1,035
Views
5
CrossRef citations to date
0
Altmetric
Research Article

Lotka–Volterra coalgebras

, &
Pages 4483-4497 | Received 15 Sep 2020, Accepted 12 Jan 2021, Published online: 03 Feb 2021

Abstract

We study Lotka–Volterra coalgebras, a new family of nonassociative coalgebras emerging from population genetics. We focus on their defining algebraic properties and deal with characterizing the existence of counital and character maps. We also provide a classification of their basis elements into (co)algebraically persistent and transient generators resulting in a semidirect sum decomposition of Lotka–Volterra coalgebras.

1. Introduction

Lotka–Volterra algebras were introduced by Itoh [Citation1] as a nonassociative framework for describing ternary interactions in competition systems. Despite being aimed by random models mirroring the kinetic theory of gases, a fact remarked by Itoh in [Citation1], these algebras were later shown to include, as particular examples, normal Bernstein algebras. Bernstein algebras became essential in the study of quadratic stochastic operators in population genetics [Citation2].

Given a field K of characteristic not 2, a commutative K-algebra A is Lotka–Volterra if it admits a basis B={e1,,en} such that eiej=ejei=(12+aij)ei+(12+aji)ej, for all i, j = 1, …, n, where A=(aij)i,j=1n is a skew-symmetric matrix. We refer to [Citation3, Citation4] and references therein for the algebraic treatment of Lotka–Volterra algebras. Alternatively, Lotka–Volterra algebras can be defined as follows:

Definition 1.1

A K-algebra A is Lotka–Volterra (for short, an LV-algebra) if it is endowed with a basis B={e1,e2,,en} and multiplication eiej=k=1nβijkek, such that for any 1 ≤ i, j, k ≤ n:

  • k{i,j}βijk=0,

  • βijk=βjik,

  • k=1nβijk=1.

From the genetic viewpoint Lotka–Volterra algebras reproduce the forward evolution of populations where offspring necessarily reproduce the genetic type of one of the progenitors, assuming also panmixia, that is, considering that no progenitor has any specific weight on offspring type. As a result, LV-algebras are, by definition, commutative. Notice however that nonnegativity of the multiplication constants has been later not required [Citation3] as it happens to be the case for (nonassociative) algebras with genetic realization [Citation5]. This was not however the case in [Citation1], where the requirement |aij|12 on the skew-symmetric matrix entries ensured the genetic realization of Itoh's Lotka–Volterra algebras.

Lotka–Volterra algebras are a family of genetic algebras different from the recently studied evolution algebras [Citation6], although both families of algebras, Lotka–Volterra and evolution, are defined by square, not cubic, structure matrices. The simplest example of Lotka–Volterra algebra is the gametic algebra for simple Mendelian inheritance [Citation7, p.135], also called an elementary algebra by Heuch and Holgate, given by the zero matrix A = 0n×n

Our aim here is, retaining those properties characterizing the genetic population under consideration, i.e. offspring reproducing some progenitor type and panmixia, to consider Tian and Li's approach based on the search for new algebraic structures focused on describing the backward inheritance within genetic systems [Citation8]. As a result, Lotka–Voterra coalgebras are introduced, providing a new framework for the treatment of such populations.

Lotka–Volterra coalgebras constitute therefore a new family of noncoassociative coalgebras whose defining conditions reflect the backward dynamics of the transference of genetic inheritance within the considered populations. Even though here we are mainly concerned with the study of their algebraic structure, the connection between Lotka–Volterra coalgebras and the theory of dynamical systems will also become apparent.

After this introduction, the paper is organized as follows. Section two provides the definition of Lotka–Volterra coalgebras, together with examples, and also focuses on their matrix realization. In the third section, we cope with the duality of Lotka–Volterra structures. A well-known result states the correspondence, through dual vector spaces, between algebraic and coalgebraic structures defined on finite-dimensional vector spaces. Here we examine up to what extent the duality between Lotka–Volterra algebras and coalgebras is retained, to continue in Section four tackling with the algebraic description of Lotka–Volterra coalgebras.

