX(N;i)
t=b(X(N;i)
t, µ(N)
t)dt + ΠTΨ(X(N;i)
t)dW(i)
t, i = 1, . . . N,
and McKean-Vlasov equation (8) becomes
dΞt=bt, µt)dt + ΠTΨ(Ξt)dWt,(14)
where µt:= Lawt).
Remark 3. Our hypotheses entail that b, when seen as a map from Sd× P(Sd), is Lipschitz
continuous for any Wasserstein distance in the measure variable, in the sense that for every
p1there exists a constant C(p, d)such that for every x,ySdand µ, ν P(Sd):
b(x, µ)b(y, ν)pC(p, d)xyp+Wasp(µ, ν).
The following result is proved in the next section.
10 VIDELA, TEJO, QUIÑINAO, MARQUET, AND REBOLLEDO
Proposition 4. Under the above assumptions, Equation (14) has a unique pathwise (hence
strong) solution.
Now we are in a position to settle the main results of this paper. We start with a proposition
that links the interacting particle system and the McKean-Vlasov replicator.
Theorem 5. Assume Φis C2(Sd), and Assumptions 1 and 3 hold. Let ˜µbe a probability
measure on the simplex, and consider the N-particle system (13) started from ˜µN. Consider
also the McKean-Vlasov replicator (14) started from ˜µ. Then the propagation-of-chaos property
holds in the following sense: for any positive time t, the sequence of random probability mea-
sures (µ(N)
[0,t])N1, converges almost surely in the topology of the weak convergence of probability
measures, to the deterministic probability measure µ[0,t]=Law[0,t]).
It is well known that the thesis of the above theorem is equivalent to the fact that, in the
limit N , a sort of asymptotic independence of the particles holds (see Proposition 2.2,
(i), at [53]). More precisely, under the above conditions, for any T > 0and any finite set
of functions gk Cb(C([0, T ]; Sd)), with E(N)the expectation operator associated to the N-
replicator system, and E(MV )the corresponding expectation associated to the McKean-Vlasov
replicator, we have:
lim
N→∞ E(N)
˜µN m
Y
k=1
gk(X(N;k)
[0,t])!=
m
Y
k=1
E(MV )
˜µ(gk[0,t])).
Let P(N)be the semigroup associated with the N-particle system. Let PN
inv (resp. PN
erg )be
the set of invariant (resp. ergodic) probability measures for P(N). The next theorem, which
extends Theorem 2, is proved in Section 4. It can be paraphrased as: if the typical replicator
of the assemblage satisfies Assumption 2, then the whole community is persistent.
Theorem 6. Assume that Φis C2(Sd), and Assumptions 1, 2, 3 hold. Fix N1,r > 0. Then
there exists ε:= ε(r)>0such that for any Ψof class C2(Sd)such that Υ+|||AΨ||| < ε, the
following holds:
(1) PN
inv(int (Sd)N)is a singleton, i.e., the N-replicator systems admits a unique invariant
probability measure µsupported on the interior of (Sd)N. Moreover µis absolutely-
continuous w.r.t. Lebesgue measure, and the function x7→ dist (x, (Sd)N)r/N is µ-
integrable.
(2) There exist positive constants C, α such that:
dT V (δxP(N)
t, µ)Ceαt
dist (x, (Sd)N)r/N ,xint((Sd)N).
Finally, let us consider the case of Φ(x) = Axand Ψ(x) = σI, and assume that for the
constant σand the payoff matrix Asatisfy the conditions:
[C1] For every i=j:
aij +aji aii ajj =σ2.
[C2] For the matrix Awith entries a
ij =aij 1
2σ2, there exists a vector α= (α1, . . . , αd)
of strictly positive entries summing up to 1such that Aαis a multiple of 1.
In the case of small σ, condition [C1] implies the existence of fitness equivalence (the diago-
nal and off-diagonal fitness payoffs are similar to each other), in the sense that any perturbation
of the fitness payoffs that results in that intraspecific and interspecific interactions compensate
INTERACTING REPLICATOR DYNAMICS 11
each other, will preserve the condition of neutrality. Notice that the condition of neutrality is
assumed in the neutral theory of ecology ([26]) but here it is derived as a condition for system
persistence. Further, a small σalso ensures the coexistence of strategies, such that the pro-
cess spends little time near the pure strategy state, and, as shown below, the process admits
a unique invariant distribution of Dirichlet type. Condition [C2] implies the existence of an
interior Nash equilibrium.
Consider the particular case where Υis given by:
Υ(x, µ) = δF (µ)x,
where F:P(Sd)7→ RdRdis a Lipschitz function with respect to the Was1distance that
ranges in the space of skew-symmetric real matrices and δRis a parameter that represents
the interaction strength of a replicator with the rest of the alike replicators in the community.
Observe that, in this case, the condition [C1] is preserved, in the sense that if Asatisfies this
condition, then for every positive time the payoff matrix associated with the dynamics of the
McKean-Vlasov replicator with characteristics ,Υ,Ψ) also satisfies the condition. Thus, if
we regard [C1] as a proxy for neutrality, the prescription for Υcan be seen as a definition that
preserves neutrality in the extended set up of mean-field replicator entities. In Section 5 we
prove the existence of invariant distribution for a class of McKean-Vlasov replicators associated
with these parameters.
Theorem 7. Assume Asatisfies [C1] and [C2]. Then there exists an interaction strength
δ0>0such that for every δ < δ0, the stochastic McKean-Vlasov replicator with Φ(x) = ˜
Ax
and Ψ(x) = σId×dconstant admits a unique invariant probability measure putting no mass on
the boundary of the simplex. Moreover, this invariant measure is of the Dirichlet family.
Although the above result states the existence of unique invariant measures for the above
class of McKean-Vlasov replicator, by now we have no rigorous arguments for the ergodicity of
the system. So, in Section 6 we provide statistical evidence to support this fact (at least in the
simple setup of the S2-Mckean-Vlasov replicator). The evidence is then confronted to a test
that allows us to accept the hypothesis that, for δ > 0small enough, every initial condition is
attracted to the Beta(s, 1s)distribution, with s=a12 a22
σ2δ.
3. Propagation of chaos for the mean-field replicators
In this section, we settle some auxiliary results regarding the well-posedness of the equa-
tions we are dealing with, and we prove the propagation of chaos for the N-replicator system
(Theorem 5).
3.1. Elementary results. We start with a proposition that regards the existence and unique-
ness of solutions to the nonlinear McKean-Vlasov equation. Its proof is relatively standard,
and consequently, it is deferred to the Apprendix A.2.
Proposition 8. Under Assumptions 1 and 3, equation (14) has a unique pathwise (hence
strong) solution.
Observe that the process Ξis not a Markov process. Indeed, the family of maps (Ps:s0)
acting on Bb(Sd)defined via
Psf(x) := Ex(fs)),
12 VIDELA, TEJO, QUIÑINAO, MARQUET, AND REBOLLEDO
do not determine a semigroup of linear operators, in neat contrast to the Markovian setup.
However, the semigroup property still holds in the dual sense. Define the one-parameter family
of operators:
Qt:P(Sd)7→ P(Sd)(15)
given by the prescription: Qtµ=Lawµt), the law of Ξtwhen started from an initial condition
distributed as µ.
Proposition 9. The family of operators (Qt:t0) enjoys the semigroup property:
Qt+s=QtQs,(16)
for all s, t 0.
Remark 10. Still, the operators Qsare non-linear: for µ, ν P(Sd),λ[0,1], in general:
(λµ + (1 λ)ν)Qt=λµQt+ (1 λ)νQt.
Proof. Let µ P(Sd), and t, s 0. Put µt:= Qtµ. Consider random variables Ξ0and Ξ1
distributed as µand µtrespectively, and the processes 0)
t+s:s0) and 1)
s:s0). By
(14), for s0, on the one hand we have:
Ξ(µ)
t+s
Law
= Ξ0)
t+s= Ξ0+Zt
0
˜
E(b0)
r,˜
Ξ0)
r))dr +Zt
0
ΠTΨ(Ξ0)
r)dWr
+Zs
0
˜
E(b0)
r+t,˜
Ξ0)
r+t))dr +Zs
0
ΠTΨ(Ξ0)
t+r)dWt+r
Law
= Ξ0+Zt
0
˜
E(b0)
r,˜
Ξ0)
r))dr +Zt
0
ΠTΨ(Ξ0)
r)dWr
+Zs
0
˜
E(b0)
r+t,˜
Ξ0)
r+t))dr +Zs
0
ΠTΨ(Ξ0)
t+r)d˜
Wr.
where ˜
Wis a Wiener process independent of (Wr: 0 rt)and ˜
Estands for expectation
with respect to the law of an independent copy of the process Ξ0). Observe that the sum of
the first three terms in the last equation is distributed as Ξ1and is independent of ˜
W. Writing:
Y·:= Ξ0)
t+·, the last line becomes:
Law
= Ξ1+Zs
0
˜
E(b(Yr,˜
Yr))dr +Zs
0
ΠTΨ(Yr)dWr
=Ys
Law
= Ξ(µt)
s,
where the first-to-last line follows from pathwise uniqueness of the solutions of (14) (Proposition
8), and the last line is obvious. Our claim follows.
For future reference, we establish a simple result. For the interested reader, we present the
proof in Appendix A.3.
Lemma 11. For every Ξ0int(Sd):
P(MV )
Ξ0min
i=1,...,d Ξ(i)
t>0for all t0= 1.
INTERACTING REPLICATOR DYNAMICS 13
An analogous result holds for the N- particle system, namely: for every
x0int ((Sd)N),
P(N)
x0min
j=1,...,N X(N;j)
t>0for all t0= 1.
In words: with probability 1, in finite time there is no absorption on the boundary of Sd(for
the McKean-Vlasov replicator) nor on the boundary of (Sd)N(for the N-particle system).
As a direct result of Lemma 11, we have that for any xint (Sd)and t > 0
Qtδx(Sd)=0,
i.e., the trajectories of the solution of (14) do not hit the boundary of the simplex in finite time.
The following result regards the existence of a density of the law of the nonlinear process.
Proposition 12. For any xint(Sd)and t > 0,QtδxLebd1.
Proof. For t0and ySd, define:
˜
b(y, t) := b(y,Qtδx),
and consider the following diffusion with time-dependent coefficients:
Ξ(xx)
t=x+Zt
0
˜
b(xx)
s, s)ds +Zt
0
ΠTΨ(Ξ(xx
s))dWs.
Observe that the coefficients are Lipschitz-continuous and bounded, and moreover, on account
of Assumption 1, for every compact Dint (Sd):
z,ΠTΨ(yTΨ(y)z> εz2,
for a certain constant ε > 0, for every y D and z T . Now, for any xint(Sd)consider:
τD:= inf{t0:Ξ(xx)
t/ D,x D},
and define the finite measure on Sd:
mD
t(A) := P(xx)
tA, t τD).
