Asymptotic Dynamics of Hamiltonian Polymatrix Replicators
Abstract.
In a previous paper [ADP2020] we have studied flows defined on polytopes, presenting a new method to encapsulate its asymptotic dynamics along the edge-vertex heteroclinic network. These results apply to the class of polymatrix replicator systems, which contains several important models in Evolutionary Game Theory. Here we establish the Hamiltonian character of the asymptotic dynamics of Hamiltonian polymatrix replicators.
Key words and phrases:
Hamiltonian polymatrix replicator system, Poisson structure, Poincaré map, Asymptotic dynamics, Heteroclinic network.2010 Mathematics Subject Classification:
34D05, 37J06, 37J46, 53D17, 91A221. Introduction
A new method to study the asymptotic dynamics of flows defined on polytopes was presented in [ADP2020]. This method allows us to analyze the asymptotic dynamics of flows defined on polytopes along the edge-vertex heteroclinic network. Examples of such dynamical systems arise naturally in the context of Evolutionary Game Theory (EGT) developed by J. Maynard Smith and G. R. Price [smith1973logic].
One such example is the polymatrix replicator, introduced in [AD2015, ADP2015]. This is a system of ordinary differential equations that models the time evolution of behavioral strategies of individuals in a stratified population.
The polymatrix replicator induces a flow on a prism (simple polytope) given by a finite product of simplices. These systems extend the class of the replicator and the bimatrix replicator equations studied e.g. in [TJ1978] and [schuster1981coyness, schuster1981selfregulation], respectively.
In [AD2015] the authors have introduced the subclass of conservative polymatrix replicators (see Definition 4.4) which are Hamiltonian systems with respect to appropriate Poisson structures. In [ADP2015] these Hamiltonian polymatrix replicators are used to describe the asymptotic dynamics of the larger class of dissipative polymatrix replicators.
For Hamiltonian vector fields on symplectic manifolds it is well known that the Poincaré map preserves the induced symplectic structure on any transversal section. In this paper we extend this fact to Hamiltonian systems on Poisson manifolds, showing that any transversal section inherits a Poisson structure and the Poincaré map preserves it. Using this result we study the Hamiltonian character of the asymptotic dynamics of conservative polymatrix replicators along their edge-vertex heteroclinic network. Our main result states that for conservative polymatrix replicators the map describing the asymptotic dynamics is Hamiltonian with respect to an appropriate Poisson structure (Theorem 6.17).
The paper is organized as follows. In Section 2 we recall the method in [ADP2020], outlining the construction of the asymptotic dynamics for a large class of flows on polytopes that includes the polymatrix replicators. In Section 3 we define Poincaré maps for Hamiltonian systems on Poisson manifolds. In Section 4 we provide a short introduction to polymatrix replicators, following [AD2015]. Namely, we state the basic definitions and results for the class of conservative polymatrix replicators, that we also designate as Hamiltonian polymatrix replicators. In Section 5 we review the main definitions and results for the polymatrix replicator systems regarding the construction outlined in Section 2. In Section 6 we analyze the Hamiltonian character of Poincaré maps in the case of Hamiltonian polymatrix replicators. Finally, in Section 7 we present an example of a five-dimensional Hamiltonian polymatrix replicator to illustrate the main concepts and results of this paper. The graphics of this section were produced with Wolfram Mathematica and Geogebra software.
2. Outline of the construction
We now outline the construction of the asymptotic dynamics for a large class of flows on polytopes that includes the polymatrix replicators. A polytope is a compact convex set in some Euclidean space obtained as the intersection of finitely many half-spaces. A polytope is called simple if the number of edges (or facets) incident with each vertex equals the polytope’s dimension. The phase space of polymatrix replicators, that are prisms given by a finite product of simplices, are examples of simple polytopes. In [ADP2020] we consider analytic vector fields defined on simple polytopes which have the property of being tangent to every face of the polytope. Such vector fields induce complete flows on the polytope which leave all faces invariant. Vertices of the polytope are singularities of the vector field, while edges without singularities, called flowing edges, consist of single orbits flowing between two end-point vertices. The vertices and flowing edges form a heteroclinic network of the vector field. The purpose of this construction is to analyze the asymptotic dynamics of the vector field along this one-dimensional skeleton. Throughout the text we assume that every vector field is non-degenerate. This means that the transversal derivative of the vector field is never identically zero along any facet of the polytope.
The analysis of the vector field’s dynamics along its edge-vertex heteroclinic network makes use of Poincaré maps between cross sections tranversal to the flowing edges. Any Poincaré map along a heteroclinic or homoclinic orbit is a composition of two types of maps, global and local Poincaré maps. A global map, denoted by , is defined in a tubular neighborhood of any flowing-edge . It maps points between two cross sections and transversal to the flow along the edge . A local map, denoted by , is defined in a neighborhood of any vertex . For any pair of flowing-edges such that is both the ending point of and the starting point of , the local map takes points from to (see Figure 1).

Asymptotically, the nonlinear character of the global Poincaré maps fade away as we approach a heteroclinic orbit. This means that these non-linearities are irrelevant for the asymptotic analysis. For regular111The reader should bare in mind that the concept of regularity used here (Definition 5.6) is more restrictive then the one in [ADP2020, Definition ]. vector fields, the skeleton character at a vertex, defined as the set of eigenvalues of the tangent map along the edge eigen-directions, completely determines the asymptotic behavior of the local Poincaré map at that vertex.