Section five is devoted to a rather classical problem in algebraic genetics [Citation7, p.103], namely here, discussing which (coalgebra) constructions retain the property of being Lotka–Volterra. To finish the algebraic description of Lotka–Volterra coalgebras we achieve in this work, Section six focuses on their algebraic structure. Distinguishing generators of Lotka–Volterra coalgebras (i.e. their natural basis elements) into group-like and nongroup-like elements provides a semidirect sum decomposition (see [Citation6, p.45]) of Lotka–Volterra coalgebras into a subcoalgebra spanned by algebraic persistent elements and a coideal spanned by algebraic transient elements. This classification of the coalgebra natural basis elements brings into a linkage between Lotka–Volterra coalgebras and dynamical systems, to be exploited in a forthcoming work.

2. Lotka Volterrra coalgebras

Throughout the paper, unless otherwise stated, K will be a field of characteristic not 2, and all objects will be defined over finite-dimensional K-vector spaces. We refer the reader to [Citation9] for basic notions on coalgebras.

A coalgebra (C,Δ) is a K-vector space C endowed with a linear map (comultiplication) Δ:CCC. Given a basis B={e1,e2,,en} of the K-vector space C, we write Δ(ek)=i,j=1nβijkeiej,k=1,,n.A subcoalgebra D of a coalgebra C is a subspace D of C such that Δ(D)⊆ D ⊗ D. A vector subspace I of C is a left (resp. right) coideal of C if Δ(I)CI (resp. Δ(I)IC) and a coideal if Δ(I)IC+CI. Subcoalgebras are one-sided coideals and, one-sided coideals are coideals. Intersections and sums of subcoalgebras (resp. left or right coideals) are again subcoalgebras (resp. left or right coideals). Sums of coideals are also still coideals, but this does not apply to intersections.

The comultiplication structure constants can be rearranged into a cubic (n × n × n) matrix P=(pijk)i,j,k=1nKn×n×n, given by pijk=βijk. Cubic matrices are usually unfolded by their frontal slices as n × n2 rectangular matrices P = (P::1|P::2| · · · |P::n), where, for any k = 1, …, n, P::k is the n × n matrix with (i, j)-th entry pijk, i, j = 1, …, n. Thus the k-th frontal slice P::k of P contains the comultiplication constants arising from Δ(ek).

Definition 2.1

A coalgebra (C,Δ) is Lotka–Volterra (for short, an LV-coalgebra) if it admits a basis B={e1,e2,,en}, satisfying for all i, j, k = 1, …, n:

  • k{i,j}βijk=0;

  • βijk=βjik;

  • i,j=1nβijk=1;

where {βijk}i,j,k=1n is a set of scalars such that Δ(ek)=i,j=1nβijkeiej,k=1,,n.

Remark 2.1

Any such a basis B will be called a natural basis. All bases considered in the paper will be assumed to be natural bases. Because of LCV-1 and LCV-2 we have that Δ(ek)=βkkkekek+i=1,iknβikk(eiek+ekei),k=1,,n.Moreover, the above relation and LCV-3 imply βkkk+2(i=1,iknβikk)=1,for allk=1,,n,or, equivalently, i=1,iknβikk=12(1βkkk).

Definition 2.2

A real (i.e. K=R) LV-coalgebra (C,Δ), with natural basis B, has genetic realization if:

  • βijk0 for all i, j, k = 1, …, n.

Example 2.3

Let G be a finite group of order n. Then the group coalgebra C=KG is an LV-coalgebra with natural basis B={eggG} and βijk=δkiδkj, for all i, j, k = 1, …, n.

Definition 2.4

An LV-coalgebra C will be said to be trivial if βijk=δkiδkj, for all i, j, k = 1, …, n., i.e. if C is a group coalgebra.

Example 2.5

Tian and Li [Citation8, p. 248] considered the case of a randomly mating population of diploid individuals differing in a locus alleles B={e1,e2,,en}. The n-dimensional K-vector space C with basis B and comultiplication given by βijk=12n(δki+δkj), for all i, j, k = 1, …, n, is an LV-coalgebra. Following [Citation7, Example 1.1, p. 8], we refer to C to be the n-dimensional gametic coalgebra for simple Mendelian inheritance.

The class of LV-coalgebras is a family of (genetic) coalgebras not containing the zygotic coalgebra or the gametic coalgebra of autotetraploids [Citation10]. Moreover, an evolution coalgebra C is Lotka–Volterra if and only if C is a trivial LV-coalgebra, that is, a group coalgebra [Citation11].