By a classical piece of the theory of parabolic PDEs with uniformly elliptic coefficients (see
for example Chapter 6 of [16] together with the representation given by Theorem 3.46, pag.
207, of [44]), we deduce that the measure mD
t(·)is absolutely-continuous with respect to the
Lebesgue measure.
Consider next a sequence of compact subsets {Dϵ}ϵin int (Sd), such that Dϵ Dϵif ϵ<ϵ,
and limϵ0Dϵ=int (Sd). Analogously as we have done with Dabove, define τDϵand mDϵ
t(·).
Let Abe a measurable subset of int (Sd)with Lebd1(A) = 0. Then mDϵ
t(A) = 0 for every
ϵ > 0. Since the trajectories of the solution of (14) do not hit the boundary of Sdin finite time,
we have that mDϵ
t(A)δxQt(A)as ϵ0. This proves the claim.
Remark 13. The result still holds for the N-replicator systems, i.e., for every initial condition
in int ((Sd)N)the law of
Xtis absolutely continuous w.r.t. the N-fold product of the Lebesgue
measure.
14 VIDELA, TEJO, QUIÑINAO, MARQUET, AND REBOLLEDO
3.2. Convergence of the empirical laws. Fix t > 0. Let µ(N)be the random probability
measure on C([0, t]; Sd)given by (6). Now we will prove that (µ(N))Nalmost surely converges,
in the weak sense, to the deterministic limit µ:= Law[0,t]). Recall that this means: if
mN P(P(C([0, t]; Sd))) is the law of the random measure µ(N), we aim to prove that for every
bounded continuous function F Cb(P(C([0, t]; Sd))), we have:
ZP(C([0,t];Sd))
mN()F(ν)F(µ),
as Ngoes to infinity. If we write E(N)for the expectation operator associated with the N-
replicator system, the above convergence is equivalent to asking that:
E(N)((µ(N)fµf)2)0,
as N , for every f C(C([0, t]; Sd). In turn, it can be proved that this will follow if we are
able to prove that the family of probability measures E(N)(µ(N)) := Law(X(N;1)
[0,t])converges in
P(C([0, t]; Sd)(see [54], pages 177 et seq.; see also [36], pages 66 et seq.). So, this is what we
will do now.
Assume that the N-particle system (X(N;1)
·, ..., X(N;N)
·)in (13) is issued from independent
and identically distributed conditions (X(N;1)
0, ..., X(N;N)
0), and are driven by independent d-
dimensional standard Brownian motions (W(1)
·, ..., W(N)
·). Notice that this implies that the
law of (X(N;1)
·, ..., X(N;N)
·)is symmetric.
Theorem 14. (Propagation of Chaos). Let (i):i= 1, . . . , N )be a family of independent
copies of the non-linear process solution of (14), such that Ξ(i)
0=X(N;i)
0for every i= 1, . . . , N
almost surely, and both X(N;i)
·and Ξ(i)
·are driven by indistinguishable Brownian motions.
Then, under the above assumptions, for any finite time horizon [0, S]there exists a positive
constant K=K(S, d)such that for every i= 1, . . . , N:
E sup
t[0,S]
X(N;i)
tΞ(i)
t
2!K
N.
Proof. Let X(N;i)
·and Ξ(i)
·be the processes specified above, and let S > 0. Then, for any
t[0, S]we have:
X(N;i)
tΞ(i)
t
22SZt
0
b(X(N;i)
s, µ(N)
s)b(i)
s, µs)
2ds
+ 2
Zt
0
TΨ(X(N;i)
s)ΠTΨ(Ξ(i)
s)]dW(i)
s
2
.
Now, for the first term of the right-hand side, by the Lipschitz condition of b, there exists a
K1=K1(S, d)>0such that:
Zt
0
b(X(N;i)
s, µ(N)
s)b(i)
s, µs)
2ds K1Zt
0
X(N;i)
sΞ(i)
s
2ds
+K1Zt
0
1
N
N
X
j=1
X(N;j)
sZSd
s(dx)
2
ds.
INTERACTING REPLICATOR DYNAMICS 15
On the other hand, notice that the term R·
0TΨ(X(N;i)
s)ΠTΨ(Ξ(i)
s)]dW(i)
sis an (Ft:t0)
square-integrable martingale, and thus by Doob’s Maximal inequality we have that
E sup
t[0,S]
Zt
0
TΨ(X(N;i)
s)ΠTΨ(Ξ(i)
s)]dW(i)
s
2!
4E
ZS
0
TΨ(X(N;i)
s)ΠTΨ(Ξ(i)
s)]dW(i)
s
2
.
By Itô’s isometry and the Lipschitz assumption on ΠTΨ(x), there exists a constant K2=
K2(S, d)>0such that:
E
ZS
0
TΨ(X(N;i)
s)ΠTΨ(Ξ(i)
s)]dW(i)
s
2
K2E ZS
0
X(N;i)
sΞ(i)
s
2ds!.
Therefore, we have that:
E sup
t[0,S]
X(N;i)
tΞ(i)
t
2!K3ZS
0
Esup
us
X(N;i)
uΞ(i)
u
2ds
+K3Zt
0
E
sup
us
1
N
N
X
j=1
X(N;j)
uZSd
u(dx)
2
ds,
for some a constant K3=K3(S, K1, K2).
Recall that we have assumed that (1)
·,...,Ξ(N)
·)are Ni.i.d. copies obeying (14) such
that for all j= 1, ..., N ,Ξ(j)
0=X(N;j)
0almost surely, and Ξ(j)
·and X(N;j)
·are driven by
indistinguishable Brownian motions. Then:
Zt
0
E
sup
us
1
N
N
X
j=1
X(N;j)
uZSd
u(dx)
2
ds
2
N2ZS
0
E
sup
us
N
X
j=1
(X(N;j)
uΞ(j)
u)
2
ds
+2
N2ZS
0
E
sup
us
N
X
j=1
(j)
uZSd
u(dx))
2
ds.
The last term above is O(N1)by the Law of Large Numbers and, on the other hand, we have
that:
1
N2ZS
0
E
sup
us
N
X
j=1
(X(N;j)
uΞ(j)
u)
2
ds ZS
0
Esup
us
X(N;i)
uΞ(i)
u
2ds,
16 VIDELA, TEJO, QUIÑINAO, MARQUET, AND REBOLLEDO
since, by symmetry, the laws of the (X(N;j)
·Ξ(j)
·)’s are exchangeable. Finally, by applying
Grönwall’s inequality, we obtain:
E sup
t[0,S]
X(N;i)
tΞ(i)
t
2!O(1/N) exp{2K3S},
and therefore the theorem follows.
We deduce from the previous result that the sequence of empirical measures (µ(N))Ncon-
verges to the deterministic limit µ:= Law[0,t]), in the weak sense. This fact plus the property
that the law of the N-replicator (X(N;1)
·, ..., X(N;N)
·)is symmetric (or exchangeable), implies
that, for any g1,g2, ...., gkcontinuous and bounded functions on Sd, with kN, we have:
lim
N→∞ E(N)
µN k
Y
i=1
gk(X(N;σ(i))
·)!=
k
Y
i=1
µ, gi,
where (X(N;σ(1))
·, ..., X(N;σ(N))
·)corresponds to any permutation of the vector (X(N;1)
·, ..., X(N;N)
·)
(see Proposition 2.2.i) in [54] or Proposition 4.2 in [36]). That is, the N-particle system in the
limit N becomes a sequence of independent and identically distributed solutions of the
McKean-Vlasov equation.
4. Strong stochastic persistence for the N-replicator system
In this section we address the problem of persistence of the Markov processes (
X(N)
t:t0).
For N1, the notation (Sd)Nstands for the N-fold product of Sd. The space (Sd)Nis endowed
with the topology of the topological sum. Since Sdis Polish, the same holds for its N-fold sum,
for the metric:
dist N(
x,
y) := max
i
dist (x(i),y(i)),
where the notation dist in the right-hand side stands for the L1distance on the simplex. We
further set:
dist N(
x, (Sd)N) := min
j
dist (x(j), Sd).
Let P(N)be the semigroup associated with the N-replicator system (7). The action of
the generator associated with the semigroup on smooth functions can be expressed as follows.
Given x(1),...,x(N)arbitrary vectors in Rd, we will write
xfor the vector in RdN resulting
from the concatenation, in lexicographic order, of the components of the x(i)’s. For a given
smooth function φ: (Sd)N7→ R, for i= 1, . . . , N and
x(Sd)Nwe denote by φ(i)
xthe map
Sdy7→ φ(x(1),x(2) ,...,y,...,x(N))where yis in the place of the i-th block of order d.
A straightforward computation shows that whenever φis a twice-continuously differentiable
real-valued function, the action of the generator is given by:
INTERACTING REPLICATOR DYNAMICS 17
L(N)φ(
x) =
N
X
i=1 Dφ(i)
x(x(i)),ΠTΦ(x(i))E
+
N
X
i=1 Dφ(i)
x(x(i)),ΠTΥ(x(i), µ
x)E
+1
2
N
X
i=1
Tr ΠTΨ(x(i))∇∇φ(i)
x(x(i)TΨ(x(i)),
where: µ
x=1
NPN
j=1 δx(i), and ∇∇φ(i)
xis the Hessian matrix associated to φ(i)
x.
Definition 15. (see [19],[59]) We say that a (Sd)N-valued, right-continuous Markov process
(Yt:t0) is strongly stochastically persistent (SSP) if there exists a unique invariant
probability measure πsuch that π(((Sd)N)) = 0 and such that for every xint ((Sd)N):
dT V (Px(Yt ·), π(·)) 0,
as t .
Here we will prove that under our running assumptions, (
X(N)
t:t0) is a SSP processes
taking values on (Sd)N. In fact, we will prove an extension of Theorem 2 and, in particular,
we will obtain geometric rates of convergence for both processes.
To this end, we introduce some notation. We will say that xint(Sd)is ε-close to extinction
if there exists iIsuch that xi< ε. For given ε, define:
Ext(ε) := {xis εclose to extinction}.
We write ExtN(ε) := {
x(Sd)N:x(i)Ext(ε)for some i}.
To prove convergence to an invariant measure for (
X(N)
t:t0), we first observe that it is
enough to prove that for every t > 0there is convergence for the chain (
X(N)
n:=
X(N)
tn :n0),
i.e. the chain sampled from
X(N)every ttime units. This claim follows from the Feller
property of the semigroup (P(N)
t:t0) (which can be derived in a rather standard way from
the Lipschitz continuity of the parameters) and from the following general fact about Feller
processes.
Lemma 16. Let (Pt:t0) be the Feller semigroup associated with a Markov process (Yt:t
0) taking values on a compact metric space (E , ρ). Denote by (P
t:t0) the dual semigroup
acting on finite measures. Fix F B(E). Assume that for every s(0,1] there exists a unique
probability measure πson (E , B(E)) with πs(FC) = 0 and such that P
sπs=πs. Then the
process Yhas a unique invariant probability measure πwith π(FC)=0.
