Research article

Continuous functions on primal topological spaces induced by group actions

  • Received: 10 October 2024 Revised: 09 November 2024 Accepted: 14 November 2024 Published: 14 January 2025
  • MSC : Primary 54A10, 54C05; Secondary 54D05, 54D30

  • If G is a group acting on a set X, then for any aG, the restriction ϕa:XX of the action to a induces a topology τa for X, called the primal topology induced by ϕa. First, we obtain a characterization of the normal subgroups in terms of the primal topologies. Later, we prove that some commutative relations among elements on the group G determine the continuity of maps among different primal spaces (X,τϕx). In particular, we prove the continuity of some maps when a,b,qG satisfy a quantum type relation, ba=qab, as is in the quaternion and Heisenberg groups.

    Citation: Luis Fernando Mejías, Jorge Vielma, Elvis Aponte, Lourival Rodrigues De Lima. Continuous functions on primal topological spaces induced by group actions[J]. AIMS Mathematics, 2025, 10(1): 793-808. doi: 10.3934/math.2025037

    Related Papers:

    [1] Ahmad Al-Omari, Mesfer H. Alqahtani . Some operators in soft primal spaces. AIMS Mathematics, 2024, 9(5): 10756-10774. doi: 10.3934/math.2024525
    [2] Hanan Al-Saadi, Huda Al-Malki . Generalized primal topological spaces. AIMS Mathematics, 2023, 8(10): 24162-24175. doi: 10.3934/math.20231232
    [3] Ahmad Al-Omari, Ohud Alghamdi . Regularity and normality on primal spaces. AIMS Mathematics, 2024, 9(3): 7662-7672. doi: 10.3934/math.2024372
    [4] Teerapong Suksumran . On the fixed space induced by a group action. AIMS Mathematics, 2022, 7(12): 20615-20626. doi: 10.3934/math.20221130
    [5] Murad ÖZKOÇ, Büşra KÖSTEL . On the topology $ \tau^{\diamond}_R $ of primal topological spaces. AIMS Mathematics, 2024, 9(7): 17171-17183. doi: 10.3934/math.2024834
    [6] Hanan Al-Saadi, Huda Al-Malki . Correction: Generalized primal topological spaces. AIMS Mathematics, 2024, 9(7): 19068-19069. doi: 10.3934/math.2024928
    [7] Ohud Alghamdi . On the compactness via primal topological spaces. AIMS Mathematics, 2024, 9(11): 32124-32137. doi: 10.3934/math.20241542
    [8] Zhanjiang Ji . The research of $({\rm{G}}, {\rm{w}})$-Chaos and G-Lipschitz shadowing property. AIMS Mathematics, 2022, 7(6): 10180-10194. doi: 10.3934/math.2022566
    [9] Abdul Razaq, Ibtisam Masmali, Harish Garg, Umer Shuaib . Picture fuzzy topological spaces and associated continuous functions. AIMS Mathematics, 2022, 7(8): 14840-14861. doi: 10.3934/math.2022814
    [10] Ohud Alghamdi, Ahmad Al-Omari, Mesfer H. Alqahtani . Novel operators in the frame of primal topological spaces. AIMS Mathematics, 2024, 9(9): 25792-25808. doi: 10.3934/math.20241260
  • If G is a group acting on a set X, then for any aG, the restriction ϕa:XX of the action to a induces a topology τa for X, called the primal topology induced by ϕa. First, we obtain a characterization of the normal subgroups in terms of the primal topologies. Later, we prove that some commutative relations among elements on the group G determine the continuity of maps among different primal spaces (X,τϕx). In particular, we prove the continuity of some maps when a,b,qG satisfy a quantum type relation, ba=qab, as is in the quaternion and Heisenberg groups.



    Primal topological spaces were first introduced by Shirazi and Golestani [12] with the name functional Alexandroff spaces, and subsequently by Echi [2] in the following way: Given a set X and a map f:XX, then the collection, τf={UX:f1(U)U} is a topology for X, which is called the primal topology induced by f, thus producing the so-called primal space (X,τf). Many important issues of this space, especially compactness and connectedness, can be described in terms of the dynamics of its points with respect to the function f. Some further developments came later, especially with the work of Echi [2] and Echi and Turki [3]. The descriptions of the topological properties of primal spaces have been used recently in applications to problems in linear algebra and number theory, see Lazaar et al. [6], Lazaar and Sabri [7], Mejías et al. [8], Vielma and Guale [13], and Vielma et al. [14].

    In addition, topologies induced by semigroup actions in a set, which was studied in [4], have been applied both in algebraic and topological contexts, as well as in some areas of computer science. The applications in algebra are established via Green's left quasi-order. In the field of topology, the main idea is to consider the relationship between Green's left quasi-order and principal topologies; see Richmond [9]. Working similarly to those works, and to generalize, we consider the primary topologies induced by group actions and investigate the continuity of maps defined on such primal topological spaces, thereby obtaining some characterizations of homeomorphisms among spaces via properties of the acting groups.

    Thus, we begin with an action Φ of a group G on a set X, that is, a map Φ:G×XX which has some nice properties of compatibility with the group operation. Then, we consider the primal topology induced on X by map the ϕa:XX obtained when Φ is restricted to one specific aG; it turns out that the continuity of maps among different primal spaces are determined by relations among elements of the group. A very special situation is obtained when the group acts on itself with both a left translation and a conjugation. In that case, we obtain a characterization of normal subgroups in terms of the concepts from the topology.