Displayed by its frontal slices, the cubic matrix associated to an LV-coalgebra has the following form: P=(β111β121β1n10β122000β1nnβ21100β212β222β2n200β2nnβn11000βn220βn1nβn2nβnnn),which, by LVC-2, is (1,2)-symmetrical with equal accompanying matrices [Citation10, Lemma 2].

Proposition 2.6

Let C be an LV-coalgebra. Then, with respect to a natural basis B, the accompanying matrices of C (i.e. those of its associated cubic matrix) are P(i)=P(j)=(i=1nβi11β122β1nnβ121i=1nβi22β2nnβ1n1βn22i=1nβinn).

Proof.

See [Citation10, 3.3].

Mimicking the algebra setting, LV-coalgebras are proposed to model backward dynamics in genetic systems when offspring necessarily reproduce the genetic type of one progenitor (LVC-1). Panmixia, i.e. considering that paternal or maternal type have no specific weight on offspring type, is also assumed, resulting in assumption LVC-2. Finally, for real LV-coalgebras, having genetic realization amounts to: βikk=βkik={P(one progenitor of type ioffspring of type k),ik;P(both progenitors of typei=koffspring of type k),i=k.

3. Duality

As remarked in the introduction, when posing the problem of introducing coalgebras as an alternative mathematical framework for modelling backward genetic inheritance, Tian and Li [Citation8] stressed on the fact that looking for new coalgebraic structures should imply more than just dualizing already existing genetic-like algebras. In this section, it is shown that, as previously defined, LV-coalgebras give rise to a new family of algebraic structures playing their own role in the mathematical study of population genetics.

The dual vector space C=HomK(C,K) of any coalgebra (C,Δ) is endowed with an algebra structure with multiplication induced by Δ*. Conversely, for any finite dimensional algebra A the isomorphism (AA)AA allows us to define a coalgebra structure on A [Citation9].

Let C be a n-dimensional K-vector space with basis B={e1,,en}. We denote by B={f1,,fn} the dual basis of B in C, i.e. fi(ej) = δij, for all i, j = 1, …, n.

Theorem 3.1

Let (C,Δ) be a coalgebra satisfying LVC-1 and LVC-2 (with respect to B). Then the induced multiplication in C is given by: {frfr=βrrrfr;frfs=fsfr=βsrrfr+βrssfs,rs.

Proof.

Use Δ*(fr ⊗ fs)(ek) = (fr ⊗ fs)Δ(ek) to obtain: (frfs)(ek)={0;k{r,s};βksk=βrsr;k=rs;βrkk=βrss;k=sr;βkkk;k=r=s;for all r, s, k = 1, …, n.

Remark 3.1

Note the multiplication in the dual algebra C is indeed defined by frfs=k=1nβrskfk, for all r, s = 1, …, n.

Corollary 3.2

Let (C,Δ) be an LV-coalgebra with genetic realization. Then C is not an LV-algebra (with respect to B).

Proof.

Assume C with the induced multiplication (see Theorem 3.1) is an LV-algebra. Then we have βkkk=1 for all k = 1, …, n, and LVC-3 and LVC-4 together imply that βrsk=δkrδks. Hence frfs = fsfr = 0 if r ≠ s, contradicting then that C is LV-algebra.

We remark that, given an LV-coalgebra C with genetic realization whose dual algebra C is also an LV-algebra Corollary 3.2 shows that C is then a group coalgebra, i.e. a trivial LV-coalgebra, and C a (nonzero) trivial evolution algebra.

Next given a coalgebra C we provide necessary and sufficient conditions for its dual algebra C to be an LV-algebra.

Theorem 3.3

Let (C,Δ) be a coalgebra with basis B={e1,,en}. Then, there exists a skew-symmetric matrix A=(aij)i,j=1n such that Δ(ek) = ek ⊗ Tk + Tk ⊗ ek, where Tk=i=1n(12+aki)ei if and only if (C,Δ) is an LV-algebra with respect to the dual basis B.

Proof.