Proof. For every t(0,1] Q, the unique invariant measure for P
tis also invariant for P
1.
Hence, πt=π1for every rational t. Then, for general tand any nN, we have:
π1=P
nt
n
π1,
Since Pis Feller and Eis compact, Pis a weakly-continuous semigroup. Hence, sending
nto in the last equation, we obtain π1f=P
tπ1ffor every f C(E). Since the class
of continuous functions is measure-determining on E, we get that π1=P
tπ1, and by the
uniqueness assumption, πt=π1.
18 VIDELA, TEJO, QUIÑINAO, MARQUET, AND REBOLLEDO
Fix t > 0, and let P:= P(N)
t. Consider the following statement.
[H] There exists a continuous function H:int((Sd)N)7→ R+and constants α(0,1)
and C > 0such that:
(1) limε0inf
xExtN(ε)H(x) = .
(2) PH(
x)αH(
x) + Cfor every
xint ((Sd)N).
Observe that [H] entails that for any xint ((Sd)N), any weak limit point, say ν, of the
sequence of probability (δ
xPn)nsatisfies ν(((Sd)N)) = 0. For suppose this is not the case,
and given
xint ((Sd)N), assume that a limit point νof (δ
xPn)nsatisfies ν(((Sd)N)) >2ε
for a given ε > 0. Apply ntimes the second condition of [H] to get:
PnH(
x)αnH(
x) + D,
where D:= C
1α. Define η:= H(
x) + D
εand consider the set Kε:= H1([0, η]). Since His
continuous, Kεis a closed (indeed, compact) set contained in int ((Sd)N)(by the first point at
[H]). Then, for every n, Markov inequality yields:
Pn1KC
ε(
x)ε.(17)
Consider a continuous function g: (Sd)N7→ [0,1] with the property g1in ((Sd)N),g0on
Kε. Since g1KC
ε, by (17) we obtain δ
xPngεfor every n. But νg > 2εwhich contradicts
the fact that νis a weak-limit point.
Remark 17. Consider for a while the dynamics given by the McKean-Vlasov replicator Ξ
discussed in the previous sections, and for every t1consider the operator ˜
Ptacting on
bounded functions as ˜
Ptf(x) := Ex(ft)). As remarked above, in general, we do not have
˜
Pnt =˜
Pn
t. So, the previous argument breaks down from the very start in the case of McKean-
Vlasov replicators, and thus new tools are needed to face the problem of persistence in this
setup.
Thus, [H] guarantees that the limit points of the sequence (Law(
X(N)
tn ) : n0) support the
whole ensemble of replicators. Also, observe that if νis an invariant probability measure for
the process
X(N)that supports the whole ensemble of replicators, then by definition νP =ν.
Let Aint ((Sd)N)be a measurable set with Leb(A)=0. Then:
ν(A) = νP (A) = Z(Sd)N
ν(d
x)δ
xP1A=Zint((Sd)N)
ν(d
x)δ
xP1A,
since νdoes not charge the boundary of (Sd)N. By Proposition 12, for every
xthere exists
a density, say p(
x,·), of the probability measure δ
xPwith respect to the Lebesgue measure.
Consequently:
ν(A) = Zint((Sd)N)
ν(d
x)δ
xP1A
=Zint(Sd)
ν(d
x)ZA
p(
x,
y)d
y=ZA
d
y Zint((Sd))N
ν(d
x)p(
x,
y)!,
INTERACTING REPLICATOR DYNAMICS 19
by Fubini’s theorem. Since Leb(A) = 0, the last integral vanishes, and we deduce ν(A) = 0.
In other words, if νis an invariant measure for the process
X(N)under which the full system
persists, then it is continuous with respect to the Lebesgue measure.
Finally, assume that we are able to prove that there exists a unique invariant measure, say
µ, for the semigroup Pwhich is supported at the interior of (Sd)N. We claim that H, as in
[H], is µ-integrable. Indeed, for n0, define Hn:= H n. Then, for every n1:
µHn= lim
k→∞ µP kHnµ(lim sup
k→∞
PkHn)D,
where the equality follows since µis invariant for P, and the first inequality is (reverse) Fatou’s
Lemma. Apply (direct) Fatou’s Lemma to obtain:
µH=Zint((Sd)N)
lim
n→∞ Hn(
x)µ(dx)lim inf
n→∞ Zint((Sd)N)
Hn(
x)µ(d
x)D,
and we deduce that His µ-integrable.
Actually, the existence of a function Has in [H] implies stronger results (see [3], page 186,
and the references therein, for the following lemma).
Lemma 18. Assume that Hexists as in [H]. Assume also that Ppossesses a density with
respect to the Lebesgue measure. Then, there exists a unique invariant probability measure π
for the kernel P, and moreover:
(1) πLeb.
(2) His π-integrable.
(3) There exists ρ(0,1) and C>0such that:
dT V (δ
xPn, π)C(1 + H(
x))ρn,
xint ((Sd)N).(18)
Remark 19. Observe that (18) implies that πattracts every initial measure νwith ν((Sd)N) =
0. Indeed, for any compact set Kint (Sd)N, by the definition of total variation norm and
Fubini’s theorem:
νP nπT V = sup
fBorel ,|f|≤1 Z(Sd)N
ν(d
x)Pnf(
x)Z(Sd)N
π(d
y)f(
y)!
= sup
fBorel ,|f|≤1Z(Sd)N
ν(d
x)Z(Sd)N
(Pn(
x, d
y)f(
y)π(d
y)f(
y))
Z(Sd)N
ν(d
x) sup
fBorel ,|f|≤1Z(Sd)N
(Pn(
x, d
y)f(
y)π(d
y)f(
y))
ZK
ν(d
x)δ
xPnπT V + 2ν(int ((Sd)N)\K),
since the total variation norm is bounded by 2and νdoes not charge the boundary. Since (Sd)N
is Polish, νis regular, and then given ε > 0we can choose Kεsuch that the last term is smaller
than ε. The result follows since the term inside the integral converges uniformly to 0on Kεas
n .
Theorem 20. Assume that Φis C2(Sd), and Assumptions 1-2-3 hold. Then for every r > 0
there exists δ > 0such that for any Υ,Ψof class C2such that ΠTΥ+|||AΨ||| < δ, the
following hold:
20 VIDELA, TEJO, QUIÑINAO, MARQUET, AND REBOLLEDO
(1) P(N)
inv (int ((Sd)N)) = {ν}; i.e., there exists a unique invariant probability measure sup-
ported on the interior of the N-fold topological sum of Sdfor the system (13). Moreover
νis absolutely continuous w.r.t. Lebesgue measure, and the function int ((Sd)N)
x7→
dist N(
x, ((Sd)N))r/N is ν-integrable.
(2) There exist positive constants C, αsuch that:
δxP(N)
tνT V Ceαt
dist N(x, ((Sd)N))r/N ,xint((Sd)N).
Proof. (of Theorem 20). We follow closely the steps in the proof Theorem 3.1 of [3]. Recall
the definition of the invasion rates (12). There exists α > 0, a neighbourhood Uof Sdand a
function W:Sd7→ Rof class C2(Sd)such that:
p,h(x)+⟨∇W(x),ΠTΦ(x)> α,(19)
for every x U (see [21], Theorems 3.4 and 4,4, and Remark 3.5), and where pis taken as in
Assumption 2. Define the map VN: (U \ Sd))N7→ Rgiven by
VN(
x) := 1
N
N
X
i=1
V(x(i)),
where V(x) := Pd
j=1 pjln(xj) + W(x). Since h(x) = diag(x)1ΠTΦ(x), we have:
N
X
i=1 D(VN(
x))(i),ΠTΦ(x(i))E=1
N
N
X
i=1 DV(x(i)),ΠTΦ(x(i))E> α,
on (U \ Sd)N. Apply the generator L(N)to VNto obtain:
L(N)VN(
x) =
N
X
i=1 DV(i)
N,
x(x(i)),ΠTΦ(x(i))E
+
N
X
i=1 DV(i)
N,
x(x(i)),ΠTΥ(x(i), µ
x)E
+1
2
N
X
i=1
Tr ΠTΨ(x(i))∇∇V(i)
N,
x(x(i)TΨ(x(i)),
where the map V(i)
N,
xis given by Sdy7→ VN(x(1),x(2),...,y,...,x(N)), where yis in the
place of the i-th block of order d.
Since Wis of class C2(Sd), the above expression can be made greater than α/2, say, for
x(U \ Sd))Nby choosing ˜ε1>0small enough and imposing ΠTΥ+|||AΨ||| <˜ε1.
Define λ:= r
inf pi
, and consider the function H: (int (Sd))N7→ R++ given by H:=
exp(λVN), which is of class C2((int (Sd))N). Furthermore
lim
η0inf
{
z:z(i)Ext(η)\Sd,i=1,...,N}H(
z) = .
INTERACTING REPLICATOR DYNAMICS 21
Since Wis bounded on Sd, with K:= exp(λinfzSdW(z)), we have:
H(
x)Kexp
λ(1/N)
N
X
i=1
d
X
j=1
pjlnx(i)
j
K
N
Y
i=1
exp
(r/N)
d
X
j=1
lnx(i)
j
=K
N
Y
i=1
1
(x(i)
1x(i)
2. . . x(i)
d)r/N
K
N
Y
i=1
1
dist (x(i), (Sd))r/N
K
dist N(
x, (Sd)N)r/N .(20)
Let Γ(N)be the carré du champ operator associated to the operator L(N), whose action on
smooth functions φ: (Sd)N7→ Ris given by:
Γ(N)φ(
x) = 1
2
N
X
i=1 X
j,k
TΨ(x(i)TΨ(x(i))]j,kjφ
x(x(i))kφ
x(x(i)).
For smooth functions f:R7→ R, φ : (Sd)N7→ R, we have:
L(N)f(φ(
x)) = f(φ(
x))L(N)φ(
x) + f′′(φ(
x))Γ(N)φ(
x).
Hence:
L(N)H(
x) = λH(
x)L(N)VN(
x)λΓ(N)VN(
x),
and then we can choose ˜ε < ˜ε1and impose ΠTΥ+|||AΨ||| <˜εsuch that the term inside the
parentheses is not smaller than α/3, say, on (U \ Sd)N. And then, for this choice:
L(N)H(
x) λα
3H(
x),
on (U \ Sd)N. Observe that L(N)His bounded on (int(Sd)\ U )Nby some absolute constant
C. Thus, with β:= λα
3, we obtain:
L(N)H(
x) βH(
x) + C,(21)
for all
x(int (Sd))N.