    Our presentation and contribution to the literature begins in Section 2, with an introduction to the basic facts about primal topological spaces and some results and conventions related to the notation. In Section 3, we set the context of the primal topologies induced by group actions with specific characterizations of normal subgroups in terms of the primal topologies. This part is followed by a quite long list of examples in Section 4, which illustrates a broad set of cases that are explained with the results presented in Section 5, where we explore the continuity of maps defined among this type of primal spaces and introduce some properties that are determined by commutation relations among elements of the groups involved. In particular, some conclusions about continuity of maps are derived when we sum the "quantum type" relation as ba=qab, which appears in interesting examples such as the quaternion group and the Heisenberg group.

    Note that the problem of applying the techniques of this paper to topologies induced by actions of some richer algebraic structures as rings, modules, and algebras may be considered.

    In this section, we present the basic notions and standard notation related to primal topological spaces and their most important properties. In particular, we present the characterization of both minimal open and closed sets, as well as those of compact and connected sets.

    The concept of a primal space was established by Shirazi and Golestani [12] in the following terms: If X and f:XX is a map, then the collection

    τf={UX:f1(U)U}

    is a topology for X, which is called the primal topology on X induced by f, and (X,τf) is said to be a primal space. In the contexts where no confusion arises, we simply denote (X,τf) by Xf. The following two facts are straightforward.

    Remark 2.1. The map f:XfXf is continuous.

    Remark 2.2. If the function f:XX is a bijective function, then Xf and Xf1 are homeomorphic.

    Some general properties of primal topological spaces are determined by the dynamics of points concerning the function f. Let N denote the set of positive integers and let N0=N{0}. If rN0, then fr is considered as the r-fold composite ff, where f0 is the identity map, which is defined by Γf(p) the trajectory of pX, and given by the following:

    Γf(p)={yX:y=fr(p) for   rN0}.

    Thus, it turns out that Γf(p)=¯{p} (the closure of {p}). If fr(p)=p for some r>0, then we say that Γf(p) is a periodic trajectory and p itself is referred to as a periodic point. If Γf(p) contains just one element, then p is said to be a fixed point.

    The basic open sets in Xf can also be described using trajectories. In fact, for any pX, it is known that the smallest open set containing p, which we denote as Pf(p), is given by the following:

    Pf(p)={yX:p=fr(y) for  rN0}.

    Remark 2.3. In general, the primal space Xf is compact if and only if there exists a finite set {p1,p2,,pn}X such that for all xX, it turns out that xPf(pi) and some pi{p1,p2,,pn}. On the other hand, the space Xf is connected if and only if for all xX, there exists a pX such that xPf(p)Γf(p).

    Next, we exhibit a few examples that illustrate the geometric aspects of primal spaces. The first two examples are well known, see Echi and Turki [3] and Shirazi and Golestani [12].

    Example 2.4. If X and qX, then consider the map f:XX, such that f(x)=q for all xX (constant). Then, for all xq, we have that {x}τf. Furthermore, the space (X,τf) satisfies the axiom T0, but it is not a T1 space.

    Example 2.5. For any X, the primal topology induced by the identity map id:XX, id(x)=x, is the discrete topology. In this case, each xX is a fixed point. This is the only case in which a primal space is T1, because any set {x} is closed if and only if Γid(x)={x}, which is true if and only if the map that induces the topology is the identity.

    Example 2.6. Let s:ZZ be the map defined by s(x)=1+x. For all xZ, we have Ps(x)Ps(x+1) and the space (Z,τs) is connected, but it is not compact. This example is due to Dahane et al. [1].

    Sometimes, in order to illustrate the properties of a topological primal space, we use a "diagram", where ab means that b is the image of a. Thus, we have the following:

    21012. (Z,τs)

    Example 2.7. Let C:NN be the map defined by the following:

    C(x)={x/2,if  x   is even, 3x+1,if  x   is odd.

    The primal space (N,τC) has a unique periodic trajectory, namely ΓC(1)={1,2,3}. This space is closely related to the so-called "Collatz conjecture". In fact, deciding whether or not (N,τC) is connected is equivalent to solving the famous Collatz problem, see Vielma and Guale [13] and Vielma et al. [14].

    Example 2.8. If A is a square matrix of the order n considered as a linear map A:RnRn, then A induces a primal topology τA on Rn. Some properties of the space (Rn,τA) are deduced from known facts about the matrix A. For instance, the space (Rn,τA) is compact if and only if A is nilpotent. Another interesting result is that A and B are similar; then, the respective primal spaces (Rn,τA) and (Rn,τB) are homeomorphic, see Mejías et al. [8].

    We conclude this section with three results about the relationship between the continuity of functions among primal topological spaces induced on the same set and the trajectories of some elements of its elements. One of the most important points about these results is that it motivates some ideas for more general cases once one considers the set of maps on a set as a semigroup.

    Lemma 2.9. Let X be a set and ϕ,ψ:XX be two maps. Let τϕ and τψ be the primal topologies induced by ϕ and ψ, respectively. If λ:XϕXψ is a function such that λϕ=ψλ, then λ(y)Γψ(λ(x)), for all x,yX with yΓϕ(x).