Let (C,Δ) be a coalgebra with comultiplication Δ(ek) = ek ⊗ Tk + Tk ⊗ ek, with Tk=i=1n(12+aki)ei defined by a skew-symmetric matrix A=(aij)i,j=1n. We note this results into: {βkkk=1;βksk=12+aks;s=1,,n;sk;βrsk=0;otherwise.Then, by Remark 3.1, frfs=k=1nβrskfk=βrsrfr+βrssfs=(12+ars)fr+(12+asr)fs, for all r, s = 1, …, n. Moreover, C is commutative and, since A is skew-symmetric, C is an LV-algebra.

Assume conversely (C,Δ) is a coalgebra, whose dual algebra C with the inherited multiplication m = Δ* is a (commutative) LV-algebra, with respect to B, and write, for all r, s = 1, …, n: m(frfs)=frfs=(12+ars)fr+(12+asr)fsfor a skew-symmetric matrix A=(ars)r,s=1n. We consider the obvious identifications (C)C and (Δ*)* ≅ Δ resulting from the finite-dimensionality of C [Citation9, p.10]. Then Δ(ek)(fr ⊗ fs) = <ek, frfs > implies Δ(ek)=r,s=1nβrskeres, with βrsk=(12+ars)δkr+(12+asr)δks, for all k, r, s = 1, …, n and, as a result:

  1. If k{r,s}, then βrsk=0,

  2. If k = r ≠ s, then βrsk=βksk=12+aks,

  3. If k = r = s, then βrsk=βkkk=1.

Moreover βrsk=βsrk for all k, r, s = 1, …, n. Thus, taking into account that akk = 0, Δ(ek)=ekek+r=1,rknβkrkeker+r=1,rknβrkkerek=ek(12+akk)ek+(12+akk)ekek+ek(r=1,rkn(12+akr)er)+(r=1,rkn(12+akr)er)ek=ek(r=1n(12+akr)er)+(r=1n(12+akr)er)ek=ekTk+Tkek.Hence the comultiplication in C is therefore given by the skew-symmetric matrix A defining the elements Tk's.

We note that, as defined in Theorem 3.3, (C,Δ) satisfies LVC-1 and LVC-2, but not LVC-3. Indeed we have i,j=1nβijk=n for all k = 1, …, n, that is, comultiplication of the basis elements is not normalized.

Example 3.4

Let A be a 3-dimensional K-vector space with basis B={e1,e2,e3}. We define the following multiplication in A: eiei=ei,i=1,2,3,e1e2=e2e1=e2,e1e3=e3e1=12e1+12e3,e2e3=e3e2=e2;and extend it by linearity to m:AAA. Let us denote MM3×9(K) the matrix associated to the multiplication m (as a linear map) in bases B={eij=eiej}i,j=13 of AA and B of A. Writing B={bi}i=19={b1=e1e1,b2=e1e2,,b9=e3e3}, we have m(bi)=k=13mkiek, for all i = 1, …, 9, and M=(101/20001/200010111010001/20001/201).Then Δ:A(AA)AA defines a coalgebra structure on A, with associated matrix (as a linear map) the transpose matrix MT of M. Moreover (A,Δ) satisfies LVC-1 and LVC-2 but not LVC-3, since columns of MT are not normalized. Hence (A,Δ) is not an LV-coalgebra. If, assuming then char(K)5, we column-normalize MT, rows of the resulting matrix M~=(1/201/40001/40001/501/51/51/501/50001/40001/401/2)i.e. columns of M~T, define a new comultiplication Δ¯:AAA given by: Δ¯(f1)=12f1f1+14(f1f3+f3f1),Δ¯(f2)=15f2f2+15(f1f2+f2f1+f2f3+f3f2),Δ¯(f3)=12f3f3+14(f1f3+f3f1),also satisfying LVC-3. Hence (A,Δ¯) is an LV-coalgebra.

4. Algebraic properties of Lotka–Volterra coalgebras

Cocommutativity of LV-coalgebras follows from LVC-2. Recall a coalgebra C is cocommutative if τΔ = Δ, where τ:CCCC is the twist map given by τ(a ⊗ b) = b ⊗ a for all a,bC. LV-coalgebras are however not necessarily coassociative.