Let ηn:= inf{t0 : H(
X(N)
t)> n}, and observe that since His bounded on compact sets
of (int (Sd))N, by Lemma 11, we have ηn a.s. For given t0, set tn:= tηn. For every
n1, the process:
Mn
t=eβtnH(X(N)
tn) H(
X(N)
t0)Ztn
0
eβs(βH(
X(N)
s) + L(N)H(
X(N)
s))ds, t 0,
is a martingale, and thus for every n1, by (21):
E
x(eβtnH(
X(N)
tn)) = H(
x) + EZtn
0
Ceβ s H(
x) + C
β(1 eβt).
22 VIDELA, TEJO, QUIÑINAO, MARQUET, AND REBOLLEDO
Hence:
E
x(eβtH(
X(N)
t)1ttn) H(
x) + C
β,
and then,
E
x(H(
X(N)
t)1ttn)eβtH(
x) + C
β.
Since tnta.s., by Monotone Convergence, we deduce that for every t0and
x(int (Sd))N:
E
x(H(
X(N)
t)) eβtH(
x) + D,
where Dis a constant. In particular, there exists γ(0,1) such that, at time 1:
P(N)
1H γH+D,
on (int (Sd))N. By Proposition 12, P(N)
1possesses a density with respect to the Lebesgue
measure. Hence, Lemma 18 applies to P(N)
1and we deduce the existence of a unique invariant
probability measure µfor P(N)
1such that µ((Sd)N)=0, and moreover µattracts every initial
condition on (int (Sd))Nat geometric rate. Of course, the same line of reasoning applies to
P(N)
tfor every t(0,1], and we deduce that for every t(0,1] there exists a unique invariant
measure, say πt, for the chain (
X(N)
nt :n0). The uniqueness follows from Lemma 16.
The remaining parts of the result are exactly those in the reference [3].
5. Invariant distributions for the McKean-Vlasov stochastic replicator
We go back to the initial model of Fudenberg and Harris, as treated in [22], and consider
a non-linear, McKean-Vlasov-type extension of the model. Assume first that d= 2, that Ais
given, and that Σ = σI. Recall that in this case, each replicator solves (3). From Theorem 3.6
of [22] we deduce the following result.
Proposition 21. Assume that a12 a22 >0and a21 a11 >0, and furthermore:
a12 +a21 a11 a22 =σ2.
Then, the system (3) has a unique invariant distribution that does not charge the boundary,
namely the absolutely continuous measure with density Beta with parameters (a12 a22)σ2
and (a21 a11)σ2.
Observe that for the deterministic system associated to (3) the unique ergodic probability
measures that support strict subsets of the community are delta distribution, namely: δ1with
all the mass at the first population, and δ2with all the mass at the second population. It is
direct to show that the average positivity of the invasion rates (Assumption 2) holds in this
case.
Assume that we choose 0< δ < min{σ2, a12 a22, a21 a11 }. For an arbitrary probability
measure µon the simplex, define Φ(x) = ˜
Ax,Ψ(x)=Σand
Υ(x, µ) = δZy1µ(dy1, dy2)0 1
1 0!x,(22)
and the associated non-linear replicator:
dΞt= ΠTΦ(Ξt)dt + ΠTΥ(Ξt, µt)dt + ΠTΨ(Ξt)dWt.(23)
INTERACTING REPLICATOR DYNAMICS 23
Theorem 22. Equation (23) admits a unique invariant distribution, namely the Beta(s, 1s)
distribution, where s=a12 a22
σ2δ.
The proof of the above theorem will be transparent as soon as we prove its extension to
higher dimensions (and the computation of the invariant measure will follow after some simple
algebraic manipulation).
Before starting the next Lemma, observe that if a payoff matrix Asatisfies [C1], for every
x= (x1, . . . , xd) T we have:
xAx=1
2X
i,j
(aij +aji )xixj=1
2X
i,j
(aij +aji aii aj j )xixj
=1
2X
i=j
(aij +aji aii aj j )xixj
=1
2σ2
d
X
i=1
xi(x1+x2+. . . +xi1+xi+1 +. . . +xd) = 1
2σ2
d
X
i=1
x2
i.
In particular, if x T is not the null vector, Axcannot be a constant vector. We deduce that
the equation Aα=constant vector has at most one solution summing up to 1.
Lemma 23. Assume that Asatisfies [C1] and [C2]. Then there exists δ > 0such that for
any skew-symmetric matrix Ewith E< δ the matrix ˆ
A:= A+Esatisfies [C2].
Proof. First, observe that for Aand Eas in the claim, the matrix A+Esatisfies [C1] as
well. Next, observe that we can assume that Ehas the block-form:
E=δE202×(d2)
0(d2)×20,with E2=0 1
1 0,
since, if the result is true for this kind of perturbation, applying repeatedly the lemma we arrive
(via permutations of rows and columns) at the desired result.
We aim to find a strictly positive vector ˆα= (ˆα1,..., ˆαd)such that ˆ
Aˆαis a constant vector.
Let α:= (α1, . . . , αd)Rdand cRbe the vector and the real number ensured by [C2]
associated to the matrix A. For i= 1,2, . . . , d 1, write ˆαi:= αi+εi, and ˆαd=αdPd1
j=1 εj.
Then for a certain constant ˆc, for i= 1, . . . , d:
ˆc=
d
X
j=1
aij σ2/2) ˆαj
=
d1
X
j=1
(aij +δeij σ2/2)(αj+εj)+(aid +δeid σ2/2)(αd
d1
X
k=1
εk),
24 VIDELA, TEJO, QUIÑINAO, MARQUET, AND REBOLLEDO
must hold. Then:
ˆc=c+δ(α2+ε2) +
d1
X
j=1
(a1ja1d)εj
=cδ(α1+ε1) +
d1
X
j=1
(a2ja2d)εj
=c+
d1
X
j=1
(aij aid)εj, i = 3,4, . . . , d,
and consequently:
δ(α2+ε2) +
d1
X
j=1
(a1ja1d)εj=δ(α1+ε1) +
d1
X
j=1
(a2ja2d)εj
=
d1
X
j=1
(aij aid)εj,
for i= 3,4, . . . , d. This reduces to a linear system of d1equations for the d1unknowns
ε1, . . . , εd1. Let H(δ)be the matrix representing the above system. Then ε= (ε1, . . . , εd1)
satisfies the system (Sδ):
H(δ)ε(δ)=δ(α1+α2)δα2δα2· · · δα2.(24)
Since the equation Aα=constant is uniquely solvable for αstrictly positive, summing up to
1and satisfying [C1], the equation H(0)ε= 0 has a unique solution, and hence H(0) is non
singular. On the other side, it is clear that (H(δ):δ[0,1]) is a Lipschitz continuous family
of matrices (with respect to a fixed reference norm in RdRd, say the L2subordinate matrix
norm), and thus there exists 1δ > 0such that for δ < δ there exists a unique solution of
(Sδ). Let (ε(δ):δ[0, δ ]) be the family of solutions, and let δbe the right hand side of (Sδ).
Then for δ1, δ2[0, δ], and some constant C > 0that can change from line to line:
ε(δ1)ε(δ2)∥≤∥(H(δ1))1(δ1δ2)+((H(δ1))1(H(δ2))1)δ2
C(H(δ1))1∥∥δ1δ2+(H(δ1))1(H(δ2))1∥∥δ2
C(H(δ1))1∥|δ1δ2|
+C(H(δ1))1∥∥(H(δ2))1∥∥H(δ1)H(δ2).
As mentioned earlier, (H(δ):δ[0, δ]) is a Lipschitz family; in particular, it has bounded norm
and, consequently, the family of inverses is bounded as well (this is so since the coefficients of
H(δ)are continuous in δ, and consequently they are bounded on [0, δ ]; the determinant across
the family, being the determinant a continuous functions of the coefficients, is a continuous
function of a parameter ranging over a compact interval, and thus is bounded; again, since [0, δ]
is compact, the determinants are bounded away from zero; Cramer’s rule gives the conclusion).
Again, by Lipschitz continuity, the second term is bounded by C|δ1δ2|. We deduce that
(ε(δ):δ[0, δ]) is a Lipschitz family. So, we can find ˆ
δ < δ such that for δ < ˆ
δthe vector ˆαis
still strictly positive. Since, by definition, Pd
j=1 ˆαi= 1, this proves our claim.
INTERACTING REPLICATOR DYNAMICS 25
As a last ingredient in the proof of the main Theorem of this section, we need an estimate of
Wasserstein’s distance between Dirichlet distributions in terms of the differences between their
parameters. Hereafter, for a positive vector a= (a1, . . . , ad)with Pd
i=1 ai= 1, the notation
Dastands for the absolutely continuous measure with Dirichlet(a)density.
Lemma 24. For Daand Dbas above:
Was1(Da, Db)
d1
X
i=1
2(di)|aibi|.
We could not find this result in the available literature, so we provide a full proof in the
Appendix.
Theorem 25. Let Abe a payoff matrix satisfying [C1] and [C2], and F:P(Sd)7→ RdRd
be a Was1-Lipschitz function with range included in the set of skew-symmetric matrices. Let
Φ(x) = ˜
Ax,Ψ(x) = Σ, and Υ(x, µ) = δF (µ)x. Then there exists δ0>0such that for every
δ < δ0the stochastic McKean-Vlasov replicator with the previous parameter admits a unique
invariant probability measure putting no mass on the boundary of the simplex.
Proof. For Aas above, let δbe the tolerance guaranteed by Lemma 23, and fix δ < δ. For
fixed µ P(int (Sd)), let (X(µ,δ)
t:t0) be the unique solution of the “frozen” replicator:
dX(µ,δ)
t= ΠTΦ(X(µ,δ)
t)dt + ΠTΥ(X(µ,δ)
t, µ)dt + ΠTΨ(X(µ,δ)
t)dWt.(25)
By Theorem 3.6 of [22], there exists a vector α=α(µ)of strictly positive entries summing
up to 1such that Dα(µ)is invariant for X(µ,δ). Moreover, by Corollary 3.8 of [22], this is the
unique invariant distribution that does not charge the boundary. Let Tµ=Dα(µ). Moreover,
since Fis Lipschitz, from the proof of Lemma 23, it transpires that there exists a constant
C > 0such that:
αµαν1δCWas1(µ, ν ),
for every µ, ν P(Sd). But then, by Lemma 24:
Was1(Tµ, Tν)2(d1)δC Was1(µ, ν).
We deduce that for δ < min{δ, (2(d1)C)1}:= δ0, the map µ7→ Tµis a contraction. By the
Contraction Mapping Theorem (see e.g. Theorem 5.7, page 138 of [4]), there exists a unique
µ P(Sd)such that µ=Tµ. Clearly, µis invariant for the McKean-Vlasov replicator. Since
the range of Tis a subset of the Dirichlet family, the last claim follows.
6. Simulations
In Section 5, for the case of the McKean-Vlasov replicator with linear interaction function
and (small) Lipschitz dependence on the instantaneous measure (see Theorem 25), we have
stated the existence of a unique invariant probability measure putting no mass on the boundary.
In this section, via numerical simulations, we investigate whether this invariant probability
attracts the solutions of the system at large times. To that aim, we use the well-known Euler-
Maruyama scheme to simulate the associated N-replicator system, with step size h= 0.01.