    Proof. Suppose that xX and yΓϕ(x). Then, there exists a kN0 such that ϕk(x)=y. In this way, we have the following:

    λ(y)=λϕk(x)=ψkλ(x).

    Therefore, λ(y)Γψ(λ(x)).

    The map λ is known as a morphism of flows of Xϕ to Xψ. The next lemma is a particular case of result presented by Haouati and Lazaar [5].

    Lemma 2.10. Let Xϕ,Xψ be primal spaces and λ:XϕXψ be a homeomorphism. Then, λ(Γϕ(x))=Γψ(λ(x)).

    Proof. Note that ¯{x}=Γϕ(x) and ¯{λ(x)}=Γψ(λ(x)). Then, since λ is a homeomorphism, the result is a direct consequence of the fact that for any set AXϕ, we have that λ(¯A)=¯λ(A).

    Corollary 2.11. Let Xϕ,Xψ be primal spaces and let λ:XϕXψ be a homeomorphism. If x,yX, then xΓϕ(y) is equivalent to λ(x)Γψ(λ(y)).

    The concept of a primal topology on a set X is based on a set-theoretical notion associated to a function f:XX, and the complexity of the topology depends on the properties of f.

    Now, we turn our attention to the primal topologies induced by functions obtained by the action of groups as a generalization of the results of Mejías et al. [8] and, somehow motivated by the works of Richmond [9,10], but working with groups rather than semigroups.

    Thus, we consider the action of a group on a set and the primal topology induced by the function obtained when we consider the restriction of the action to a particular element of the group. A very special situation arises when the set is the group itself; in that case, we characterize the normal subgroups in terms of the topology.

    Our research deals with the primal topologies induced on sets by a very specific types of maps: The actions of semigroups and, mainly, groups. In this section, we introduce some basic facts about those spaces. Recall that if X is a set and G is a semigroup, then an action of G on X is a map Φ:G×XX, such that for all a,bG and for all xX, it turns out that

    Φ(a,Φ(b,x))=Φ(ab,x).

    Furthermore, if G is a monoid, then for an action of G on X, it is required that for all xX, it is verified that Φ(ϵ,x)=x. In that case, for each aG, the map ϕa:XX, which is defined by ϕa(x)=Φ(a,x), satisfies the following:

    ϕa(ϕb(x))=ϕab(x)andϕϵ(x)=x.

    Thus, for each aG, we focus on the primal topology τϕa induced by ϕa. As mentioned in Section 2, we use both (X,τϕa) and Xϕa to denote the corresponding primal space. Let us note that if G is a group and Φ is an action, then for all aG, ϕa is invertible and ϕa1=ϕa1.

    It is well known that the arbitrary intersection of topologies is another topology. In the context of semigroup actions, such an intersection will be the trivial topology, if the action is transitive.

    Theorem 3.1. Let X be a nonempty set and F is the collection of all maps ϕ:XX. If G is a semigroup and Φ:G×XX is a transitive action, then

    ϕFτϕ={,X}.

    Proof. Suppose that there exists a set Uτϕ for all ϕF, U, and UX. Given xU, let us take yXU; then, there is aG such that Φ(a,y)=x, so yPϕa(x). Since Pϕa(x) is the smallest open set containing x, then yU, which is a contradiction.

    The following examples show some spaces of particular interest. They play an important role when we consider some concrete cases, especially if G is a group, since we will use the notation given in them.

    Example 3.2. If G is a semigroup and aG, then the left translation by a, which is denoted by La:GG, is given by the following:

    La(x)=axfor all  xG,

    which induces an action Φ:G×GG defined by the following:

    Φ(a,x)=La(x).

    When no confusion arises, we denote the topology by τa and the space GLa by Ga.

    Example 3.3. Let G be a semigroup. For aG, a invertible, we define the conjugation Ka:GG which is given for the following:

    Ka(x)=axa1,xG.

    In this special case, we denote the primal topology induced by Ka on G as κa, that is, κa=τKa. Note that Ka1:GG is the inverse map of Ka.

    With the conjugation K, it is possible to characterize the normal subgroups of a given group.

    Theorem 3.4. Let G be a group and H be a subgroup of G. Then, H is a normal subgroup if and only if H is a closed set in the space (G,κ), where

    κ=aGκa.

    Proof. By definition, HG means that Ka(H)H for all aG. Hence, H is closed in the space (G,κa) for all aG. Therefore, H is closed in (G,κ).

    Now, suppose that H is closed in the space (G,κ); then, H is closed in the space (G,κa) and Ka(H)H for all aG. Therefore, HG.

    In the following, we use the multiplicative notation for the semigroup operation and denote the unit by ϵ. One of the most basic problems in this context is to determine the topological properties of the trajectory of ϵ. We finish this section with three results about this issue.

    Theorem 3.5. If G=(G,,ϵ) is a group and ϕ:GG is a homomorphism, then Pϕ(ϵ) is a subgroup of G, which is a closed set in the space Gϕ.

    Proof. Clearly Pϕ(ϵ), because ϕ(ϵ)=ϵ. If aPϕ(ϵ), then there exists rN0 such that ϕr(a)=ϵ. Thus,

    ϕr(a1)=ϵϕr(a1)=ϕr(a)ϕr(a1)=ϕr(aa1)=ϵ.