Example 4.1

Let C be a 2-dimensional coalgebra with comultiplication (with respect to a basis B={e1,e2}): Δ(e1)=e1e1,Δ(e2)=12e2e2+14(e1e2+e2e1).Clearly (C,Δ) is an LV-coalgebra and (id ⊗ Δ)Δ(e2) ≠ (Δ ⊗ id)Δ(e2). Hence C is not coassociative.

Definition 4.2

A coalgebra C is counital if it exists a linear map (counit) ε:CK such that (εid)Δ=id=(idε)Δ.

Proposition 4.3

Let C be an LV-coalgebra. Then C is counital if and only if there exists a nonzero a=(γ1,,γn)TKn such that, for all k = 1, …, n, (1) i=1nβikkγi=1,(1) (2) γk(1βkkk)=0.(2)

Proof.

Assume ε:CK is a (nonzero) counit for C and write ε(ei)=γi, i = 1, …, n. Then, by LVC-2, we have: ek=(idε)Δ(ek)=(εid)Δ(ek)=(εid)(βkkkekek+i=1,iknβikk(eiek+ekei))=βkkkγkek+i=1,iknβikk(γiek+γkei)=(i=1nβikkγi)ek+i=1,iknβikkγkei.Thus, ε is a counit of C if and only if, for all k = 1, …, n it holds that i=1nβikkγi=1,andγki=1,iknβikk=0;or equivalently, by LVC-3, i=1nβikkγi=1,andγk(1βkkk)=0.Conversely, any such nonzero a=(γ1,,γn)TKn defines a (nonzero) counit for C.

Example 4.4

  1. Let C=KG be a trivial LV-coalgebra, i.e. a group coalgebra. Since then we have βijk=δkiδkj for all i, j, k = 1, …, n, γk(1βkkk)=0 holds straightforwardly for all k, and i=1nβikkγi=1 becomes βkkkγk=1. Thus, by Proposition 4.3, C=KG has a unique counit defined by ε(ek)=1, for all k = 1, …, n.

  2. The 2-dimensional LV-coalgebra defined in Example 4.1 admits no nonzero counit.

It follows from Example 4.4(ii) above that having natural basis elements that are group-like is not enough for ensuring LV-coalgebras to be counital. However, having a nonzero counit is a sufficient condition for the existence of group-like elements (i.e. elements xC such that Δ(x) = x ⊗ x) in the case of LV-coalgebras with genetic realization.

Theorem 4.5

Let C be an LV-coalgebra with genetic realization and natural basis B={e1,,en}. If C has a nontrivial counit ε, then D=spanR(ekkΛ), where Λ={kε(ek)=γk0} is a (group) subcoalgebra of C. Moreover γk = 1 for all k ∈ Λ.

Proof.

Let ε be a nonzero counit for C. Then Λ={kε(ek)=γk0}. Take k0 ∈ Λ. Then, by Proposition 4.3(2), we have γk0(1βk0k0k0)=0, which implies βk0k0k0=1 and, therefore, by LVC-4, βrsk0=δk0rδk0s, for all r, s = 1, …, n. Thus Δ(ek0)=ek0ek0 and ek0 is a group-like element, and, as a result, D=spanR(ekkΛ) is a subcoalgebra of C. The last assertion, γk = 1 for all k ∈ Λ, follows from Proposition 4.3(1).

Definition 4.6

A coalgebra (C,Δ) is baric if there exists a nonzero linear map ϕ:CK such that (ϕ ⊗ ϕ)Δ = ϕ. The map ϕ is called character or weight function.

Any character ϕ is an element of C=HomK(C,K) and therefore can be uniquely written ϕ=i=1nαiei, for some a=(α1,,αn)TKn.

Proposition 4.7

Let C be an LV-coalgebra. A linear map ϕ=i=1nαieiC is a character of C if and only if (3) αk=αk(βkkkαk+2i=1,iknβikkαi),forallk=1,,n.(3)

Proof.

Let ϕ=i=1nαieiC. Then ϕ is a character of C if and only if, for all ekB αk=ϕ(ek)=(ϕϕ)Δ(ek)=(ϕϕ)(βkkkekek+i=1,iknβikk(eiek+ekei))=βkkkαk2+2i=1,iknβikkαiαk.Hence, ϕ is a character if and only if for all k = 1, …, n, αk=αk(βkkkαk+2i=1,iknβikkαi).

Corollary 4.8

The linear map ϕ=i=1nei is a character of any LV-coalgebra C.

Proof.

Let a = (α1, …, αn)T = 1. By Proposition 4.7, ϕ is a character if and only if 1=βkkk+2i=1,iknβikkfor allk=1,,nwhich follows straightforwardly for any LV-coalgebra C. (See also [Citation8, Theorem 4.3].)

Group-like elements of LV-coalgebras behave, with respect to weight maps (or characters), as do idempotent elements in genetic algebras [Citation7, Lemma 4.1, p.65]. Indeed from the biological viewpoint, the existence of group-like elements can be understood as equilibrium states for the underlying biological system.

Proposition 4.9

Let ϕ be a character of an LV-coalgebra C. For any group-like element xC, either ϕ(x) = 0 or ϕ(x) = 1.

Proof.

Let xC be group-like. Then ϕ(x) = (ϕ ⊗ ϕ)Δ(x) = (ϕ ⊗ ϕ)(x ⊗ x) = ϕ(x)2. Hence either ϕ(x) = 0 or ϕ(x) = 1.

Corollary 4.10

Characters of LV-coalgebras, if there exist, are not necessarily unique.

Proof.

Let C be the LV-coalgebra introduced in Example 3.4. The cubic matrix P associated to C unfolded by its frontal slices is P=(12014015000140001515150001400015014012)and Proposition 4.7(3) gives that ϕ=i=1nαiei is a (nonzero) character of C if and only if a = (α1, α2, α3)T is of the form: a={(α,0,2α),αK;(α,1,2α),αK;(0,5,0).Thus the LV-coalgebra C defined in Example 3.4 has two 1-parameter families of characters together to one additional nontrivial character.

5. Constructions of LV-coalgebras

Achieving the construction of general objects enclosing simultaneously different (genetic) features has been a recurrent topic in the literature of genetic algebraic structures. See, for instance, [Citation7, p.103] or [Citation8, Section 5]. Here we consider different constructions of coalgebras preserving the property of being Lotka–Volterra.

Following [Citation8, Subsection 5.1], given a n-dimensional K-vector space V we denote by C0(V) the set of all coalgebra structures defined on V or, equivalently, the set of all linear maps Δ: V → V ⊗ V. The set C0(V) was shown to be a vector space in [Citation8, Proposition 5.1].

Let pN and B={e1,,en} be an (ordered) basis of V. Given elements Δ1,,ΔpC0(V), with Δt(ek)=i,j=1nβijt,keiejk=1,,n,t=1,,p;and (α1,,αp)TKp, we write Δ=t=1pαtΔt. Then Δ(ek)=i,j=1nβijkeiejk=1,,n,where for all i, j, k = 1, …, n, we have βijk=t=1pαtβijt,k.

Theorem 5.1

Convex linear combinations of real LV-coalgebras are Lotka–Volterra.

Proof.

Let {Ct=(V,Δt)}t=1p be real LV-coalgebras and (α1,,αp)TRp be nonnegative and such that t=1pαt=1. We claim that C=(V,Δ), with Δ=t=1pαtΔt, is also Lotka–Volterra. Indeed LVC-1 and LVC-2 follow straightforwardly from being each Ct LV-coalgebra, and LVC-3 follows from the Cauchy–Schwarz inequality.

Corollary 5.2

Any convex linear combination of real LV-coalgebras with genetic realization has genetic realization.

Proof.

LVC-4 follows from the nonnegativity of (α1,,αp)TRp.

Remark 5.1

The direct sum C=t=1pCt of any finite family of coalgebras {Ct=(Vt,Δt)}t=1p has a coalgebra structure resulting from CC being isomorphic to r,s=1p(CrCs) and, therefore, containing a copy of t=1p(CtCt) [Citation9, p.50].

Let V=t=1pVt, with Vt being a nt-dimensional K-vector space with basis Bt={et,1,et,nt}. If Ct=(Vt,Δt) is a coalgebra, then Δ:V → V ⊗ V, given by Δ(xt) = Δt(xt), for all xt ∈ Vt, defines a coalgebra structure on V. We write B={et,k}1knt1tp

Theorem 5.3

Let {Ct=(Vt,Δt)}t=1p be LV-coalgebras. Then, the direct sum C=(V=t=1pVt,Δ) is an LV-coalgebra. If, moreover, each Ct is a real LV-coalgebra with genetic realization, so is C.