26 VIDELA, TEJO, QUIÑINAO, MARQUET, AND REBOLLEDO
Our objective is to study the behaviour of the solutions to (23) as t , i.e. the long term
solutions to
dX(N;i)
t= ΠTΦ(X(N;i)
t)dt +1
N
N
X
j=1
ΠTΥ(X(N;i)
t, X(N;j)
t)dt + ΠTΨ(X(N;i)
t)dWt,
where X(N;i)
t= (x(N;i)
1(t), x(N;i)
2(t)) for Nlarge. Here Φ,Ψand Υ, correspond to those defined
in (22). In particular, we simulated the following sets of parameters
(PS1) A=0.5 1
1 0.5, σ = 1, δ = 0.05, s = 0.5263,
and
(PS2) A=0.6 0.9
1 0.4, σ = 0.9487, δ = 0.04, s = 0.5814,
For each of the two sets of parameters, to test the effect of the initial distribution, we perform
the simulations for initial conditions uniformly scattered in the unit interval, and also for
initial conditions uniformly scattered in the subset [0.2,0.4]. We called those distributions
uniform and lump initial conditions (UIC and LIC for short). For each of the four experiments
described, we approximate the probability density function from a normalized histogram of
the first coordinate of each of the N-replicator {x(N;i)
1}N
i=1. This gives us some qualitative
understanding of the time convergence of the probability density.
The contour plot shown in Figure 1 (top panel) corresponds to the time evolution of such
histograms for PS1 and UIC. Results for LIC and PS2 can be found in the appendix Figure 3.
We notice that the histograms stabilize after approximately 3000-time steps (t= 30), with small
fluctuations due to finite-size effects. At the bottom panel of Figure 1 we see that the final
empirical stationary pattern is similar to the Beta distribution stated at that Theorem 22, and
it seems to not depend on the initial distributions UIC or LIC (see appendix Figure 3). If such
convergence holds true, then as the time tgoes by, the position of the first coordinate of each
of the Ninteracting replicators would constitute a sample from the unique Beta distribution.
In that case, the empirical mean 1
NPN
i=1 x(N;i)
1would be converging towards the mean of the
limit distribution which is, in this case, s/(s+ (1 s)) = s. To test this conjecture we calculate
- for each parameter set and initial distribution - at each time step the empirical mean and
calculate the relative percentage error with respect to s. For any of the numerical experiments
performed, the results showed a relatively fast convergence of the empirical mean towards s,
with small fluctuations around ssmaller than 2%.
To test that the convergence is not only restricted to the empirical mean but also to the
empirical distribution given by the simulated N-replicator system, we calculate the empirical
cumulative distribution function (ECDF) for each set of parameters and initial distributions.
In Figure 2 we plot the ECDF for different times and each numerical experiment. In all cases,
the shape of the empirical CDF and the theoretical CDF given by Theorem 22, are relatively
similar. To further study this hypothesis, we performed at each time step tnlarger than
t= 30 an Anderson-Darling test to see whether the sample vector x(N;i)
1(tn)is drawn from
the theoretical Beta(s,1-s) distribution. For a 1% of significance level, we see that there is no
statistical evidence to reject the null hypothesis for 64.91%, 72.47%, 76.32%, and 80.31% of
the times for the four PS1-UIC, PS1-LIC, PS2-UIC and PS2-LIC respectively.
As a consequence, it seems reasonable to conjecture that ergodicity must hold true. We
perform an additional statistical analysis, but now consider true independent trajectories of
INTERACTING REPLICATOR DYNAMICS 27
PS1-UIC
0 10 20 30 40 50 60 70 80 90 100
time
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1
relative abundance probability
0
0.05
0.1
0.15
0.2
0.25
0.3
Figure 1. Simulations for the (S2)10000 interacting particle system. For the
N-replicator system, at each time step, we calculate the empirical histogram
from the first coordinate of each replicator under probability normalization.
The top panel shows the time evolution of those empirical relative abundance
probabilities for a set of parameters and initial conditions PS1-UIC. The bot-
tom panel shows the histogram for N-replicator at t= 100 (bins) and for a
sample of 107random numbers drawn for the theoretical Beta(s,1-s) (red line).
N-replicators at a significantly lower Nthan the previously used ones. To that end, we now
simulate 103times the 500-replicator dynamics with UIC for both sets parameters, saving the
trajectory of only one randomly selected replicator each time. We report a goodness-of-fit test
regarding the theoretical stationary Beta distribution specified in Theorem 22 at times t= 40,
and t= 50. In all cases, we will see now that the laws of the trajectories are quite close to
the corresponding theoretical stationary Beta distribution (for more details see also appendix
Figure 4).
The implemented test is based on an L2-type test statistic proposed in [12, Theorem 1 -
Corollary 1]: based on a random sample Y1, ..., Yn, their test statistic is given by:
Tnan,ˆ
bn) = nZ1
0
1
n
n
X
j=1 an+ˆ
bn)Yjˆan1{Yjt}tˆan(1 t)ˆ
bn
Ban,ˆ
bn)
2
dt,
where B(·,·)is the Beta function, and ˆanand ˆ
bnare consistent estimators of the Beta pa-
rameters, such as moment estimators or maximum likelihood estimators (MLEs). However,
28 VIDELA, TEJO, QUIÑINAO, MARQUET, AND REBOLLEDO
0 0.2 0.4 0.6 0.8 1
Relative abundance
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1
Empirical CDF
PS1-UIC
Empirical samples (t=10)
Empirical samples (t=25)
Empirical samples (t=50)
Empirical samples (t=75)
Empirical samples (t=100)
Theoretical CDF
0 0.2 0.4 0.6 0.8 1
Relative abundance
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1
Empirical CDF
PS1-LIC
Empirical samples (t=10)
Empirical samples (t=25)
Empirical samples (t=50)
Empirical samples (t=75)
Empirical samples (t=100)
Theoretical CDF
0 0.2 0.4 0.6 0.8 1
Relative abundance
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1
Empirical CDF
PS2-UIC
Empirical samples (t=10)
Empirical samples (t=25)
Empirical samples (t=50)
Empirical samples (t=75)
Empirical samples (t=100)
Theoretical CDF
0 0.2 0.4 0.6 0.8 1
Relative abundance
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1
Empirical CDF
PS2-LIC
Empirical samples (t=10)
Empirical samples (t=25)
Empirical samples (t=50)
Empirical samples (t=75)
Empirical samples (t=100)
Theoretical CDF
Figure 2. Empirical cumulative distribution function of the first coordinate of
the N-replicator system. At the top (left) panels, we show the first parameter
set (uniform initial distribution), and in the bottom (right) panels we show
the second parameter set (lump initial distribution). In all plots, we show the
ECDF for times t= 10,25,50,75 and t= 100, and the dashed black line is
the CDF associated with the theoretical Beta(s,1-s) distribution. Again, for
times larger than 25, the shape of the empirical and theoretical distribution
are qualitatively similar supporting the conjecture of ergodicity.
since in our simulations, the stationary distribution is completely specified by the resulting
parameterization given in Theorem 22, we can introduce such values, aand b, instead of its
estimates. Thus, for our null hypothesis, H0: “the law of the first coordinate trajectories of
the one-dimensional McKean-Vlasov replicator (23) follows a Beta(a, b)for time t”, our test
statistic will become:
(26) Tn=Tn(a, b) = nZ1
0
1
n
n
X
j=1
((a+b)Yja)1{Yjt}ta(1 t)b
B(a, b)
2
dt,
where in our case a=sand b= 1 a. The critical region to reject the H0is delimited by
1αquantiles of the distribution of the above test statistic under H0. Since Tnis a distance,
the larger the observed values of Tnare, the further we will be from the null hypothesis. To
obtain an approximation of the distribution of Tnunder H0, we construct the density of B
values of Tn, each one obtained by an independent simulated random sample from a Beta(a, b).
In our case, we used B= 5000. We found that the 90% quantiles are 0.1056 (PS1) and 0.1119
INTERACTING REPLICATOR DYNAMICS 29
(PS2). For t= 40 the computed tests statistics are Tn= 0.0175 (PS1) and Tn= 0.0046
(PS2). For t= 50 the computed tests statistics are Tn= 0.0817 (PS1) and Tn= 0.0492 (PS2).
In all cases, the null hypothesis is not rejected. The corresponding R-code can be found at
https://github.com/leonardo-videla/beta_test.git.
Given these results, it seems reasonable to conjecture that ergodicity must hold, at least in
this case. We have no proof yet, and thus this remains an open question.
7. Discussion and outlook
This paper advances the modeling of biological phenomena, which we have proposed in
previous papers using the framework of Open System Dynamics ([33], [45], [15], [34], [56]). We
regard our contribution as a step forward in understanding how the living develops in continuous
interaction with the environment, including this time the interaction within communities of
living entities in what we envisioned as an ecology of primordial or early life represented by
a community of replicators. In particular, we were interested in assessing if neutrality and
associated fitness equivalence could emerge in this system and under which conditions. To this
end, we generalized the stochastic replicator equation by adding a new component that affects
the instantaneous fitness, the idea being that this component accounts for the interaction of
one particular replicator population with others, alike ones, within a community in a mean-field
regime.
We show that under suitable hypotheses, the asymptotic behavior of interacting replicator
dynamics may be approached by independent non-markovian, McKean-Vlasov particles (Sec-
tion 3). For the mean-field replicator aggregates, we have shown that the conditions used in
[3], with some simple modifications, still apply to guarantee the coexistence (in the sense of
persistence) of large (but finite) systems of interacting communities (Section 4). In simple
cases (Section 5), we show the existence of a unique non-trivial invariant probability measure
for the McKean-Vlasov replicator. Indeed, in simple setups, we have been able to prove that
the Dirichlet family is stable (in the obvious sense that a system with Dirichlet invariant still
has a Dirichlet invariant when we introduce the dependence on the measure). This result,
which partially extends the work [22] and [29], admits further extensions, and it is a part of our
current research efforts. Likewise, regarding the ergodicity of the McKean-Vlasov replicator,
we still have no concluding results (although we have shown in Section 6 that numerical exper-
iments give strong empirical support in this direction). This is an open question we continue
investigating.
We point out some salient features of this article. The first one refers to the fact that, to the
best of our knowledge, this is the first attempt to import the McKean-Vlasov machinery into
the evolutionary-game-theoretic realm. Also, the study of persistence for interacting, finite
aggregates of complex communities seems to be a novel feature. The use of the fixed-point
argument to prove existence/uniqueness of invariant measures for McKean-Vlasov systems (see
Theorem 7) also seems to be a technique that has not been previously informed in this literature.
From the point of view of theoretical ecology, we note that the condition [C1] can be regarded
as the expression of neutrality and fitness equivalence among types, as assumed in the neutral
ecological theory ([26]); in turn, in previous work ([33]) we have shown that this condition is
associated with Beta invariant probability measure. In light of the results of Section 5, this
observation provides theoretical support for the idea that neutrality, or fitness equivalence,
is associated with the emergence of ecologies from simple, replicating molecules to complex
ecosystems, and is an emergent condition that is related to the persistence of the ecological
30 VIDELA, TEJO, QUIÑINAO, MARQUET, AND REBOLLEDO
system. Thus, unlike traditional models in ecology, we did not assume neutrality but neutrality
naturally emerged as a condition for persistence.
There are many open directions to continue this work. The first one was alluded to above,
namely: a proof for ergodicity of the McKean-Vlasov replicator is still missing. Secondly, we
were able to prove persistence for the finite-size particle system under small mean-field and
noise terms. The problem of proving the analogous property for the McKean-Vlasov replicator
was not faced in this article, but we notice in advance that the techniques used in Section 4 (a
modification of the well-known Foster-Lyapunov techniques as treated for example in the classic
articles of Meyn and Tweedie [38, 39, 40]) are unlikely to work in the non-linear case since they
are strongly grounded on the Markovian nature of the objects they are meant to be applied
to. Finally, another aspect that must be investigated is the extension of the model proposed
here to the setup of many types of interacting replicators. More specifically, we can think of a
myriad of entities of different types, say Ni1replicators of type ifor i= 1, . . . , M , where M
is the fixed number of types. A preliminary assessment of this extended setup shows that the
propagation-of-chaos (under suitable conditions) must still hold. The problem of persistence in
this multi-type interacting replicator seems to be a challenging issue, and we think it deserves
further investigation.
Acknowledgements This work was funded by project FONDECYT 1200925 The emergence of ecolo-
gies through metabolic cooperation and recursive organization, Centro de Modelamiento Matemático
(CMM), Grant FB210005, BASAL funds for centers of excellence from ANID-Chile, Exploration-ANID
13220168, Biological and quantum Open System Dynamics: evolution, innovation and mathematical
foundations, FONDECYT Iniciación project number 11240158-2024 Adaptive behavior in stochastic
population dynamics and non-linear Markov processes in ecoevolutionary modeling, and FONDECYT
Iniciación 11200436, Excitation and inhibition balance as a dynamical process, and “Programa de In-
serción Académica 2024 Vicerrectoría Académica y Prorrectoría de la Pontificia Universidad Católica
de Chile”. We thank Professor Rodrigo Plaza for his valuable help in the implementation of statistical
tests. We also thank the editor and an anonymous referee who provided us with valuable comments
that greatly improved our manuscript.
Declarations and conflict of interest
The authors declare that have no financial or personal relationship with other people or
organizations that could inappropriately influence or bias the content of this paper. Also, the
authors declare that they have no conflicts of interest.
References
[1] M. Amir and S. Berninghaus. Scale functions in equilibrium selection games. Journal of Evolutionary
Economics, 8:1–13, 1998.
[2] K. Avrachenkov and V. S. Borkar. Metastability in stochastic replicator dynamics. Dynamic Games and
Applications, 9:366–390, 2019.
[3] M. Benaïm, J. Hofbauer, and W. H. Sandholm. Robust permanence and impermanence for stochastic
replicator dynamics. Journal of Biological Dynamics, 2(2):180–195, 2008.
[4] H. Brezis. Functional Analysis, Sobolev Spaces and Partial Differential Equations. Springer New York,
NY, 2010.
[5] L.-P. Chaintron and A. Diez. Propagation of chaos: a review of models, methods and applications. i. models
and methods, 2022.
[6] F. Collet. Macroscopic limit of a bipartite curie–weiss model: a dynamical approach. Journal of Statistical
Physics, 157:1301–1319, 2014.
[7] P. Contucci, I. Gallo, and G. Menconi. Phase transitions in social sciences: two-population mean field
theory. International Journal of Modern Physics B, 22(14):2199–2212, 2008.
INTERACTING REPLICATOR DYNAMICS 31
[8] D. B. Cooney. The replicator dynamics for multilevel selection in evolutionary games. Journal of mathe-
matical biology, 79(1):101–154, 2019.
[9] R. Cressman. Evolutionary dynamics and extensive form games, volume 5. MIT Press, 2003.
[10] R. Cressman and Y. Tao. The replicator equation and other game dynamics. Proceedings of the National
Academy of Sciences, 111(supplement_3):10810–10817, 2014.
[11] G. Da Prato. Introduction to stochastic analysis and Malliavin calculus, volume 13. Springer, 2014.
[12] B. Ebner and S. C. Liebenberg. On a new test of fit to the beta distribution. Stat, 10(1):e341, 2021.
[13] O. D. Faugeras, J. D. Touboul, and B. Cessac. A constructive mean-field analysis of multi population neural
networks with random synaptic weights and stochastic inputs. Frontiers in computational neuroscience,
3:323, 2009.
[14] D. Foster and P. Young. Stochastic evolutionary game dyanamics. Theoret. Population Biol., pages 219–
232, 1990.
[15] M. A. Freilich, R. Rebolledo, D. Corcoran, and P. A. Marquet. Reconstructing ecological networks with
noisy dynamics. Proceedings of the Royal Society A, 476(2237):20190739, 2020.
[16] A. Friedman. Partial Differential Equations of Parabolic Type. Robert E. Krieger Publishing Company,
1983.
[17] D. Fudenberg and C. Harris. Evolutionary dynamics with aggregate shocks. J. Econon. Theory, 57:420–441,
1992.
[18] C. Hauert, S. De Monte, J. Hofbauer, and K. Sigmund. Volunteering as red queen mechanism for cooper-
ation in public goods games. Science, 296(5570):1129–1132, 2002.
[19] A. Hening and D. Nguyen. Coexistence and extinction for stochastic kolmogorov systems. Ann.Appl.Prob.,
28(3):1893–1942, 2018.
[20] J. Hofbauer. A general cooperation theorem for hypercycles. Monatshefte für Mathematik, 91(3):233–240,
1981.
[21] J. Hofbauer and B. Garay. Robust permanence for ecological differential equations, minimax, and dis-
cretizations. SIAM Journal of Mathematical Analysis, 34:1007–1039, 2003.
[22] J. Hofbauer and L. Imhof. Time averages, recurrence and transience in the stochastic replicator dynamics.
The Annals of Applied Probability, 19(4):1347–1368, 2009.
[23] J. Hofbauer, P. Schuster, and K. Sigmund. Competition and cooperation in catalytic selfreplication. Journal
of Mathematical Biology, 11(2):155–168, 1981.
[24] J. Hofbauer and K. Sigmund. The Theory of Evolution and Dynamical Systems. Mathematical Aspects of
Selection. Cambridge University Press, 1988.
[25] J. Hofbauer and K. Sigmund. Evolutionary games and population dynamics. Cambridge University Press,
1998.
[26] S. Hubbell. The Unified Neutral Theory of Biodiversity and Biogeography. Princeton University Press,
2001.
[27] S. P. Hubbell. Neutral theory in community ecology and the hypothesis of functional equivalence. Functional
ecology, 19(1):166–172, 2005.
[28] S. P. Hubbell. Neutral theory and the evolution of ecological equivalence. Ecology, 87(6):1387–1398, 2006.
[29] L. Imhof. The long-run behavior of the stochastic replicator dynamics. The Annals of Applied Probability,
15(1B):1019–1045, 2005.
[30] R. Khasminskii and N. Potsepun. On the replicator dynamics behavior under stratonovich type random
perturbations. Stochastics and Dynamics, 6(02):197–211, 2006.
[31] A. N. Kolmogorov. Deviations from hardy’s formulas under partial isolation. Dokl. Akad. Nauk. SSSR,
3:129–132, 1935.
[32] E. V. Koonin and A. S. Novozhilov. Origin and evolution of the genetic code: the universal enigma. IUBMB
life, 61(2):99–111, 2009.
[33] P. A. Marquet, G. Espinoza, S. R. Abades, A. Ganz, and R. Rebolledo. On the proportional abundance of
species: Integrating population genetics and community ecology. Scientific reports, 7(1):16815, 2017.
[34] P. A. Marquet, M. Tejo, and R. Rebolledo. What is the species richness distribution? Unsolved Problems
in Ecology, page 177, 2020.
[35] H. McKean. A class of markov processes associated with non-linear parabolic equations. Proceedings of the
National Academy of Sciences, 56(6):1907–1911, 1966.
[36] S. Méléard. Asymptotic behaviour of some interacting particle systems; mckean-vlasov and boltzmann
models. In D. Talay and L. Tubaro, editors, Probabilistic Models for Nonlinear Partial Differential Equa-
tions: Lectures given at the 1st Session of the Centro Internazionale Matematico Estivo (C.I.M.E.) held
32 VIDELA, TEJO, QUIÑINAO, MARQUET, AND REBOLLEDO
in Montecatini Terme, Italy, May 22–30, 1995, pages 42–95, Berlin, Heidelberg, 1996. Springer Berlin
Heidelberg.
[37] P. Mertikopoulos and Y. Viossat. Imitation dynamics with payoff shocks. International Journal of Game
Theory, 45:291–320, 2016.
[38] S. Meyn and R. Tweedie. Stability of markovian processes i: Criteria for discrete-time chains. Adv. Appl.
Prob-, 24(3):542–574, 1992.
[39] S. Meyn and R. Tweedie. Stability of markovian processes ii: Continuous time processes and sampled
chains. Adv. Appl. Prob., 25(3):487–517, 1992.
[40] S. Meyn and R. Tweedie. Stability of markovian processes iii: Foster-lyapunov criteria for continuous-time
processes. Adv. Appl. Prob., 25(3):518–548, 1993.
[41] M. A. Nowak and K. Sigmund. Evolutionary dynamics of biological games. Science, 303(5659):793–799,
2004.
[42] H. Ohtsuki and M. A. Nowak. The replicator equation on graphs. Journal of theoretical biology, 243(1):86–
97, 2006.
[43] K. M. Page and M. A. Nowak. Unifying Evolutionary Dynamics. Journal of Theoretical Biology, 219(1):93–
98, 2002.
[44] E. Pardoux and A. Răşcanu. Stochastic Differential Equations, Backward SDEs, Partial Differential Equa-
tions, volume 69 of Stochastic Modelling and Applied Probability. Springer, July 2014.
[45] R. Rebolledo, S. A. Navarrete, S. Kéfi, S. Rojas, and P. A. Marquet. An Open-System Approach to Complex
Biological Networks. SIAM Journal on Applied Mathematics, 79(2):619–640, 2019.
[46] M. Saito. A note on ergodic distributions in two-agent economies. Journal of Mathematical Economics,
27(2):133–141, 1997.
[47] W. H. Sandholm. Population games and evolutionary dynamics. MIT Press, 2010.
[48] S. Schreiber, M. Benaïm, and K. Atchadé. Persistence in fluctuating environments. J. Math. Biol.,
62(5):655—-683, 2011.
[49] P. Schuster and K. Sigmund. Replicator dynamics. Journal of Theoretical Biology, 100(3):533–538, 1983.
[50] S. Shahshahani. A New Mathematical Framework for the Study of Linkage and Selection. AMS, memoirs
of the american mathematical society edition, 1979.
[51] E. Smith and H. J. Morowitz. The Origin and Nature of Life on Earth: the emergence of the fourth
geosphere. Cambridge University Press, 2016.
[52] J. M. Smith and G. R. Price. The logic of animal conflict. Nature, 246(5427):15–18, 1973.
[53] A.-S. Sznitman. A propagation of chaos result for burgers’ equation. Probability theory and related fields,
71:581–613, 1986.
[54] A.-S. Sznitman. Topics in propagation of chaos. In P.-L. Hennequin, editor, Ecole d’Eté de Probabilités de
Saint-Flour XIX 1989, pages 165–251, Berlin, Heidelberg, 1991. Springer Berlin Heidelberg.
[55] P. Taylor and L. Jonker. Evolutionarily stable strategies and game dynamics. Mathematical Biosciences,
40:145–156, 1978.
[56] M. Tejo, C. Quiñinao, R. Rebolledo, and P. A. Marquet. Coexistence, dispersal and spatial structure in
metacommunities: a stochastic model approach. Theoretical Ecology, pages 1–24, 2021.
[57] V. Van. A new evolutionary law. Evolutionary theory, 1:1, 1973.
[58] A. Y. Veretennikov. On pathwise uniqueness of solutions for multidimensional mckean–vlasov equation.
Theory of Probability & Its Applications, 66(3):469–473, 2021.
[59] L. Videla. Strong stochastic persistence of some lévy-driven lotka-volterra systems. Journal of Mathematical
Biology, 84(11), 2022.
[60] S. I. Walker, P. C. Davies, and G. F. Ellis. From matter to life: information and causality. Cambridge
University Press, 2017.
[61] J. W. Weibull. Evolutionary game theory. MIT Press, 1995.
[62] J. J. Wiens, D. D. Ackerly, A. P. Allen, B. L. Anacker, L. B. Buckley, H. V. Cornell, E. I. Damschen,
T. Jonathan Davies, J.-A. Grytnes, S. P. Harrison, et al. Niche conservatism as an emerging principle in
ecology and conservation biology. Ecology letters, 13(10):1310–1324, 2010.
[63] S. Wright. Evolution in mendelian populations. Genetics, 16(2):97, 1931.
INTERACTING REPLICATOR DYNAMICS 33
Appendix A. Appendix
A.1. Derivation of the replicator equation from a model of population dynamics.
Define G:Rd\ {0} 7→ Sdvia:
G(y) = y1
1y,
and consider the SDE:
dYt=Yt(˜
Φ(G(Yt))dt + Ψ(G(Yt))dWt).
Here, Yis to be understood as a Rd
+-valued process of abundances. The characteristic ˜
Φ
and Ψare suitable vector- and matrix-valued functions, respectively. Observe that in this
formulation, the instantaneous fitness rates of the populations depend on Ytonly through the
relative proportions of the populations. Consider the case where Ψis a diagonal matrix (the
general case is analogous). Write Xt= (X(i)
t)i=1,...,d =G(Yt)=(gi(Yt))i=1,...,d, and let ˜
ϕi, ψi
the components of ˜
Φand the diagonal elements of Ψ, respectively. We easily compute:
kgi(y) = δik
1
y1
yi
y2.
kkgi(y) = 2 δik
1
y2+yi
y3.
Thus, Itô’s formula gives:
dX(i)
t= X
k
Y(k)
t˜
ϕk(Xt) δik
YtY(i)
t
Yt2!
+X
k
(ψk(Xt)Y(k)
t)2 Y(i)
t
Yt3δik
Yt2!!dt
+X
k
Y(k)
t˜
ψk(Xt) δik
YtY(i)
t
Yt2!dW (i)
t
= X
k
X(k)
t˜
ϕk(Xt)δik X(i)
t
+X
k
(ψk(Xt)X(k)
t)2X(i)
tδik!dt
+X
k
X(k)
t˜
ψk(Xt)δik X(i)
tdW (i)
t.
Hence, we see that Xsatisfies the equation (4) with:
Φ(x) = ˜
Φ(x)ΨΨ(x)x,
A.2. Proof of Proposition 8. Proposition 2: Under Assumptions 1 and 3, equation (14)
has a unique pathwise (hence strong) solution.
Proof. Since b= ΠTΦ + ΠTΥis Lipschitz, in particular, it is Dini-continuous. On the
other hand, Sdis compact, and consequently the drift and diffusion coefficients are bounded.
34 VIDELA, TEJO, QUIÑINAO, MARQUET, AND REBOLLEDO
Finally, Assumption 4.2 implies that for every xSd, we have ΠTΨ(x)y = 0 whenever
y T ,y= 0. But then:
inf
xSd,y∈T ,y=1 yΠTΨ(xTΨ(x)y>0,
since the expression inside the infimum is a continuous function of x,yand it is taken on a
compact set. Theorem 1 of [58] yields the claim.
A.3. Proof of Lemma 11. Lemma 1: For every Ξ0int (Sd):
P(MV )
Ξ0min
i=1,...,d Ξ(i)
t>0for all t0= 1.
An analogous result holds for the N- particle system, namely: for every
x0int ((Sd)N),
P(N)
x0min
j=1,...,N X(N;j)
t>0for all t0= 1.
In words: with probability 1, in finite time there is no absorption on the boundary of Sd(for
the McKean-Vlasov replicator) nor on the boundary of (Sd)N(for the N-particle system).
Proof. We prove the result for the McKean-Vlasov system, since the proof for the N-replicator
is completely analogous. For any kN, set:
Vk:= {xSd: min
ixi< ek},
and consider the function Uk:Sd\Vk7→ R+given by:
Uk(x) := X
i
ln(xi).
Then Ukis uniformly continuous for each k, and indeed it is a Cfunction in its domain. For
µ P(Sd)consider the operator Lµacting on a smooth function φ C2(Sd)as:
Lµφ(x) = b(x, µ),φ(x)+1
2TraceTΨ(x)∇∇φ(xTΨ(x)),(27)
where µtsolves the non-linear Fokker-Planck equation:
d
dt µt, φ=µt, Lµtφ,(28)
for every φ C2(Sd)([36], [5], [54]).
So, for xSd\Vkand µ P(Sd), we can compute:
LµUk(x) = X
iΦi(x) x,Φ(x)X
iΥi(x, µ) x,Υ(x, µ)+1
2X
i
Tr(AΨA
Ψ)(x),
and we observe that LµUkis, for each k, a uniformly continuous function on Sd. Moreover,
there exists an absolute constant, say M, such that LµUk< M for every k. Let:
τk:= inf{t0:ΞtVk}.
INTERACTING REPLICATOR DYNAMICS 35
Now, fix Ξ0int (Sd). There exists k0such that Ξ0Sd\Vkfor every kk0. It is known that
for kk0, the process (Uktτk)Uk0)Rt
0LLawsτk)sτk)ds :t0) is a martingale.
Thus, for every NN:
EΞ0(UkτkN)) = Uk0) + EΞ0 ZτkN
0
LLawsτk)Uksτk)ds!
Uk0) + MN.(29)
Let:
N:= \
kk0
{ω : τk< N},
and define ϵN:= PΞ0(ΩN). We have:
EΞ0(UkτkN)) EΞ0(Ukτk)·1N)
N,
by path-continuity. Using (29), we deduce that for fixed Nand every kk0:
NUk0) + M N C+M N.
for some constant Cthat depends on Ξ0but not on k. Letting kgo to infinity, we obtain
ϵN= 0 for every N, and consequently:
PΞ0(lim
kτk<) = PΞ0 [
NN
N!= 0,
and this proves our claim.
A.4. Proof of Lemma 24. Lemma 5: Let Daand Dbbe the Dirichlet distributions with
paramters aand b, both summing up to 1. Then:
Was1(Da, Db)
d1
X
i=1
2(di)|aibi|.
Proof. By dominated convergence, it is direct that the map a7→ Dais weakly continuous, i.e.
continuous when we endow the set of probability measures on the simplex with the topology of
weak convergence of probability measures. Thus, it suffices to consider the case where a,bhave
rational entries. So, for a certain natural number Nthere exists natural numbers m1, . . . , md
and n1, . . . , ndsuch that for ai=mi/N , bi=ni/N. Put Mi=Pjimj, and analogously
Ni=Pjinj. Let (Gj:j= 1, . . . , N )be an i.i.d. family of Gamma(1/N, 1) random variables.
Put:
Vi:=
Mi
X
j=Mi1+1
Gj;Wi:=
Ni
X
j=Ni1+1
GjZ:=
n
X
j=1
Gj.
Then X= (Xi)d
i=1 := (Vi/Z)d
i=1 and Y= (Yi)d
i=1 := (Wi/Z)d
i=1 have law Daand Dbrespec-
tively. We say that the ivariates overlap if [Mi1+ 1, Mi][Ni1+ 1, Ni]=. Observe that
if the ivariates overlap, then |ViWi|is bounded by the sum of two independent Gamma-
distributed random variables of parameters (|Ni1Mi1|
N,1) and (|NiMi|
N,1). If the ivariates
36 VIDELA, TEJO, QUIÑINAO, MARQUET, AND REBOLLEDO
do not overlap, then |ViWi|is bounded by the sum of two (not independent) Gamma-
distributed random variables of parameters (|Ni1Mi1|
N,1) and (|NiMi|
N,1). Denote by Ri
and Sithese random variables. We have:
E(|XiYi|)E(Ri/Z) + E(Si/Z)
=|Ni1Mi1|
N+|NiMi|
N
i1
X
j=1
|ajbj|+
i
X
j=1
|ajbj|.
Finally, by considering the first and last variates separately, we obtain:
Was1(Da, Db)
d
X
i=1
E(|XiYi|)
2(d1)|a1b1|+ 2(d2)|a2b2|+. . .
+. . . + 2 ×2|ad2bd2|+ 2|ad1bd1|,
and this is the claim.
A.5. Supplementary figures. In this section, we present some supplementary results for the
model studied in Section 6. Figure 3, shows the time evolution of the N-replicator for as second
set of initial conditions (LIC) and also for the second set of parameters (PS2-UIC and PS2-
LIC) along with the comparative between the final histogram for the N-replicator dynamics
and the Beta theoretical distribution of Theorem 22. These numerical results support the claim
that the long-term convergence and the shape of the steady state do not depend on the initial
conditions chosen or the used parameter set.
The second supplementary figure shows the result for the goodness-of-fit for different param-
eter sets at times t= 40 and t= 50. This is a graphical representation of the results explained
in the main text, along with the 90%, 95%, and 99% quantiles. We can see that in none of the
cases we can reject the null hypothesis, supporting the conjecture that as the system evolves
in time, the solutions converge to the stationary distribution given by Theorem 22
1Departamento de Matemática y Ciencia de la Computación, Facultad de Ciencia, Universidad
de Santiago de Chile, Las Sophoras 173, Estación Central, Santiago, Chile
Email address:leonardo.videla@usach.cl
2Instituto de Estadística, Universidad de Valparaíso, Valparaíso, Chile
Email address:mauricio.tejo@uv.cl
3Facultad de Ciencias Biológicas, Pontificia Universidad Católica de Chile, Santiago, Chile
Email address:cquininao@uc.cl
Email address:pmarquet@puc.bio.cl
4Centro de Modelamiento Matemático (CMM), Universidad de Chile-IRL 2807 CNRS Beauchef
851, Santiago, Chile
4The Santa Fe Institute, 1399 Hyde Park Road, Santa Fe NM 87501, USA
4Instituto de Sistemas Complejos de Valparaíso, Subida Artillería 470, Valparaíso, Chile
5Instituto de Ingeniería Matemática, Universidad de Valparaíso
Email address:rolando.rebolledo@uv.cl
INTERACTING REPLICATOR DYNAMICS 37
PS1-LIC
0 10 20 30 40 50 60 70 80 90 100
time
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1
relative abundance probability
0
0.05
0.1
0.15
0.2
0.25
0.3
PS2-UIC
0 10 20 30 40 50 60 70 80 90 100
time
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1
relative abundance probability
0
0.05
0.1
0.15
0.2
0.25
0.3
PS2-LIC
0 10 20 30 40 50 60 70 80 90 100
time
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1
relative abundance probability
0
0.05
0.1
0.15
0.2
0.25
0.3
Figure 3. Simulations for the (S2)10000 interacting particle system. For the
N-replicator system, at each time step, we calculate the empirical histogram
from the first coordinate of each replicator under probability normalization.
From top to bottom, the plot shows the time evolution of the empirical relative
abundance probability for PS1-LIC, PS2-UIC, and PS2-LIC.
38 VIDELA, TEJO, QUIÑINAO, MARQUET, AND REBOLLEDO
Tn obs = 0.01756
quantile 99% = 0.24979
quantile 95% = 0.15579
quantile 90% = 0.10562
0
5
10
15
20
0.0 0.1 0.2 0.3
Tn
Probability density
Tn obs = 0.08174
quantile 99% = 0.24979
quantile 95% = 0.15579
quantile 90% = 0.10562
0
5
10
15
20
0.0 0.1 0.2 0.3
Tn
Probability density
Tn obs = 0.00465
quantile 99% = 0.23385
quantile 95% = 0.14588
quantile 90% = 0.11191
0
5
10
15
20
0.0 0.1 0.2 0.3
Tn
Probability density
Tn obs = 0.04922
quantile 99% = 0.23385
quantile 95% = 0.14588
quantile 90% = 0.11191
0
5
10
15
20
0.0 0.1 0.2 0.3
Tn
Probability density
Figure 4. Graphical hypothesis testing for contrasting H0: “the law of the
first coordinate trajectories of (23) follows a Beta(a, b)at time t”. The sta-
tistical test is based on the work of [12]. The upper panels show the results
for PS1 and the lower panels are the corresponding results for PS2, with their
corresponding theoretical sgiven in the main text. The left column corre-
sponds to time t= 40 and the right one to time t= 50. The test is based
on the statistic Tn, constructed according to a L2distance given in Equation
(26), whose observed values are shown by the red dashed lines. We can notice
that, in all the cases, they are much lower than the quantiles that delimit the
rejecting region at each indicated (1 α)×100 percentile, with α= 0.1(yel-
low dashed lines), α= 0.05 (purple dashed lines) and α= 0.01 (green dashed
lines), regarding the theoretical probability density of the statistic Tnunder
H0, approximated by simulations.
ResearchGate has not been able to resolve any citations for this publication.
Article
Full-text available
In many biological systems, natural selection acts simultaneously on multiple levels of organization. This scenario typically presents an evolutionary conflict between the incentive of individuals to cheat and the collective incentive to establish cooperation within a group. Generalizing previous work on multilevel selection in evolutionary game theory, we consider a hyperbolic PDE model of a group-structured population, in which members within a single group compete with each other for individual-level replication; while the group also competes against other groups for group-level replication. We derive a threshold level of the relative strength of between-group competition such that defectors take over the population below the threshold while cooperation persists in the long-time population above the threshold. Under stronger assumptions on the initial distribution of group compositions, we further prove that the population converges to a steady state density supporting cooperation for between-group selection strength above the threshold. We further establish long-time bounds on the time-average of the collective payoff of the population, showing that the long-run population cannot outperform the payoff of a full-cooperator group even in the limit of infinitely-strong between-group competition. When the group replication rate is maximized by an intermediate level of within-group cooperation, individual-level selection casts a long shadow on the dynamics of multilevel selection: no level of between-group competition can erase the effects of the individual incentive to defect. We further extend our model to study the case of multiple types of groups, showing how the games that groups play can coevolve with the level of cooperation.
Article
Full-text available
We study a class of Lotka–Volterra stochastic differential equations with continuous and pure-jump noise components, and derive conditions that guarantee the strong stochastic persistence (SSP) of the populations engaged in the ecological dynamics. More specifically, we prove that, under certain technical assumptions on the jumpsizes and rates, there is convergence of the laws of the stochastic process to a uniquestationary distribution supported far away from extinction. We show how the tech-niques and conditions used in proving SSP for general Kolmogorov systems driven solely by Brownian motion must be adapted and tailored in order to account for the jumps of the driving noise. We provide examples of applications to the case where the underlying food-web is: (a) a 1-predator, 2-prey food-web, and (b) a multi-layer food-chain. Share link provided by Springer: https://rdcu.be/cEThL
Article
Full-text available
We propose a stochastic model for interacting species in a metacommunity in order to study the factors affecting the intensity of the competition/colonization trade-off as a coexistence mechanism in metacommunities. We particularly focus on the role of the number of local communities and the number of refuges for the inferior competitor. The stochastic component is associated with the dispersal process and is represented by Poisson random measures. Thus, this stochastic model includes two dynamic scales: a continuous one, which refers to the interactions among species, and a low frequency one, referring to dispersal following a Poisson scheme. We show the well-posedness of the model and that it is possible to study its long-term behavior using Lyapunov exponents; the extinction of a species is associated with a negative slope in the time trajectory of the Lyapunov exponent, otherwise, it is equal to zero. We show that the competition/colonization trade-off is a function of the dispersal rate of the inferior competitor, and that it becomes less intense as the number of local communities increases, while the opposite is true with an increase in the number of refuges for the inferior competitor. We also show that under a priority effect type of scenario, dispersal can reverse priority effects and generate coexistence. Our results emphasize the importance of coexistence mechanisms related to the topology of the system of local communities, and its relationship with dispersal, in affecting the result of competition in local communities.
Article
Full-text available
We propose a new L²‐type goodness‐of‐fit test for the family of beta distributions based on a conditional moment characterisation. The asymptotic null distribution is identified, and since it depends on the underlying parameters, a parametric bootstrap procedure is proposed. Consistency against all alternatives that satisfy a convergence criterion is shown, and a Monte Carlo simulation study indicates that the new procedure outperforms most of the classical tests. Finally, the procedure is applied to a real data set related to air humidity.
Article
Full-text available
Ecosystems functioning is based on an intricate web of interactions among living entities. Most of these interactions are difficult to observe, especially when the diversity of interacting entities is large and they are of small size and abundance. To sidestep this limitation, it has become common to infer the network structure of ecosystems from time series of species abundance, but it is not clear how well can networks be reconstructed, especially in the presence of stochasticity that propagates through ecological networks. We evaluate the effects of intrinsic noise and network topology on the performance of different methods of inferring network structure from time-series data. Analysis of seven different four-species motifs using a stochastic model demonstrates that star-shaped motifs are differentially detected by these methods while rings are differentially constructed. The ability to reconstruct the network is unaffected by the magnitude of stochasticity in the population dynamics. Instead, interaction between the stochastic and deterministic parts of the system determines the path that the whole system takes to equilibrium and shapes the species covariance. We highlight the effects of long transients on the path to equilibrium and suggest a path forward for developing more ecologically sound statistical techniques.
Article
The notion of propagation of chaos for large systems of interacting particles originates in statistical physics and has recently become a central notion in many areas of applied mathematics. The present review describes old and new methods as well as several important results in the field. The models considered include the McKean-Vlasov diffusion, the mean-field jump models and the Boltzmann models. The first part of this review is an introduction to modelling aspects of stochastic particle systems and to the notion of propagation of chaos. The second part presents concrete applications and a more detailed study of some of the important models in the field.
Article
The notion of propagation of chaos for large systems of interacting particles originates in statistical physics and has recently become a central notion in many areas of applied mathematics. The present review describes old and new methods as well as several important results in the field. The models considered include the McKean-Vlasov diffusion, the mean-field jump models and the Boltzmann models. The first part of this review is an introduction to modelling aspects of stochastic particle systems and to the notion of propagation of chaos. The second part presents concrete applications and a more detailed study of some of the important models in the field.
Article
Since Darwin, individuals and more recently genes, have been the focus of evolutionary thinking. The idea that selection operates on nonreproducing, higher-level systems including ecosystems or societies, has met with scepticism. But research emphasising that natural selection can be based solely on differential persistence invites reconsideration of their evolution. Self-perpetuating feedback cycles involving biotic as well as abiotic components are critical to determining persistence. Evolution of autocatalytic networks of molecules is well studied, but the principles hold for any ‘self-perpetuating’ system. Ecosystem examples include coral reefs, rainforests, and savannahs. Societal examples include agricultural systems, dominant belief systems, and economies. Persistence-based selection of feedbacks can help us understand how ecological and societal systems survive or fail in a changing world.
Article
The process by which chemistry can give rise to biology remains one of the biggest mysteries in contemporary science. The de novo synthesis and origin of life both require the functional integration of three key characteristics — replication, metabolism and compartmentalization — into a system that is maintained out of equilibrium and is capable of open-ended Darwinian evolution. This Review takes systems of self-replicating molecules as starting points and describes the steps necessary to integrate additional characteristics of life. We analyse how far experimental self-replicators have come in terms of Darwinian evolution. We also cover models of replicator communities that attempt to solve Eigen’s paradox, whereby accurate replication needs complex machinery yet obtaining such complex self-replicators through evolution requires accurate replication. Successful models rely on a collective metabolism and a way of (transient) compartmentalization, suggesting that the invention and integration of these two characteristics is driven by evolution. Despite our growing knowledge, there remain numerous key challenges that may be addressed by a combined theoretical and experimental approach. Self-replicating systems play a central role in the emergence of life. This Review describes the features that self-replicating systems need to acquire to transition from chemistry to biology and surveys the progress made in theoretical and experimental approaches.