    Hence, a1Pϕ(ϵ).

    If a,bPϕ(ϵ), then there exist r,sN0 such that ϕr(a)=ϕs(b)=ϵ. Then,

    ϕr+s(ab)=ϕr+s(a)ϕr+s(b)=ϕs(ϕr(a))ϕr(ϕs(b))=ϕs(ϵ)ϕr(ϵ)=ϵ.

    Thus, abPϕ(ϵ). We conclude that Pϕ(ϵ) is a nontrivial subgroup of G.

    On the other hand, if aPϕ(ϵ) and ϕk(a)=ϵ for a kN0, then ϕk1(ϕ(a))=ϵ; thus ϕ(a)Pϕ(ϵ). Therefore, Pϕ(ϵ)ϕ1(Pϕ(ϵ)) and Pϕ(ϵ) is a closed set.

    One may expect that whenever ϵPϕa(b), we have that b1Pϕa(ϵ), though such a claim has not been proven. However, the following lemma shows a positive result concerning this matter.

    Lemma 3.6. Let G be a group with a,bG, and Φ be an action of G on itself. If ϕa commutes with Lb and ϕa(b)=ϵ, then ϕa(ϵ)=b1.

    Proof. Since ϕa commutes with Lb, it also commutes with Lb1. Therefore,

    ϕa(ϵ)=ϕa(b1b)=ϕa(Lb1(b))=Lb1(ϕa(b))=Lb1(ϵ)=Lb1(ϵ).

    Thus, ϕa(ϵ)=b1.

    In the context of Lemma 3.6, note that we have (ϕar(ϵ))1=ϕa1r(ϵ) for all rN0. In this case, the relationship between Pϕa(ϵ) and Γϕa(ϵ) is even deeper, thus revealing its algebraic structure as the following proposition shows.

    Lemma 3.7. Let G be a group and Φ be an action of G into itself. Let a,bG such that ϕa commutes with Lb and ϕa(b)=ϵ. If Pϕa(ϵ) is a subgroup of G, then Γϕa(ϵ) is periodic.

    Proof. Clearly, (ϕa(ϵ))1=ϕa1(ϵ). However, ϕa1(ϵ)Pϕa(ϵ), which is a subgroup of G. Thus, ϕa(ϵ)Pϕa(ϵ), that is, for some rN0, it turns out that ϕra(ϵ)=ϵ. Therefore, Γϕa(ϵ) is periodic.

    Example 2.6 shows that the hypothesis Pϕa(ϵ) is a subgroup in Lemma 3.7, which cannot be dropped; in that case, a=1 and ϕa is the left translation.

    Next, we introduce some concrete examples of primal topologies induced by the group actions. In each case, we describe some topological properties that can be proved directly; however, they can also be derived as consequences of some of the results presented in Section 5.

    Example 4.1. For nZ, let us consider the additive group (Zn,+,0) acting on itself by the left translation, that is, La(x)=a+x for all xZn. If n is a prime number, then the primal topological space (Zn,τLa) is connected, as illustrated in the following diagrams:

    03124, (Z5,τL3)
    04312. (Z5,τL4)

    If n is not a prime number, then the space (Zn,τLa) may be not connected depending on whether or not a is prime with n and a does not divide n. For example, for n=6 and a=4, we have that 4 does not divide 6 but (Z6,τL4) is not connected. This situation is illustrated as follows:

    021345, (Z6,τL2)
    031425, (Z6,τL3)
    054123. (Z6,τL5)

    Example 4.2. We consider the general linear group GL2(R) acting on R2 as indicated in Example 2.8, that is, AGL2(R) is considered as a linear map A:R2R2. Then, A induces a primal topology τA on Rn. For instance, let us note that if

    A=(1101),andB=(1011),

    then the subgroup H=A of GL2(R) generated by A is not closed in the primal space (GL2(R),κB), because the trajectory of A is not contained in H. In fact,

    KB(A)=BAB1=(0112)H.

    We may also consider the primal topology induced by A as an element of the additive group of square matrices; however, that scenario is weaker than the other because of the commutativity of addition of the matrices.

    Example 4.3. Let S3 be the symmetric group of order 3, that is, the permutation group of the set {1,2,3}. Then,

    S3={1,σ,σ2,ρ,σρ,σ2ρ},

    with the following relations:

    σ3=1,ρ2=1,ρσ=σ2ρ,

    namely permutations σ=(1,2,3) and ρ=(1,2).

    Next, we consider the different primal spaces induced on S3 by the left translation. The following diagrams illustrate that primal spaces (S3,τσ) and (S3,τσ2) are homeomorphic, which is a consequence of the fact that σ1=σ2:

    1σρσρσ2σ2ρ, (S3,τσ)
    1σ2ρσ2ρσσρ. (S3,τσ2)

    On the other hand, it is easy to prove that the spaces (S3,τσ), (S3,τσρ), and (S3,τσ2ρ) are homeomorphic:

    1ρσσ2ρσ2σρ, (S3,τρ)
    1σρ σρσ2σ2ρ, (S3,τσρ)
    1σ2ρσσ2ρσ2ρ. (S3,τσ2ρ)

    It is easy to verify that the left translation Lρ:(S3,τσ)(S3,τσ) is continuous, since τσ={,Γσ(1),Γσ(ρ)},S3}, Lρ(Γσ(1))=Γσ(ρ), and Lρ(Γσ(ρ))=Γσ(1).