Proof.

Write, for k = 1, …, nt and t = 1, …, p, Δ(et,k)=r,s=11inr1jnspβr,i;s,jt,ker,ies,t.On the other hand, Δ(et,k)=Δt(et,k)=i,j=1ntβijt,ket,iet,j. Thus we have βr,i;s,jt,k={βijt,k,r=s=t;0,otherwise.Assume now Ct=(Vt,Δt) is an LV-coalgebra, for all t = 1, …, p. Then (C,Δ) clearly satisfies LVC-1 and LVC-2, whereas LVC-3 follows from the fact that Δ(xt) = Δt(xt), for all xt ∈ Vt. The last statement about the genetic realization (i.e. C satisfying LVC-4) is clear.

The family of LV-coalgebras is therefore closed under convex linear combinations (when K=R) and direct sums, preserving also both constructions the genetic realization (i.e. LVC-4). LV-coalgebras are not however closed under tensor products or the cocommutative duplication.

Example 5.4

Let (C,Δ) be a coalgebra. It is a well-known result that the (K-vector space) CC admits a comultiplication Δ¯=(idτid)(ΔΔ), where τ denotes the twist map, endowing CC with a coalgebra structure [Citation9, p.49].

Consider now the 2-dimensional coalgebra (C,Δ) with basis B={e1,e2} and comultiplication: Δ(e1)=12e1e1+14(e1e2+e2e1),Δ(e2)=12e2e2+14(e1e2+e2e1).Clearly C is Lotka–Volterra with respect to B, but it is not difficult to check that (CC,Δ¯) is not an LV-coalgebra (with respect to the basis B¯={eij=eiej}i,j=12).

Finally let (CC,) be the cocommutative duplication of the (cocommutative) coalgebra C [Citation8, Definition 3.1], that is, CC=CC/Σ, where Σ={iI(xiyiyixi)xi,yiC,iI,I∣<},is the symmetric tensor product of the (cocommutative) coalgebra C. Recall CC is a cocommutative coalgebra with comultiplication (using Sweedler notation [Citation9, p.32]) (xy)=(x),(y)(x(1)y(1))(x(2)y(2)).Then CC is a 3-dimensional coalgebra with basis B={e1e1,e1e2,e2e2} and, it holds that (e1e1)=14(e1e1)(e1e1)+14((e1e1)(e1e2)+(e1e2)(e1e1))+116((e1e1)(e2e2)+(e2e2)(e1e1))+116((e1e2)(e1e2)+(e1e2)(e1e2)).Hence (CC,) is not Lotka–Volterra (with respect to B).

6. Structure of LV-coalgebras

The different algebraic properties of the generators (i.e. natural basis elements) of algebraic objects with genetic significance usually reproduce their different behaviour within the underlying genetic system, when considered from the dynamical viewpoint. The classification of a natural basis elements into algebraically persistent and transient generators considered in [Citation6, Subsection 3.4.2] led Tian to consider semi-direct sum decompositions of evolution algebras. Here we recover the notion of semi-direct sum decomposition, and apply it to Lotka–Volterra coalgebras.

Let C be an LV-coalgebra with natural basis B={e1,,en}. We write [n] = {1, …, n} and Ik=Kek, for all k ∈ [n].

Proposition 6.1

Let k ∈ [n]. If Ik is a subcoalgebra of C, then ek is a group-like element. Otherwise Ik is a coideal.

Proof.

Take k ∈ [n] and assume Ik is a subcoalgebra of C. This implies Δ(ek)=βkkkekek and, therefore, by LVC-3, βkkk=1. Hence ek is group-like. Assume otherwise, ek is not group-like. Then βikk0 for some i ≠ k, and Δ(ek)=βkkkekek+i=1,ikkβikk(eiek+ekei)=ek(i=1nβikkei)+(i=1,iknβikkei)ek.Therefore Δ(ek)IkC+CIk. Hence Ik is a coideal.

Let us write [n]=ΛΛI (disjoint union), where Λ={k[n]ekisgrouplike} and, ΛI=[n]Λ. Obviously Λ= or ΛI= remains possible. Sufficient conditions for Λ are given in Theorem 4.5.

Lemma 6.2

  • D = k∈Λ Ik is a subcoalgebra of C.

  • I=kΛIIk is a subcoideal of C.

Proof.

It follows from [Citation9, p.18].

Theorem 6.3

Let C be an LV-coalgebra. Then C=DI, where denotes the direct sum of vector subspaces.

Proof.

Clear.

Remark 6.1

Following [Citation6, Theorem 11] we will call such an LV-coalgebra decomposition to be a semidirect sum decomposition of C. In such a decomposition (co)algebraic persistent elements can be identified to group-like elements, being (co)algebraic transient otherwise.

The dynamical behaviour of LV-coalgebras basis elements becomes apparent when considering their accompanying matrices. Write r = |Λ| and t = nr. Then, by Theorem 6.3, the accompanying matrices of P have the following form (see Proposition 2.6):

Corollary 6.4

Reordering the natural basis elements, if necessary, so that, B={e1,,er,er+1,,en}, being ek group-like, for k = 1, …, r: P(i)=P(j)=(Ir×rAr×t0t×rBt×t)

Proof.

It suffices to note that for group-like elements, i.e. for all k = 1, …, r, it holds that βijk=δikδjk, i, j = 1, …, n and, therefore, i=1nβikk=βkkk=1. On the other hand, by Theorem 6.3, for all j = 1, …, t, there exists some i = 1, …, r, such that aij=βi,r+jr+j0.

6.1. Further comments

Under the additional assumption of being endowed with genetic realization (i.e. LVC-4), accompanying matrices of LV-coalgebras evidence the connection between these new family of coalgebras and dynamical systems. Indeed, then P(i) turns out to be not only nonnegative, by LVC-4, but also column stochastic as a result of LVC-3. This settles a new connection between LV-coalgebras and Markov processes, where the topology of P(i) classifying the matrix indices into essential and inessential classes translates into the (co)algebraic persistency and transiency of the natural basis elements. It is a forthcoming work to provide a detailed approach to the ergodic behaviour of LV-coalgebras from the viewpoint of Markov chains.

Acknowledgments

This paper was written while the third author was visiting the Department of Mathematics of the University of Chile, supported by FONDECYT 1170547, and wants to thank the members of the department there for their hospitality. The authors want to thank the referee for his/her comments.

Disclosure statement

No potential conflict of interest was reported by the author(s).

Additional information

Funding

The authors are supported by Fondo Nacional de Desarrollo Científico y Tecnológico (FONDECYT) grant 1170547. The third author is also supported by MTM2017-83506-C2-1-P (AEI/FEDER, UE).

References

  • Itoh Y. Nonassociative algebra and Lotka–Volterra equation with ternary interaction. Nonlinear Anal. 1981;5(1):53–56.
  • Lyubich Y. Mathematical structures in population genetics. Berlin: Springer-Verlag; 1983.
  • Gutierrez JC, Garcia CI. On Lotka–Volterra algebras. J Algebra Appl. 2019;18(10):1950187.
  • Gutierrez JC, Garcia CI. Derivations of Lotka–Volterra algebras. Sao Paulo J Math Sci. 2019;13:292–304.
  • Reed ML. Algebraic structure of genetic inheritance. Bull Amer Math Soc. 1997;34(2):107–131.
  • Tian JP. Evolution algebras and their applications, Berlin: Springer-Verlag; 2008. (Lecture Notes in Mathematics; 1921).
  • Wörz-Busekros AAlgebras in genetics, New York: Springer-Verlag; 1980. (Lecture Notes in Biomathematics; 36).
  • Tian J, Li B-L. Coalgebraic structure of genetic inheritance. Math Biosci Eng. 2004;1(2):243–266.
  • Sweedler ME. Hopf algebras, New York: W. A. Benjamin, Inc.; 1969. (Mathematics Lecture Note Series).
  • Paniello I. On evolution operators of genetic coalgebras. J Math Biology. 2017;74(1-2):149–168.
  • Paniello I. Evolution coalgebras. Linear Multilinear Algebra. 2019;67(8):1539–1553.