To describe the limit dynamical behavior we introduce the dual cone of a polytope where the asymptotic piecewise linear dynamics unfolds. This space lies inside , where is the set of the polytope’s facets. The dual cone of a -dimensional simple polytope is the union
where for each vertex , is the -dimensional sector consisting of points with non-negative coordinates such that for every facet that does not contain . See Figure 2.
Given a vector field on a -dimensional polytope , we now describe a rescaling change of coordinates , depending on a blow up parameter . See Figure 3.

This change of coordinates maps tubular neighborhoods of edges and vertices to the dual cone . For instance, the tubular neighborhood of a vertex is defined as follows. Consider a system of affine coordinates around , which assigns coordinates to and such that the hyperplanes are precisely the facets of the polytope through . Then is defined by
The sets are called the outer facets of . The remaining facets of , defined by equations like , are called the inner facets of . The previous cross sections can be chosen to match these inner facets of the neighborhoods .
The rescaling change of coordinates maps to the sector . Enumerating so that the facets through are precisely , the map is defined on the neighborhood by
Similarly, given an edge , maps a tubular neighborhood of to the facet sector of where is the other end-point of . The map sends interior facets of and respectively to boundary facets of and while it maps outer facets of and to infinity. As the rescaling parameter tends to , the rescaled push-forward of the vector field converges to a constant vector field on each sector . This means that asymptotically, as , trajectories become lines in the coordinates . Given a flowing-edge between vertices and , the map over depends only on the coordinates transversal to . Moreover, as the global Poincaré map converges to the identity map in the coordinates . Hence the sector is naturally identified as the common facet between the sectors and . Hence the asymptotic dynamics along the edge-vertex heteroclinic network is completely determined by the vector field’s geometry at the vertex singularities and can be described by a piecewise constant vector field on the dual cone, whose components are precisely those of the skeleton character of . We refer to this piecewise constant vector field as the skeleton vector field of . This vector field induces a piecewise linear flow on the dual cone whose dynamics can be computationally explored.
We use Poincaré maps for a global analysis of the asymptotic dynamics of the flow of . We consider a subset of flowing-edges with the property that every heteroclinic cycle goes through at least one edge in . Such sets are called structural sets. The flow of induces a Poincaré map on the system of cross sections . Each branch of the Poincaré map is associated with a heteroclinic path starting with an edge in and ending at its first return to another edge in . These heteroclinic paths are the branches of . The flow of the skeleton vector field also induces a first return map on the system of cross sections . This map , called the skeleton flow map, is piecewise linear and its domain is a finite union of open convex cones. In some cases, see Proposition 5.18, the map becomes a closed dynamical system.
We can now recall the main result in [ADP2020], Theorem 5.20 below, which says that under the rescaling change of coordinates , the Poincaré map converges in the topology to the skeleton flow map , in the sense that the following limit holds
with uniform convergence of the map and its derivatives over any compact set contained in the domain .
Consider now, for each facet of the polytope, an affine function which vanishes on and is strictly positive on the rest of the polytope. With this family of affine functions we can present the polytope as . Any function function of the form
rescales to the following piecewise linear function on the dual cone
in the sense that . When all coefficients have the same sign then is a proper function on the dual cone and all levels of are compact sets. If the function is invariant under the flow of , i.e. , then the piecewise linear function is also invariant under the skeleton flow, i.e. . Thus integrals of motion (of vector fields on polytopes) of the previous form give rise to (asymptotic) piecewise linear integrals of motion for the skeleton flow.
3. Poisson Poincaré maps
In this section we will define Poincaré map for Hamiltonian systems on Poisson manifolds. For Hamiltonian vector fields on symplectic manifolds it is well known that the Poincaré map preserves the induced symplectic structure on any transversal section (see [MR2554208]*Theorem 1.8.). We extend this fact to Hamiltonian systems on Poisson manifolds, showing that any transversal section inherits a Poisson structure and the Poincaré map preserves this structure.
A Poisson manifold is a pair where is a smooth manifold without boundary and a Poisson structure on . Recall that a Poisson structure is a smooth bivector field with the property that , where is the Schouten bracket (cf. e.g. [MR2178041]). The bivector field defines a vector bundle map
(3.1) |
The image of this map is an integrable singular distribution which integrates to a symplectic foliation, i.e., a foliation whose leaves have a symplectic structure induced by the Poisson structure.
Notice that a Poisson structure can also be defined as a Lie bracket on satisfying the Leibniz rule
These two descriptions are related by . In a local coordinate chart , or equivalently when , a Poisson bracket takes the form
where is a skew symmetric matrix valued smooth function satisfying
Definition 3.1.
Let and be two Poisson manifolds. A smooth map will be called a Poisson map iff
(3.2) |
Using the map , defined in (3.1), this condition reads as
(3.3) |
where we use the notation to denote the adjoint operator of . Notice that, if is the Jacobian matrix of in local coordinates, then the matrix representative of the pullback will be .
Remark 3.2.
When is a diffeomorphism and only one of the manifolds or is Poisson manifold, Definition 3.1 can be used to push-forward or pullback the Poisson structure to the other manifold.
Definition 3.3.
Let be a Poisson manifold. The Hamiltonian vector field associated to a given function is defined by derivation for , or equivalently .
As in the symplectic case, to define the Poincaré map we will consider the traversal sections inside the level set of the Hamiltonian. We will show that such a transversal section is a cosymplectic submanifolds of the ambient Poisson manifold and naturally inherits a Poisson structure. For more details on cosymplectic submanifolds see [CALVO2010259]*Section .