    We may also consider S3 acting on itself by the conjugation Kg. Again, since σ1=σ2, we have that the spaces (S3,κσ) and (S3,κσ2) are homeomorphic:

    1σσ2ρσ2ρσρ, (S3,κσ)
    1σσ2ρσρσ2ρ. (S3,τσ2)

    In this case, we have that the spaces (S3,κσ), (S3,κσρ), and (S3,κσ2ρ) are homeomorphic, as illustrated in the following diagram:

    1σσ2ρσρσ2ρ, (S3,κρ)
    1σσ2σρρσ2ρ, (S3,κσρ)
    1σσ2σ2ρ ρσρ. (S3,κσ2ρ)

    Example 4.4. Let S3 be the symmetric group of the set {1,2,3}, acting on R3, which is the same as taking the action on R3 by the subgroup S3(R) of the general linear group GL3(R) generated by the following matrices:

    (010001100)and(010100001).

    Let us note that for all MS3(R), the primal space (R3,τM) is not compact.

    Example 4.5. Let us consider the following quaternion group:

    H={±1,±i,±k,±j},

    with the relations

    i2=j2=k2=ijk=1.

    We may easily prove, that for any ab, a,b±1, the left translation Lb:(H,τa)(H,τa) is continuous. For example, τi={,Γi(1),Γi(j),H}, Lj(Γi(1))=Γi(j), and Lj(Γi(j))=Γi(1).

    1ijki1kj, (H,τi)
    1jkij1ik, (H,τj)
    1kijk1ji. (H,τk)

    These diagrams suggest that all the primal spaces (H,τg) are homeomorphic for g±1. In fact, the diagrams themselves indicate the respective homeomorphisms.

    Again, let us consider the quaternion group H acting on itself, though now with a conjugation. Similar arguments can be introduced to find homeomorphisms among the different spaces (H,κg), for g±1. Let us note that Kg(x)=x, for x=±1,±g, Thus, in this case, the situation looks as follows:

    11ii jjkk, (H,κi)
    11jjkkii, (H,κj)
    11kkiijj. (H,κk)

    Let us note that Kk:(H,κi)(H,κj) and KkKj:(H,κi)(H,κi) are continuous functions and homeomorphisms. Additionally, let us note also that, in general, Ka does not commute with Lb, a,b=i,j,k, ab. For example, LiKj(i)=1, though KjLi(i)=1.

    The next example brings one important mathematical object to the context of a primal topology. The n-th Heisenberg group Hn(R) is considered as the set Hn(R)=Rn×Rn×R with the operation defined for all a,b,a,bRn and c,cR as follows:

    (a,b,c)(a,b,c)=(a+a,b+b,c+c+ab),

    where on the right hand side of the equation represents the standard inner product on Rn. A discrete version of the Heisenberg group may be obtained by considering Z instead of R. Furthermore, in a similar fashion, we can take any commutative ring R instead of R, see Semmes [11]. In particular, it makes sense to consider Hn(Zp) for any pN. In this paper, it is enough to consider n=1 and p=2 for the point that we want to illustrate.

    Example 4.6. Let H=H1(Z2) be the group the Heisenberg associated to Z2={0,1}, that is, the set of triplets (a,b,c) where a,b,cZ2, with the following operation:

    (a,b,c)(a,b,c)=(a+a,b+b,c+c+ab),

    where on the right hand side represents the usual product in Z2. It is easy to verify that (H,) is a noncommutative group with the identity ϵ=(0,0,0) and (a,b,c)1=(a,b,abc).

    Let us note that H is generated by the elements α=(1,0,0), β=(0,1,0), and γ=(0,0,1), with the following relations:

    α2=β2=γ2=ϵ,αγ=γα,βγ=γβ,βα=γαβ.

    Thus, H={ϵ,α,β,γ,αβ,βα,αγ,βγ}. It is clear that H is isomorphic to D8, which is the dihedral group of order 8 (the group of symmetries of the square); however, in the context of this research, we prefer to stick to the the name "Heisenberg group" in order to consider the possibility of more general results.

    The following diagrams illustrate the different primal spaces Ha, aϵ, with H acting on itself by the left translation La(x)=ax:

    ϵα  βαβγαγβαβγ, (Hα)
    ϵβ  αβαγβγαβαγ, (Hβ)
    ϵγ  ααγββγαββα, (Hγ)
    ϵαγαγ  ββγαββγ, (Hαγ)
    ϵβγααββγ  αγβα, (Hβγ)
    ϵαβαβγβαγβαγ. (Hαβ)

    It is easy to verify that the primal spaces Ha with a{ϵ,αβ,βα} are homeomorphic to each other. On the other hand, Hαβ and Hβα are homeomorphic, because (αβ)1=βα. Additionally, let us note that Lβ:HαHα is not continuous, since Lβ(Pα(ϵ))={β,βα} is not connected.

    Now, consider the different primal spaces generated by the conjugation on the Heisenberg group, (H,κa), for a=α,β,αβ. Note that (H,κγ) is homeomorphic to (H,κϵ) because γ commutes with all the elements of H, and (H,καγ) y (H,κβγ) are homeomorphic to (H,κα), because Kαγ=Kα.

    ϵβ  γβγααγαββα, (H,κβ)
    ϵα  γαγββγαββα, (H,κα)
    ϵγαββαααγ  ββγ. (H,καβ)

    Related to the remarks about Example 4.6, we consider the Heisenberg group associated to Z3. It is a non-abelian group of order 27 generated by three elements.

    Example 4.7. The Heisenberg group H1(Z3) is generated by the elements α=(1,0,0), β=(0,1,0), and γ=(0,0,1) under the following relations:

    α3=β3=γ3=ϵ,αγ=γα,βγ=γβ,βα=γαβ.

    The following diagram describes the primal space (H1(Z3),τα):

    ϵαβαβγαγα2α2βα2γ
    αβ2α2β2αβ2γα2β2γαβγα2βγβ2β2γβγ
    αβ2γ2α2β2γ2αγα2γγ2αγ2β2γ2γα2γ2.

    Similarly, we can verify that the primal space (H1(Z3),τβ) contains nine connected components and each of them is a cycle with three elements; therefore, the spaces are homeomorphic.

    Our main purposes in this work is to obtain links between properties of a group and the topological properties of the primal spaces that it induces by actions. Thus, it seems to be natural to look for characterizations of the primal spaces, which are determined by relations among some elements of the group. In that sense, we begin by showing that if a and b belong to a group G such that a is in the centralizer of b, then the spaces (X,τa) and (X,κa) are homeomorphic to (X,τb) and (X,κb), respectively. Before considering that situation, we will establish some general facts.

    Example 5.1. Let us consider the following "rotation matrix":

    Aθ=(cosθsinθsinθcosθ).

    For θ=π/2 and θ=π, the corresponding matrices, which are considered as linear maps, induce the primal topologies τAπ/2 and τAπ on R2, respectively, see Example 2.8. Note that any nonempty set in τAπ/2 other than {(0,0)} has at least four elements. Thus, the function ι:(R2,τAπ/2)(R2,τAπ), which is defined by ι(x)=x for all x, is not continuous. In fact, if e1=(1,0), then e1=(0,1)R2 and ΓAπ(e1)={e1,e1}τAπ/2.

    Lemma 5.2. If G is a semigroup, then a,bG, and Rb:GaGa is defined by the following:

    Rb(x)=xb.

    Then, Rb is continuous.

    Proof. We consider Vτa, that is, La1(V)V. If xLa1(Rb1(V)), then (ax)b=a(xb)V. In other words, xbLa1(V)V. This implies that xRb1(V), so we conclude that La1(Rb1(V))Rb1(V). Therefore, Rb1(V)τa.

    It is important to know what sort of relations among the elements of a group G acting on set X may allow us to decide whether or not two primal spaces are homeomorphic. In that order of ideas, Mejías et al. [8] proved that if A and B are two similar matrices of order n, that is, A=PBP1 for some P, then they induce homeomorphic topologies in Rn. Besides this, as we saw in Example 4.3, the primal spaces (S3,τσ), (S3,τσρ), and (S3,τσ2ρ) are homeomorphic. This conclusion may be obtained from the following relations:

    σρ=(σρ)ρ(σρ)1,andσ2ρ=σρσ1,

    where σ and ρ are the generators of the symmetric group S3. Motivated by these examples, we have derived the following general result.

    Theorem 5.3. Let Φ:G×XX be an action of a group G on a set X. Suppose that there exist a,b,gG such that a=gbg1. Then, the primal spaces Xa and Xb are homeomorphic.

    Proof. We prove that the left translation Lg:XbXa is a homeomorphism. If Vτa, then a1(V)V and

    b1(Lg1(V))=b1(g1(V))=(gb)1(V)=(ag)1(V)=g1(a1(V)).

    Note that g1(a1(V))g1(V)=Lg1(V). Thus, Lg1(V)τb.

    Now, let F be a closed set in Xa, that is, a(F)F. Then, we have the following:

    b(Lg(F))=(bg)(F)=(ga)(F)=g(a(F))g(F)=Lg(F).

    Then, Lg is a closed map, which means that Lg1:XaXb is continuous; therefore, Lg is a homeomorphism.

    By applying the same hypothesis of Theorem 5.3 to the conjugation by g, Kg, we obtain that the topological primal spaces (X,κa) and (X,κb) are homeomorphic.

    Theorem 5.4. Let Φ:G×XX be an action of a group G on a set X. Suppose that there exist a,b,gG such that a=gbg1. Then, the conjugation by g, Kg:(X,κb)(X,κa) is a homeomorphism.

    Proof. Let us prove that the conjugation by g, Kg:(X,κb)(X,κa) is an open map. If Vκb and xKa1(Kg(V))=Ka1(Kg(V)), then g1axa1gV. However, a=gbg1, so bg1xgb1V, which means that g1xgK1b(V)V because Vκb. Hence, xKg(V) and Ka1(Kg(V))Kg(V). In other words, Kg(V)κa. The same kind of argument shows that Kg1:(X,κa)(X,κb) is open and, since Kg is a bijection, we conclude that Kg is a homeomorphism.