Definition 3.4.
is a cosymplectic submanifold if it is the level set of second class constraints i.e., where are functions such that is an invertible matrix at all points .
Remark 3.5.
A constraint is called first class if it Poisson commutes with other constraints of the system. Sometimes, in the literature, a constraint that has non-zero Poisson bracket with at least one other constraint of the system is called a second class constraint. Definition 3.4 demands a stronger condition, but the cosymplectic submanifolds that we will use have codimension , where having non-zero Poisson bracket with the other constraint is the same as being an invertible matrix.
Every, cosympletic submanifold is naturally equipped with a Poisson bracket called Dirac bracket. Paul Dirac, [paul-dirac-book], developed this bracket to treat classical systems with second class constraints in Hamiltonian mechanics.
Definition 3.6.
For cosymplectic submanifold , let
be its second class constraints, where is a small enough neighborhood of in such that the matrix is invertible at all points . The Dirac bracket is defined on by
(3.4) |
where is the column matrix with components , .
Dirac bracket is actually a Poisson bracket on the open submanifold , see [CALVO2010259]. It takes an easy calculation to see that constraint functions are Casimirs of Dirac bracket. This fact allows the restriction of Dirac bracket to the cosymplectic submanifold . Note that, in general, restricting (pulling back) a Poisson structure to an arbitrary submanifold is not straightforward. Actually, the decomposition
(3.5) |
that holds for every point and a strait forward calculation yield the independence of the extension in the following definition. In Equation (3.5), the term is the annihilator of in . We will use this notation in the rest of the paper. Equation (3.5) can be used as definition of a cosymplectic submanifold, see [CALVO2010259], but for our propose it suits better to use the second class constraints to define cosymplectic submanifolds.
Definition 3.7.
The restricted Dirac bracket on cosymplectic submanifold , which will be also referred to as Dirac bracket, is simply defined by extending in any arbitrary way functions on to functions on , calculating their Dirac bracket on and restricting the result back to .
We consider a Hamiltonian on the -dimensional Poisson manifold and its associated Hamiltonian vector field defined by . For a given point let be a neighborhood around it such that , and be the energy surface passing through , i.e., the connected component of containing . We call level transversal section to at a regular point any -dimensional transversal section through .
The following lemma shows that is a cosymplectic submanifold.
Lemma 3.8.
Every level transversal section is a cosymplectic submanifold of .
Proof.
Since , there exist a function locally defined in (shrink if necessary) and linearly independent from such that
Then, we have
by transversality. This finishes the proof. ∎
Remark 3.9.
The second term in the right hand side of Equation (3.4) is
Then, using extensions and of such that at every point their differentials vanish on , yields
We will use this fact to simplify our proofs but arbitrary extensions are more suitable for calculating the Dirac structure.
We will use the same notation for arbitrary extension and reserve the notation for extension that their differentials vanishes on at every point . To avoid any possible confusion, we observe that in [MR2128714]*Section 8 and [MR2128714]*Section 8 the notation is used in a slightly different sense.
Remark 3.10.
Cosymplectic submanifolds are special examples of the so called Poisson-Dirac submanifolds, see [MR2128714]*Section 8. The induced Poisson structure on a Poisson-Dirac submanifold is defined by using extensions such that their differentials vanish on . In [MR2128714]*Section 8 and [CALVO2010259]*Lemma the notation is used for this type of extensions. For a cosymplectic submanifold given by second class constraints , we have
(3.6) |
and the Dirac bracket coincides with the bracket induced in this way, see [CALVO2010259]*Section . In our case, we only have two constraints and requiring the vanishing of the differential only on (or ) at every point is enough to obtain the same induced Poisson bracket.
For a fixed time , let , where is the flow of the Hamiltonian vector field , and be level transversal sections at and , respectively. As usual, a Poincaré map can be defined from an appropriate neighborhood of in to a neighborhood of in . The existence of the smooth function is guaranteed by the Implicit Function Theorem. We replace and by the domain and the image of the Poincaré map .
By Lemma 3.8 both , , are cosymplectic submanifolds equipped with Dirac brackets . We will show that the Poincaré map is a Poisson map (see Definition 3.1).
Proposition 3.11.
The Poincaré map
is a Poisson map.
Proof.
We define by
where is an extension of to a neighborhood of such that its differential, , vanishes on . Both neighborhood and can be shrunk, if necessary, in a way that both Dirac brackets around and are defined in and , respectively. A straightforward calculation shows that for every point in the domain of , we have
where . Furthermore, for every , we have
(3.7) |
Note that in Equation (3) is the derivative of time- flow of and the fixed time flow maps of are Poisson maps, i.e. it sends Hamiltonian vector fields to Hamiltonian vector field. Furthermore, the flow of preserves , this means that
As we set in Remark 3.9, let be an extension of a given such that
then for every ,
Now, for and we have
and consequently,
where we used Remark 3.9. This finishes the proof. ∎
4. Hamiltonian polymatrix replicators
In this section we provide a short introduction to polymatrix replicators, following [AD2015]. In particular we will focus on the class of conservative polymatrix replicators that we designate as Hamiltonian polymatrix replicators.
Consider a population divided in groups where each group is labeled by an integer , and the individuals of each group have exactly strategies to interact with other members of the population (including of the same group). In total we have strategies that we label by the integers , denoting by
the set (interval) of strategies of group .
Given , consider a real matrix, say , whose entries , with and , represent the payoff of an individual of the group using the strategy when interacting with an individual of the group using the strategy. Thus the matrix with entries , where , and , is a square matrix of order , consisting of the block matrices .
Let . The state of the population is described by a point in the polytope
where , and the entry represents the usage frequency of the strategy within the group . We denote by the boundary of .
Assuming random encounters between individuals, for each group , the average payoff of a strategy within a population with state is given by
where the overall average payoff of group is given by
Demanding that the logarithmic growth rate of the frequency of each strategy , , is equal to the payoff difference between strategy and the overall average payoff of group yields the system of ordinary differential equations defined on the polytope ,
(4.1) |
that will be designated as a polymatrix replicator.
If equation (4.1) becomes the usual replicator equation with payoff matrix . When and are null matrices, equation (4.1) becomes the bimatrix replicator equation with payoff matrices and .
The flow of this equation leaves the polytope invariant. The proof of this fact is analogous to that for the bimatrix replicator equation, see [HS1998, Section 10.3]. Hence, by compactness of , the flow is complete. From now on the term polymatrix replicator will also refer to the flow and the underlying vector field on , denoted by .
Given , let
A set determines the facet of . The correspondence between labels in and facets of is bijective.
Remark 4.1.
The partition of into the interiors , with , is a smooth stratification of with strata . Every stratum is a connected open submanifold and for any pair if then . For more on smooth stratification see [FO] and references therein.
For a set consider the pair , where with , and .
Proposition 4.2.
[AD2015, Proposition ] Given , the facet of is invariant under the flow of and the restriction of (4.1) to is the polymatrix replicator .
For a fixed the correspondence is linear and its kernel consists of the matrices where each block has equal rows, i.e., has the form
Thus if and only if for every the matrix has equal rows (see [AD2015, Proposition 1]).
We have now the following characterization of the interior equilibria.
Proposition 4.3.
[AD2015, Proposition ] Given a polymatrix replicator , a point is an equilibrium of iff for all and .
In particular the set of interior equilibria of is the intersection of some affine subspace with .
Definition 4.4.
A polymatrix replicator is said to be conservative if there exists:
-
(a)
a point , called formal equilibrium, such that for all , and all and ;
-
(b)
matrices such that
-
(i)
,
-
(ii)
is a skew symmetric, and
-
(iii)
with for all .
-
(i)
The matrix will be referred to as a skew symmetric model for , and as a scaling co-vector.
In [ADP2015], another characterization of conservative polymatrix replicators, using quadratic forms, is provided. Furthermore, in [hassan-JGM-2020] the concept of conservative replicator equations (where ) is generalized using Dirac structures.
In what follows, the vectors in , or , are identified with column vectors. Let . We will omit the subscript whenever the dimension of this vector is clear from the context. Similarly, we write for the identity matrix. Given , we denote by the diagonal matrix . For each we define the matrix
and the block diagonal matrix .
Given an anti-symmetric matrix , we define the skew symmetric matrix valued mapping
(4.2) |
The interior of the polytope , denoted by , equipped with is a Poisson manifold, see [AD2015]*Theorem 3.5. Furthermore, we have the following theorem.
Theorem 4.5.
[AD2015]*Theorem Consider a conservative polymatrix replicator with formal equilibrium , skew symmetric model and scaling co-vector . Then , restricted to , is Hamiltonian with Hamiltonian function
(4.3) |
5. Asymptotic dynamics
Given a polymatrix replicator , the edges and vertices of the polytope form a (edge-vertex) heteroclinic network for the associated flow. In this section we recall the technique developed in [ADP2020] to analyze the asymptotic dynamics of a flow on a polytope along its heteroclinic edge network. In particular we review the main definitions and results for the polymatrix replicator on the polytope .
The affine support of is the smallest affine subspace of that contains . It is the subspace where for ,
Following [ADP2020, Definition 3.1] we introduce a defining family for the polytope . The affine functions where , form a defining family for because they satisfy:
-
(a)
,
-
(b)
for all , and
-
(c)
given such that , the linear -forms are linearly independent at every point .
Next we introduce convenient labels for vertices, facets and edges of . Let be the canonical basis of and denote by the Cartesian product which contains elements. Each label determines the vertex of . This labeling is one-to-one. The set can be used to label the facets of . Each integer labels the facet of . Edges can be labeled by the set . Given there exists a unique (unordered) pair of labels such that is the union of the strategies in and . The label determines the edge . Again the correspondence between labels and edges of is one-to-one.
Given a vertex of , we denote by and respectively the sets of facets and edges of that contain . Given
and this set of facets contains exactly elements.
Triples in
are called corners. Any pair of elements in a corner uniquely determines the third one. Therefore, sometimes we will shortly refer to a corner as or . An edge with end-points determines two corners and , called the end corners of . The facets are referred to as the opposite facets of .
Remark 5.1.
In a small neighborhood of a given vertex , where , the affine functions , , with , can be used as a coordinate system for .
Given a polymatrix replicator and a facet with , , the component of is given by
A polymatrix replicator is called non-degenerate if for any , the function ,
is not identically zero along .
Clearly generic polymatrix replicators are non-degenerate. Using the concept of order of a vector field along a facet [ADP2020, Definition ], is non-degenerate if and only if all facets of have order . From now on we will only consider non-degenerate polymatrix replicators.
Definition 5.2.
The skeleton character of polymatrix replicator is defined to be the matrix where
where stands for when with . For a fixed vertex , the vector is referred to as the skeleton character at .
Remark 5.3.
Given a corner of , is the eigenvalue of the tangent map along the eigen-direction parallel to .
Proposition 5.4.
If is a non-degenerate polymatrix replicator for every vertex with label , and every facet with and the skeleton character of is given by
Proof.
Straightforward calculation. ∎
Remark 5.5.
For a given corner of ,
-
•
if then is the -limit of an orbit in , and
-
•
if then is the -limit of an orbit in .
Let be an edge with end-points and and opposite facets and , respectively. This means that and are corners of . If does not have singularities in , then consists of a single heteroclinic orbit with -limit and -limit if and only if and . This type of edges will be referred to as flowing edges. The vertices and are respectively called the source and target of the flowing edge and we will write to express it. When the two characters the edge its called neutral.
Definition 5.6.
A polymatrix replicator is called regular if it is non-degenerate and moreover every edge is either neutral or a flowing edge.
Given with consider the vertex neighborhood
Rescaling the defining functions we may assume these neighborhoods are pairwise disjoint. See Remark 5.1.
For any edge with end-points and we define a tubular neighborhood connecting to by
Again we may assume that these neighborhoods are pairwise disjoint between themselves. Finally we define the edge skeleton’s tubular neighborhood of to be
(5.1) |
The next step is to define the rescaling map on . See [ADP2020, Definition ]. We will write to denote the affine function associated with the facet with .
Definition 5.7.
Let be a small parameter. The -rescaling coordinate system
maps to where
-
if for some vertex :
-
if for some edge :
We now turn to the space where these rescaling coordinates take values. For a given vertex we define
(5.2) |
where . Since is a coordinate system over and the function , , is a diffeomorphism, the restriction is also a diffeomorphism denoted by .
If is an edge connecting two corners and , and we define
(5.3) |
Then has dimension while has dimension . In particular the map is not injective over . See Figure 4.

Definition 5.8.
Hence .
Denote by the flow of the vector field . Given a flowing edge with source and target we introduce the cross-sections
transversal to the flow . The sets and are inner facets of the tubular neighborhoods and respectively. Let be the set of points such that the forward orbit has a first transversal intersection with . The global Poincaré map
is defined by , where
To simplify some of the following convergence statements we use the terminology in [ADP2020, Definition ].
Definition 5.9.
Suppose we are given a family of functions with varying domains . Let be another function with domain . Assume that all these functions have the same target and source spaces, which are assumed to be linear spaces. We will say that in the topology, to mean that:
-
(1)
domain convergence: for every compact subset , we have for every small enough , and
-
(2)
uniform convergence on compact sets:
Convergence in the topology means convergence in the topology for all . If is a composition of two or more mappings then its domain should be understood as the composition domain.
Let now
(5.4) |
and define
Notice that .
Lemma 5.10.
For a given , there exists a number such that the following limit holds in the topology,
where is the domain of .
Proof.
See [ADP2020, Lemma 7.2]. ∎
Hence, since the global Poincaré maps converge towards the identity map as we approach the heteroclinic orbit, the asymptotic behavior of the flow is solely determined by local Poincaré maps.
From Definition 5.2, for any vertex , the vector is tangent to , in the sense that belongs to the linear span of the sector . Let
(5.5) |
Using the notation of Definition 5.2 we have
Lemma 5.11.
We have
where
Moreover, given there exists such that the following limit holds in the topology
Proof.
See [ADP2020, Lemma ]. ∎
Consider a vertex with an incoming flowing-edge and an outgoing flowing-edge . Denote by the facet opposed to at . We define the sector
(5.6) |
and the linear map by
(5.7) |
Notice that as well as are facets to .
Proposition 5.12.
The sector consists of all points which can be connected to some point by a line segment inside the ray . Moreover, if then the other endpoint is .
Proof.
See [ADP2020, Proposition ]. ∎
Given flowing-edges and such that we denote by the set of points such that the forward orbit has a first transversal intersection with . The local Poincaré map
is defined by , where
Lemma 5.13.
Let be the domain of the map
Then for a given there exist such that
in the topology.
Proof.
See [ADP2020, Lemma ]. ∎
Given a chain of flowing-edges
the sequence is called a heteroclinic path, or a heteroclinic cycle when .
Definition 5.14.
Given a heteroclinic path :
-
1)
The Poincaré map of a polymatrix replicator along is the composition
whose domain is denoted by .
-
2)
The skeleton flow map (of ) along is the composition map defined by
whose domain is
The previous lemmas 5.10 and 5.13 imply that given a heteroclinic path , the asymptotic behavior of the Poincaré map along is given by the Poincaré map of .
Proposition 5.15.
Let be the domain of the map
from into . Then
in the topology.
Proof.
See [ADP2020, Proposition ]. ∎
To analyze the dynamics of the flow of the skeleton vector field we introduce the concept of structural set and its associated skeleton flow map. See [ADP2020, Definition ].
Definition 5.16.
A non-empty set of flowing-edges is said to be a structural set for if every heteroclinic cycle contains an edge in .
Structural sets are in general not unique. We say that a heteroclinic path is an -branch if
-
(1)
,
-
(2)
for all .
Denote by the set of all -branches.
Definition 5.17.
The skeleton flow map is defined by
where
The reader should picture as the first return map of the piecewise linear flow of on to the system of cross-sections . The following, see [ADP2020, Proposition ], provides a sufficient condition for the skeleton flow map to be a closed dynamical system.
Proposition 5.18.
Given a skeleton vector field on with a structural set , assume
-
(1)
every edge of is either neutral or a flowing-edge,
-
(2)
every vertex is of saddle type, i.e., for some facets .
Then
is a Baire space with full Lebesgue measure in and is a homeomorphism.
Given a structural set any orbit of the flow that shadows some heteroclinic cycle must intersect the cross-sections recurrently. The following map encapsulates the semi-global dynamics of these orbits.
Definition 5.19.
Given , let be a structural set of its skeleton vector field. We define setting , and for all . The domain components and are disjoint for branches in .
Up to a time reparametrization, the map embeds in the flow . In this sense the dynamics of encapsulates the qualitative behavior of the flow of along the edges of .
Theorem 5.20.
Let be a regular polymatrix replicator with skeleton vector field . If is a structural set of then
in the topology, in the sense of Definition 5.9.
Proof.
See [ADP2020, Theorem ]. ∎
6. Hamiltonian character of the asymptotic dynamics
In this section we discuss the Poisson geometric properties of the Poincaré maps in the case of Hamiltonian polymatrix replicator equations. Given a generic Hamiltonian polymatrix replicator, , we study its asymptotic Poincaré maps, proving that they are Poisson maps.
Let be a conservative polymatrix replicator, a formal equilibrium, and as in Definition 4.4, and
(6.1) |
its Hamiltonian function as in Theorem 4.5. The Hamiltonian (6.1) belongs to a class of prospective constants of motion for vector fields on polytopes discussed in [ADP2020, Section ]. Since the polymatrix replicator is fixed we drop superscript “” and use for the rescaling coordinate systems defined in Definition 5.7. The following proposition gives the asymptotic constant of motion, on the dual cone, associated
Proposition 6.1.
Given defined by
(6.2) |
-
(1)
over for any vertex , with convergence in the topology;
-
(2)
over for any vertex , with convergence in the topology;
-
(3)
Since is invariant under the flow of , i.e., , the function is invariant under the skeleton flow of , i.e., .
Proof.
See [ADP2020, Proposition ]. ∎
We will use the following family of coordinate charts for the Poisson manifold where is defined in (4.2).
Definition 6.2.
Given a vertex of , we set and , and define the projection map
is a diffeomorphism onto its image and the inverse map can be regarded as a local chart for the manifold .
Remark 6.3.
The projection map extends linearly to and it is represented by the block diagonal matrix
where , for , is the constant matrix obtained from the identity matrix by removing its row .
Using the definitions of and given in Section 4 we can state the following lemma.
Lemma 6.4.
Consider the Poisson manifold where is defined in (4.2). Then for any vertex , the matrix representative of in the local chart is
(6.3) |
Proof.
Notice that with and , . ∎
We used the notation instead of to make it clear that the representing matrix is with respect to the local chart . The following trivial lemma gives us the differential of the -rescaling map (in Definition 5.7) for the coordinate chart . Given a vertex and using the notation introduced in Definition 6.2 we write and denote by the diagonal matrix .
Lemma 6.5.
The differential of the diffeomorphism
is given by
We push forward, by the diffeomorphism , the Poisson structure defined on to . The following lemma provides the matrix representative of the push forwarded Poisson structure. In order to simplify the notation we set
(6.4) |
and for every
(6.5) |
Notice that .
Lemma 6.6.
The diffeomorphism pushes forward the Poisson structure to the Poisson structure on where
(6.6) |
The Poisson structure is asymptotically equivalent to a linear Poisson structure. Let
(6.7) |
be the matrix defined by diagonal blocks , for , where the block is the matrix in which the column is equal to and every other column is equal to .
Lemma 6.7.
Given a vertex , if is the matrix in (6.7) and , then
over with convergence in topology. Consequently,
over with convergence in topology.
Proof.
A simple calculation shows that for every
For every and we have
Considering that , we get the first claim of the lemma and the second claim is an immediate consequence. ∎
Figure 5 illustrates the case .
Remark 6.8.
The same linear Poisson structure appears in [AD2015, Theorem ].
Lemma 6.9.
Proof.
We use the notation for the local expression of the replicator vector field in the local chart . If we write the function , defined in (6.1), as then
Notice that . By Theorem 4.5, . Locally,
Similarly, writing we have
The vector field defined in Lemma 5.11 is
where in the second equality we use . Then, applying Lemma 5.11, Lemma 6.7, and Proposition 6.1, the result follows. Notice that . ∎
Our aim is to show that for a given heteroclinic path , the skeleton flow map of along (see Definition 5.14),
restricted to the level set of , is a Poisson map. Notice that the Poisson structure is only defined in and neither nor are submanifolds of . So we need to define Poisson structures on the sections for all .
For the heteroclinic path
(6.8) |
we store in the column of the matrix
the indices of the non zero components of the vertex . By construction of , there exists such that for and , i.e. is the group containing the nonzero component that differ between the end points of the edge . In order to simplify notations for every vertex in we denote
First we consider the vertex with incoming and outgoing edges
respectively, i.e. , where . Notice that
Since for and we have
The opposite facet to at is then where we omitted the superscript from since it is evident that . We keep omitting the superscript whenever there is no ambiguity. The sector defined in (5.6) is
(6.9) |
The skeleton flow map of at vertex is the linear map defined by
(6.10) |
Notice that where is the flow of the skeleton vector field and . We denote where . More precisely
Definition 6.10.
We define by
(6.11) |
the convex cone containing the line segments of the flow of connecting the points in the domain of to their images.
We consider two cosymplectic foliations interior to each sector in order to use the techniques introduced in Section 3. In the following lemma, we describe the Poisson structures on and .
Lemma 6.11.
With the notation adopted in Lemma 6.9, let be the restriction of function , defined in (6.2), to . Consider two functions defined by and then:
-
1)
Level sets of partition into a cosymplectic foliation and , i.e. every leaf of these foliations is a cosymplectic submanifold of . Furthermore, every leaf of these foliations is a level transversal section to at every point ;
-
2)
Given two leafs222Notice that the flow of preserves . , of and two leafs , of , then the Poincaré map between any pair of these four leafs is a Poisson map.
Proof.
Clearly,
and similarly . As before, using the notation for the facet , we see that is a flowing edge from the corner to the corner . So . In a similar way we have . What we actually need is both of them to be nonzero. Then, both and are second class constraints and consequently, their level sets are cosymplectic submanifolds (see Definition 3.4). The fact that is a level transversal section is clear.
The Poincaré map between is the translation
(6.12) |
and a similar translation for . Clearly, these translations are Poisson maps.
The Poincaré map between two level sets and is
(6.13) |
By Proposition 3.11 this map is a Poisson map as well. ∎
Remark 6.12.
Note that is not constant on , so the map (6.13) is not a fixed time map of the flow . Therefore, being Poisson is not a direct consequence of the flow being Hamiltonian. Furthermore, proving that this map is Poisson by direct calculation is not straightforward. This makes the contents of Section 3 inevitable.
Since the Poincaré maps can be considered between level sets of the functions and , we state the following definition.
Definition 6.13.
Let and be the foliations constituted by the level sets of the functions and , respectively.
Every leaf of is equipped with a Poisson structure, which has as a Casimir, and the level sets of this Casimir are the leafs of the cosymplectic foliation . The leafs of can be identified (as Poison manifolds) through translations of type (6.12).
Definition 6.14.
By we denote a typical leaf of the Poisson foliation .
Ignoring (for a moment) the fact that the function is only defined on , we may consider as the zero level set of . Hence a typical leaf is diffeompric to through a translation of type (6.12). Through this, diffeomorphism secures a Poisson structure. Similarly, gains a Poisson structure from .
Proposition 6.15.
Proof.
We decompose into three maps , and , where and are the translations used to define the Poisson structures on and , respectively, and is the Poincaré map from to . By the construction of these two sections, together with Lemma 6.11, is a Poisson map, which ends the proof. ∎
Notice that is a union of Poisson submanifolds equipped with Dirac bracket. We describe now the matrix representative of this Dirac bracket.
Lemma 6.16.
The matrix representative, in the coordinate system , of the Dirac bracket generated in by the second class constrains and is
(6.14) |
where with and
In the matrix the line and column are null. Removing these line and column one obtains the matrix representative, in the coordinate system obtained by removing from , of the Poisson structure on . Similarly, for the second class constrains and we have
(6.15) |
where
and removing the line and column yields the matrix representative, in the coordinate system obtained by removing from , of the Poisson structure on .
Proof.
By definition,
So we need to calculate
(6.16) |
see the definition of in (3.4). Reminding that
together with a simple calculation, yields (6.14). The line and column are zero simply because, by definition, is a Casimir of the Dirac bracket. Note that the representative matrix is with respect to the coordinate system as of , and by omitting the component from this coordinate system one obtains a coordinate system on . Therefore, removing the null line and column yields the representative matrix of with respect to the obtained coordinate. The same reasoning holds for . ∎
We now extend Proposition 6.15 to the whole heteroclinc path . Our main result is the following.
Theorem 6.17.
Let
(6.17) |
be a heteroclinic path. Then for every , the Poisson structures induced on the intersection
(6.18) |
from Poisson submanifolds and is the same. Consequently, the skeleton flow map of along (see Definition 5.14),
is a Poisson map w.r.t. the Poisson structures induced by and on its domain and range, respectively.
Considering the segment
the key point is to show that the Poisson structure induced from on and the one induced from on , match on the intersection (6.18) (see Figure 6). To prove Theorem 6.17 we need to state and prove two preliminary lemmas regarding this key point.
The two sectors and are only different in the group , where for the elements of and for the elements . Let be the diffeomorphism of the form
where:
-
1)
For the associated component is the identity map;
-
2)
For any
-
3)
For the following notation to be consistent, without loss of generality we assume that . Notice that for any given point the map acts on the component as
where the notation means that the entry is missing in the corresponding vector.
The image point is not in . However composing with the translation
we get
We restrict the diffeomorphism to an open set around to get
where is an open set around .
Lemma 6.18.
The diffeomorphism
is Poisson, i.e. preserves the ambient Poisson structure.
Proof.
Lemma 6.19.
For the diffeomorphism we have that:
-
1)
-
2)
Proof.
The first equality is trivial since, for any , the component of is . For the second equality we have
Then, using the fact that we get
∎
Proof of Theorem 6.17:.
By Lemma 6.18
Since are second class constraints, then
(6.19) |
are also second class constraints. Considering the equalities obtained in Lemma 6.19, this fact can be obtained by direct calculations and being second class constraints. Furthermore, the Dirac structure on generated by the second class constraints is the same as the one generated by (6.19). To see this, note that the foliation constituted from the level sets of is the same as the one made up from the level set of the constraints (6.19). Also, Dirac bracket (see (3.4)) defined by them is the same, since the second term in Definition (3.4) is the same whether it is computed using the constraints or the constraints (6.19). Simply compare the following equations
where . The constant terms are ignored and we used the fact that to simplify the middle term in the second equation.
We conclude that the diffeomorphism , in addition to preserving the ambient Poisson structures, preserves the Dirac brackets as well, and consequently
Let and be the translations as defined in the proof of Proposition 6.15. The restriction map is also a translation, so there exists a vector such that the following diagram is commutative.
This shows that the Poisson structures coming from different sides of match and we can compose the Poisson map for . This finishes the proof. ∎
For a given edge if there are more than one edge going out from the vertex , say , with , the are disjoint open subsets of . Considering all these disjoint Poisson submanifold all together we can state the following result whose proof is immediate from the previous results.
Theorem 6.20.
Let denote the set of all -branches of the skeleton vector field (see Definition 5.16) and set to be the open submanifold of
with the same Poisson structure. Then the skeleton flow map is Poisson.
7. Example
We will now present an example of a Hamiltonian polymatrix replicator system with a non trivial dimension. This example was chosen to provide an illustration of the concepts and main results of this paper. In particular it has a small structural set with a simple heteroclinic network.
7.1. The fish example
Consider the polymatrix replicator system defined by matrix
We denote by the vector field associated to this polymatrix replicator that is defined on the polytope
The point
satisfies
-
(1)
;
-
(2)
and ,
where stands for the -th component of vector , and hence is an equilibrium of (see Proposition 4.3). Since matrix is skew-symmetric, the associated polymatrix replicator is conservative (see Definition 4.4).
The polytope has seven facets labeled by an index ranging from to , and designated by . The vertices of the phase space are also labeled by , and designated by , as described in Table 1.
Vertex | |
---|---|
Vertex | |
---|---|
The edges of are designated by , according to Table 3, where we write to mean that is an edge connecting the vertices and . This model has edges: neutral edges,
and flowing-edges,
The flowing-edge directed graph of is depicted in Figure 7.
From this graph we can see that
is a structural set for (see Definition 5.16) whose -branches denoted by are displayed in Table 4, where we write to indicate that is a path from vertex passing along vertices .

From\To | |
---|---|
Considering the vertex , which has the incoming edge and the outgoing edge , we will now illustrate Proposition 6.15.
For , the constant Poisson structures induced by asymptotic rescaling on each (see Lemma 6.7) can be easily calculated:
and
The matrix represents the Poisson structure on in the coordinates . Notice that on . Similarly, the matrix represents the Poisson structure on in the same coordinates . Notice again that on . Now the matrix representative of in the coordinates is
A simple calculation shows that
confirming the fact that the asymptotic Poincaré map is Poisson (see (3.3) in Definition 3.1).
Consider now the subspaces of
and
For the given matrix , its null space has dimension . Take a non-zero vector . For example,
The set of equilibria of the natural extension of to the affine hyperplane is
The Hamiltonian of is the function
where is the -th component of the equilibrium point (see Theorem 4.5). Another integral of motion of is the function
where is the -th component of , which is a Casimir of the underlying Poisson structure.
Consider the skeleton flow map of (see Definition 5.17). Notice that , where by Proposition 5.18, . By Proposition 6.1 the function is invariant under . Moreover, the skeleton flow map is Hamiltonian with respect to a Poisson structure on the system of cross sections (see Theorem 6.17).
For all , the polyhedral cone has dimension . Hence, each polytope is a -dimensional polygon.
Remark 7.1.
We came from dimension to . This will happen for any other conservative polymatrix replicator with the same number of groups and the same number of strategies per group. In fact when is odd, where is the total number of strategies in the population and is the number of groups, we will have a minimum drop of dimensions. The reason is that a Poisson manifold with odd dimension (in this example is ) has at least one Casimir, and considering the transversal section we drop two dimensions from the symplectic part (not from the Casimir). So in total we drop a minimum of 3 dimensions. If the original Poisson structure has more Casimirs, the invariant submanifolds yielded geometrically, are going to have even less dimensions, which is good as long as it not zero. In the case of an even dimension, the drop will be at least of two dimensions.
By invariance of , the set is also invariant under . Consider now the restriction of to . This is a piecewise affine area preserving map. Figure 8 shows the domain and iterates by of a point in . Following the itinerary of a random point we have picked the following heteroclinic cycle consisting of -branches
The map is represented by the matrix
The eigenvalues of , besides and (with geometric multiplicity and , respectively), are
Remark 7.2.
The determinant of is zero which means that the Poisson structure on is non-degenerate. So, has a two dimensional symplectic foliation invariant under the asymptotic Poincaré map. The leaf of this foliation are affine spaces parallel to the kernel of
i.e. the set of the form
The restriction of the asymptotic Poincaré map to these leaves is a symplectic map. One important consequence is that its eigenvalues are of the form and .

An eigenvector associated to the eigenvalue is
We have chosen so that , i.e., . In fact we have . Hence is a periodic point of the skeleton flow map with period (whose iterates are represented by the green dots in Figure 8).
Figure 8 also depicts the polygons contained in , and the orbit of another periodic point of the skeleton flow map with period (represented by the blue dots in Figure 8).
Following the procedure to analyze the dynamics in [ADP2020, Section ] and using Theorem also in [ADP2020] we could deduce the existence of chaotic behavior for the flow of in some level set , with the chosen above and for all small enough .
Acknowledgements
The first author was supported by mathematics department of UFMG. The second author was supported by FCT - Fundação para a Ciência e a Tecnologia, under the projects UIDB/04561/2020 and UIDP/04561/2020. The third author was supported by FCT - Fundação para a Ciência e a Tecnologia, under the project UIDB/05069/2020.