    With respect to Example 4.3, we can use Theorem 5.4 to prove that the spaces (S3,κσ), (S3,κσρ), and (S3,κσ2ρ) are homeomorphic.

    Let us note that an argument similar to that in the proofs of Theorems 5.3 and 5.4 gives us a result about the primal spaces Xϕa and Xϕb that involve an arbitrary action.

    Theorem 5.5. Let G be a semigroup and Φ:G×XX be an action. If a,bG with ab=ba, then the maps ϕa:XϕbXϕb and ϕb:XϕaXϕa are both continuous.

    Proof. Of course, it is enough to prove that ϕa:XϕbXϕb is continuous.

    Let Vτϕb and xϕb1(ϕa1(V)); then, ϕa(ϕb(x))V and

    ϕb1(ϕa(ϕb(x)))ϕb1(V).

    However, a and b commute, thus ϕaϕb=ϕbϕa and ϕa(x)ϕb1(V)V. Therefore, xϕa1(V), which means that ϕa1(V)τϕb.

    Corollary 5.6. If G is a group and aG, then the left translation La:GaGa is a homeomorphism.

    Proof. We know that La is continuous by Remark 2.1. Let us note that Theorem 5.5 implies that La1=La1:GaGa is continuous.

    Corollary 5.7. If G is a group and aG, then the conjugation Ka:GaGa is continuous.

    Proof. We know that Ka=LaRa1; therefore, we obtain the continuity of Ka by Theorem 5.5 and Lemma 5.2 applied to b=a1.

    It may seem that the hypothesis of the commutativity of a and b in Theorem 5.5 cannot be dropped. For instance, if we consider A,BGL2(R), as in Example 2.8,

    A=(1101)andB=(1011),

    then ABBA and PB(0,1)={(0,1)}τB, but A1(PB(0,1))={(1,1)}τB because (1,2)PB(1,1).

    However, as the next theorem shows, the hypothesis of commutativity can be replaced by a sort of weaker condition.

    Theorem 5.8. Let Φ:G×XX be an action of a semigroup on a set X. If a,b,qG with ba=qab and τaτq, then ϕb:XaXa is continuous.

    Proof. Suppose that Vτa=τϕa, which means that ϕa1(V)V. Then, if xϕa1(ϕb1(V)), then we have ϕb(ϕa(x))=ϕba(x)=ϕqab(x)V. Thus, ϕab(x)ϕq1(V); from τaτq, it turns out that ϕab(x)V. Then, ϕb(x)ϕa1(V)V, hence xϕb1(V). Thus, ϕa1(ϕb1(V))ϕb1(V), meaning that ϕb1(V)τa, so the map ϕb is continuous.

    Let us note that the hypothesis τaτq in Theorem 5.5 cannot be dropped. In fact, in the case of the Heisenberg group (Example 4.6), we have βα=γαβ, though Lβ:HαHα is not continuous.

    Note that Theorem 5.5 is actually a particular case of Theorem 5.8 which takes q=ϵ. Between these two results, we have the following corollary, which suits very nicely to the investigation of some specific examples.

    Corollary 5.9. Let Φ:G×XX be an action of a semigroup on a set X. If a,bG with akb=ba for a number kN, then ϕb:XaXa is continuous.

    As an application of Corollary 5.9, let us note that for the symmetric group S3 (Example 4.3) from the relation ρσ=σ2ρ, we deduce that the left translation Lρ:(S3,τσ)(S3,τσ) is continuous.

    Similarly, in Example 4.5, the equality a3b=ba whenever ab, a,b±1 and Corollary 5.9 imply that the left translation Lb:(H,τa)(H,τa) is continuous.

    Next, we consider other relevant consequences of Theorem 5.8.

    Corollary 5.10. Let Φ:G×XX and Ψ:G×YY be actions of the semi-group G on some sets X and Y. Suppose that a,b,qG with ba=qab and τϕaτϕq. If f:XϕaYψa is a continuous function, then the composite fϕb:XϕaYψa is continuous.

    Corollary 5.11. Let Φ:G×XX and Ψ:G×YY be actions of the semi-group G on some sets X and Y. Suppose that a,b,qG with ba=qab and τϕaτϕq. If f:XϕaYψb is a continuous function, then the composite fϕb:XϕaYψb is continuous.

    Corollary 5.11 allows us to prove that the functions Kk:(H,κi)(H,κj), Kj:(H,κk)(H,κi), and KkKj:(H,κi)(H,κj) in Example 4.5 are homeomorphisms.

    Lemma 5.12. Let Φ:G×XX be an action of G on a set X. If a,bG with τϕbτϕab1, then τϕbτϕa.

    Proof. Let Vτϕb. Then, Vτϕab1, which means that (ϕaϕb1)1(V)V. Therefore,

    ϕa1(V)ϕb1(V)V,

    because Vτϕb. In other words, Vτϕa.

    Corollary 5.13. Let Φ:G×XX be an action of G on a set X. If a,bG with τϕbτϕab1, then the map ι:XϕaXϕb is continuous.

    Theorem 5.14. Let Φ:G×XX be an action of G on a set X. Let G be a group and G×XX be an action. If a,bG commute and τϕa,τϕbτϕba1, then Xϕa and Xϕb are homeomorphic.

    Proof. Let us consider the composite φ=ϕaι:XϕaXϕb:

    XϕaιXϕbϕaXϕb.

    Obviously, φ is bijective. By Corollary 5.13, we have that both ι and ι1 are continuous. On the other hand, by Theorem 5.5, we have that ϕa and ϕa1 are continuous. Thus, both φ and φ1 are continuous.

    A primal space (X,τ) is a topological structure constructed from a set-theoretical basis, where the topology τ is defined in terms of sets determined by a function f:XX, namely

    τ={UX:f1(U)U}.

    Therefore, it seems to be a natural option to investigate the properties of the topology τ when the function f has some associated structures other than the ones from the set theory. With that idea in mind, we considered functions that were induced by actions of a group G on a set X. It turns out that the algebraic structure has remarkable implications on the properties of τ and the other way around. In particular, we proved the following results:

    1. If H is a subgroup of a group G, then H is normal if and only if H is a closed set in the space (G,κ), where

    κ=aGκa.

    2. If Φ:G×XX is an action of a group G on a set X and if there are a,b,gG such that a=gbg1, then the conjugation by g, Kg:(X,κb)(X,κa) is a homeomorphism.

    3. If Φ:G×XX is an action of a group on a set X, a,b,qG with ba=qab and τaτq, then ϕb:XaXa is continuous.

    From these results, it makes sense to consider similar problems with some more complex algebraic structures such as rings, algebras, etc., and their eventual applications.

    All authors contributed equally to the manuscript. All authors have read and approved the final version of the manuscript for publication.

    The authors declare they have not used Artificial Intelligence (AI) tools in the creation of this article.

    The authors express their gratitude to the referees for their contributions to the final version of this manuscript. This research has been funded by the of Escuela Superior Politécnica del Litoral through project number FCNM-006-2024.

    The authors have no conflict of interest to declare.



    [1] I. Dahane, S. Lazaar, T. Richmond, T. Turki, On resolvable primal spaces, Quaest. Math., 42 (2019), 15–35. https://doi.org/10.2989/16073606.2018.1437093
    [2] O. Echi, The categories of flows of set and top, Topol. Appl., 159 (2012), 2357–2366. https://doi.org/10.1016/j.topol.2011.11.059 doi: 10.1016/j.topol.2011.11.059
    [3] O. Echi, T. Turki, Spectral primal spaces, J. Algebra Appl., 18 (2019), 1–12. https://doi.org/10.1142/S0219498819500300
    [4] D. B. Ellis, R. Ellis, M. Nerurkar, The topological dynamics of semigroup actions, Trans. Amer. Math. Soc., 353 (2001), 1279–1320. https://doi.org/10.1090/S0002-9947-00-02704-5 doi: 10.1090/S0002-9947-00-02704-5
    [5] A. Haouati, S. Lazaar, Primal spaces and quasihomeomorphisms, Appl. Gen. Topol., 16 (2015), 109–118. https://doi.org/10.51768/dbr.v16i1.161201510 doi: 10.51768/dbr.v16i1.161201510
    [6] S. Lazaar, T. Richmond, H. Sabri, The autohomeomorphism group of connected homogeneous functionally Alexandroff spaces, Commun. Algebra, 7 (2019), 3818–3829. https://doi.org/10.1080/00927872.2019.1570240 doi: 10.1080/00927872.2019.1570240
    [7] S. Lazaar, H. Sabri, The topological group of autohomeomorphisms of homogeneous functionally Alexandroff spaces, Topology Proceedings, 55 (2020), 243–264.
    [8] L. Mejías, J. Vielma, A. Guale, E. Pineda, Primal topologies on finite-dimensional vector spaces induced by matrices, Int. J. Math. Math. Sci., 2023 (2023), 1–7. https://doi.org/10.1155/2023/9393234 doi: 10.1155/2023/9393234
    [9] B. Richmond, Principal topologies and transformation semigroups, Topol. Appl., 155 (2008), 1644–1649. https://doi.org/10.1016/j.topol.2008.04.007 doi: 10.1016/j.topol.2008.04.007
    [10] B. Richmond, Semigroups and their topologies arising from Green's left quasiorder, Appl. Gen. Topol., 9 (2008), 143–168. https://doi.org/10.4995/agt.2008.1795 doi: 10.4995/agt.2008.1795
    [11] S. Semmes, An introduction to Heisenberg groups in analysis and geometry, Notices Amer. Math. Soc., 50 (2003), 640–646.
    [12] F. A. Z. Shirazi, N. Golestani, Functional Alexandroff spaces, Hacet. J. Math. Stat., 40 (2011), 515–522.
    [13] J. Vielma, A. Guale, A topological approach to the Ulam-Kakutani-Collatz conjecture, Topol. Appl., 256 (2019), 1–6. https://doi.org/10.1016/j.topol.2019.01.012 doi: 10.1016/j.topol.2019.01.012
    [14] J. Vielma, A. Guale, F. Mejías, Paths in primal spaces and the Collatz conjecture, Quaest. Math., 43 (2020), 1–7. https://doi.org/10.2989/16073606.2020.1806939 doi: 10.2989/16073606.2020.1806939
  • Reader Comments
  • © 2025 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(526) PDF downloads(58) Cited by(0)

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog