Processing math: 96%
Research article

Uniform boundedness of (SL2(C))n and (PSL2(C))n

  • Received: 11 September 2024 Revised: 22 November 2024 Accepted: 22 November 2024 Published: 28 November 2024
  • MSC : 05E16, 20G20, 58D19

  • Let G be a group and S be a subset of G. We say that S normally generates G if G is the normal closure of S in G. In this situation, every element gG can be written as a product of conjugates of elements of S and their inverses. If SG normally generates G, then the length gSN of gG with respect to S is the shortest possible length of a word in ConjG(S±1):={h1sh|hG,sSors1S} expressing g. We write GS=sup{gS|gG} for any normally generating subset S of G. The conjugacy diameter of any group G is Δ(G):=sup{GS|S is a finite normally generating subset of G}. We say that G is uniformly bounded if Δ(G)<. This concept is a strengthening of boundedness. Motivated by previously known results approximating Δ(G) for any algebraic group G, we find the exact values of the conjugacy diameters of the direct product of finitely many copies of SL2(C) and the direct product of finitely many copies of PSL2(C). We also prove that if G1,,Gn be quasisimple groups such that Gi is uniformly bounded for each i{1,,n}, then G1××Gn is uniformly bounded. This is also a generalization of some previously known results in the literature.

    Citation: Fawaz Aseeri. Uniform boundedness of (SL2(C))n and (PSL2(C))n[J]. AIMS Mathematics, 2024, 9(12): 33712-33730. doi: 10.3934/math.20241609

    Related Papers:

    [1] Fawaz Aseeri, Julian Kaspczyk . The conjugacy diameters of non-abelian finite $ p $-groups with cyclic maximal subgroups. AIMS Mathematics, 2024, 9(5): 10734-10755. doi: 10.3934/math.2024524
    [2] Fuya Hu, Chengshi Huang, Zhijie Jiang . Weighted composition operators on $ \alpha $-Bloch-Orlicz spaces over the unit polydisc. AIMS Mathematics, 2025, 10(2): 3672-3690. doi: 10.3934/math.2025170
    [3] Huafeng Liu, Rui Liu . The sum of a hybrid arithmetic function over a sparse sequence. AIMS Mathematics, 2024, 9(2): 4830-4843. doi: 10.3934/math.2024234
    [4] Cheng-shi Huang, Zhi-jie Jiang, Yan-fu Xue . Sum of some product-type operators from mixed-norm spaces to weighted-type spaces on the unit ball. AIMS Mathematics, 2022, 7(10): 18194-18217. doi: 10.3934/math.20221001
    [5] Ermin Wang, Jiajia Xu . Toeplitz operators between large Fock spaces in several complex variables. AIMS Mathematics, 2022, 7(1): 1293-1306. doi: 10.3934/math.2022076
    [6] Mourad Berraho . On a problem concerning the ring of Nash germs and the Borel mapping. AIMS Mathematics, 2020, 5(2): 923-929. doi: 10.3934/math.2020063
    [7] Kai Zhou, Zhenhua Gu, Hongfeng Wu . Orthogonality preserving transformations on complex projective spaces. AIMS Mathematics, 2025, 10(5): 11411-11434. doi: 10.3934/math.2025519
    [8] Jiao Xu, Yinlan Chen . On a class of inverse palindromic eigenvalue problem. AIMS Mathematics, 2021, 6(8): 7971-7983. doi: 10.3934/math.2021463
    [9] Badriya Al-Azri, Ahmad Al-Salman . Weighted $ L^{p} $ norms of Marcinkiewicz functions on product domains along surfaces. AIMS Mathematics, 2024, 9(4): 8386-8405. doi: 10.3934/math.2024408
    [10] Junyong Zhao, Shaofang Hong, Chaoxi Zhu . The number of rational points of certain quartic diagonal hypersurfaces over finite fields. AIMS Mathematics, 2020, 5(3): 2710-2731. doi: 10.3934/math.2020175
  • Let G be a group and S be a subset of G. We say that S normally generates G if G is the normal closure of S in G. In this situation, every element gG can be written as a product of conjugates of elements of S and their inverses. If SG normally generates G, then the length gSN of gG with respect to S is the shortest possible length of a word in ConjG(S±1):={h1sh|hG,sSors1S} expressing g. We write GS=sup{gS|gG} for any normally generating subset S of G. The conjugacy diameter of any group G is Δ(G):=sup{GS|S is a finite normally generating subset of G}. We say that G is uniformly bounded if Δ(G)<. This concept is a strengthening of boundedness. Motivated by previously known results approximating Δ(G) for any algebraic group G, we find the exact values of the conjugacy diameters of the direct product of finitely many copies of SL2(C) and the direct product of finitely many copies of PSL2(C). We also prove that if G1,,Gn be quasisimple groups such that Gi is uniformly bounded for each i{1,,n}, then G1××Gn is uniformly bounded. This is also a generalization of some previously known results in the literature.



    Let G be a group. A norm on G is a function ν:G[0,), which satisfies the following axioms:

    (ⅰ) ν(g)=0 if and only if g=1;

    (ⅱ) ν(g)=ν(g1)gG;

    (ⅲ) ν(gh)ν(g)+ν(h)g,hG.

    We call ν conjugation-invariant if in addition

    (ⅳ) ν(g1hg)=ν(h)g,hG.

    Examples of conjugation-invariant norms are word norms associated with generating sets invariant under conjugations [5], the commutator length [9], verbal norms [10], the Hofer norm on the group of Hamiltonian diffeomorphisms of a symplectic manifold [21], and others [7,18,22].

    The diameter of a group G with respect to a conjugation-invariant norm ν on G, denoted diamν(G) or diam(ν), is sup{ν(g)|gG}. Burago et al. [8] introduced the concept of bounded groups, namely groups for which every conjugation invariant norm has a finite diameter. Examples of bounded groups include finite index subgroups in S-arithmetic Chevalley groups [14], the commutator subgroup of Thompson's group [13], the automorphism groups of regular trees [15], and others [12,18,25]. On the other hand, if G is not bounded, then G is said to be unbounded. Groups with infinite abelianization are examples of unbounded groups [8]. There are also other examples of unbounded groups, and for these examples, we refer the reader to [3,4,6].

    A normally generating subset of a group is one of the most important sources of conjugation invariant norms. Assume that G is a group and S be a subset of G. The normal closure of S in G, denoted by S, is the smallest normal subgroup of G containing S. In other words, it is the subgroup of G that is generated by all conjugates of elements of S. We say that G is normally generated by S if G=S. In this situation, every element of G can be written as a product of elements of

    ConjG(S±1):={h1sh|hG,sSors1S}. (1.1)

    If SG normally generates G, then the length gS[0,) of gG with respect to S is defined to be

    gS=inf{nN|g=s1snfor somes1,,snConjG(S±1)}.

    In other words, it is the shortest possible length of a word in ConjG(S±1) expressing g.

    The word norm is

    .S:G[0,),
    ggS.

    It is easy to see that the word norm .S is a conjugation-invariant norm on G. The diameter of a group G with respect to the conjugation-invariant word norm .S is given by

    GS:=sup{gS|gG}.

    Let G be a group normally generated by a finite set S. It has been shown in [18, Corollary 2.5] that G is bounded if and only if GS<. This highlights the significance of using word norms for finite normally generating subsets and their diameters to study the boundedness of groups.

    In light of [18, Corollary 2.5], there are some refinements of the notion of boundedness. We will introduce some notation to describe these refinements. For any group G and any n1, let

    Γn(G):={SG||S|nandSnormally generatesG},
    Γ(G):={SG||S|<andSnormally generatesG}.

    Set

    Δn(G):=sup{GS|SΓn(G)},Δ(G):=sup{GS|SΓ(G)}.

    The conjugacy diameter of any group G is denoted as Δ(G). The group G is said to be strongly bounded if Δn(G)< for all nN (see [18, Definition 1.1]). The group G is said to be uniformly bounded if Δ(G)< (see [18, Definition 1.1]).

    Recently, conjugacy diameters were studied for several classes of finite and infinite groups. For the case of finite groups, Kaspczyk and the author of this paper studied the conjugacy diameters of non-abelian finite p-groups with cyclic maximal subgroups in [1]. They determined the conjugacy diameters of the semidihedral 2-groups (see [1, Theorem 1.4]), the generalized quaternion groups (see [1, Theorem 1.5]), and the modular p-groups (see [1, Theorem 1.6]). Kȩdra et al. also proved that Δ(PSL(n,q))12(n1) for any n3 and any prime power q (see [18, Example 7.2]). Also, Libman and Tarry showed that if G is a non-abelian group of order pq, where p and q are prime numbers with p<q, then Δ(G)=max{p12,2} (see [20, Thereom 1.1]), and that Δ(Sn)=n1 for any n2 (see [20, Thereom 1.2]). For the case of infinite groups, we have Δ(D)4 (see [18, Example 2.8]). Also, the conjugacy diameters of SL2(C) and PSL2(C) have been studied, and we have:

    Theorem 1.1. ([17, Theorem 3.3]) Let G:=SL2(C). Then G is normally generated by any gGZ(G), and moreover

    Gg={2,iftrace(g)=0,3,otherwise.

    Hence, G is uniformly bounded and Δ(G)=3.

    Theorem 1.2. ([17, Theorem 3.4]) Let G:=PSL2(C). Then G is normally generated by any gGZ(G), and moreover, Gg=2. Hence G is uniformly bounded and Δ(G)=2.

    The next theorem provides a family of uniformly bounded groups.

    Theorem 1.3. ([18, Theorem 4.3]) Let G be a finitely normally generated algebraic group over an algebraically closed field. Let G0 denote the connected component of the identity. Then G is uniformly bounded and Δ(G)4dim(G)+Δ(G/G0).

    Note that by [18, Proposition 4.2], a linear algebraic group over an algebraically closed field is finitely normally generated if and only if it has a finite abelianization.

    In general, calculating Δ(G) for a group G is not easy, and the goal of this paper is to compute this value for some finitely normally generated algebraic groups over C. We show that the bound for Δ(G) in Theorem 1.3 is not sharp, and we obtain the exact value for the conjugacy diameters of the direct product of finitely many copies of SL2(C) and the direct product of finitely many copies of PSL2(C). In particular, we prove the following:

    Theorem 1.4. Let n1 be a natural number and let G:=PSL2(C). Then,

    Δ(Gn)=Δn(Gn)=2n.

    Theorem 1.5. Let n1 be a natural number and let G:=SL2(C). Then,

    Δ(Gn)=Δn(Gn)=3n.

    The bound in Theorem 1.3 is far from sharp. As an example, let G:=SLm(C) where m1. We have that G is a linear algebraic group and dim(G)=m21 (see [16, Section 7.2]). By [16, Proposition 3.1]), we have dim(Gn)=n(m21) for every n1. Now, Theorem 1.5 shows that Δ((SL2(C))n)=3n, whereas Theorem 1.3 yields that Δ((SL2(C))n)4n(221). Also, there is some improvement on the bound in Theorem 1.3 for some special cases (see [19, Theorem 1.3]). However, this improvement does not give exact values of Δ((SL2(C))n) and Δ((PSL2(C))n).

    In order to prove Theorems 1.4 and 1.5, we need first to give an alternative method to prove Theorems 1.1 and 1.2. The new method we used there is based on the so-called rational canonical form and a new definition (see Definition 2.2). The importance of this definition will be mentioned in Remark 5.1.

    The other main result of this paper is to generalise the following result:

    Theorem 1.6. [17, Lemma 2.20 (c)] Let n1 be a natural number, and let G1,,Gn be finitely normally generated groups. Then G=G1××Gn is finitely normally generated, and if the groups Gi are uniformly bounded and simple then G is uniformly bounded.

    Our main result generalises the above result as follows:

    Theorem 1.7. Let n1 be a natural number, and let G1,,Gn be quasisimple groups. Suppose that Gi is uniformly bounded for each i{1,,n}. Then G=G1××Gn is uniformly bounded.

    In this section, we present some results and notation needed for the proofs of Theorems 1.4, 1.5, and 1.7.

    Definition 2.1. ([18, Section 2]) Let X be a subset of a group G. For any n0, we define BX(n) to be the set of all elements of G that can be expressed as a product of at most n conjugates of elements of X and their inverses.

    By Definition 2.1, we have

    {1}=BX(0)BX(1)BX(2)

    The next lemma is [18, Lemma 2.3]. We prove the last two parts of this lemma since the proofs are not given in [18].

    Lemma 2.1. Let G be a group; let X,YG and n,mN. Then,

    (i) BX(n)1=BX(n) and BX(n) is invariant under conjugation in G.

    (ii) If XY then BX(n)BY(n).

    (iii) BX(n)BX(m)=BX(n+m).

    (iv) YBX(n)BY(m)BX(mn).

    (v) If π:GH is a surjective group homomorphism, then BHπ(X)(n)=π(BGX(n)) for any XG.

    (vi) If π:GH is a surjective group homomorphism, then BGπ1(Y)(n)=π1(BHY(n)) for any YH.

    Proof. (ⅰ)–(ⅳ) follow from the definition.

    We now prove (ⅴ). Let π(g)π(BGX(n)). Then gBGX(n). So there exist x1,,xmX for some 0mn such that

    g=mi=1h1ixϵiihi,

    where hiG and ϵi{1,1} for each 1im. We have

    π(g)=mi=1π(hi)1π(xi)ϵiπ(hi).

    This shows that π(g)BHπ(X)(n). Therefore, π(BGX(n))BHπ(X)(n).

    Now let hBHπ(X)(n). So there exist π(x1),,π(xm)π(X) for some 0mn such that

    h=mi=1h1iπ(xi)ϵihi,

    where hiH and ϵi{1,1} for each 1im. For each 1im, let ˜hiG with π(˜hi)=hi. Then

    h=mi=1π(˜hi)1π(xi)ϵiπ(˜hi)=mi=1π(˜h1ixϵii˜hi).

    This shows that hπ(BGX(n)). Therefore, BHπ(X)(n)π(BGX(n)). The proof is complete.

    Finally, the proof of (ⅵ). Let gBGπ1(Y)(n). Then there are elements x1,,xm of π1(Y) for some 0mn such that

    g=mi=1h1ixϵiihi,

    where hiG and ϵi{1,1} for each 1im. We have

    π(g)=mi=1π(hi)1π(xi)ϵiπ(hi).

    Since π(xi)Y for each 1im, it follows that π(g)BHY(n). Therefore, gπ1(BHY(n)), and it follows that BGπ1(Y)(n)π1(BHY(n)).

    Suppose now that gπ1(BHY(n)). Then π(g)BHY(n). So there exist y1,,ymY for some 0mn such that

    π(g)=mi=1h1iyϵiihi,

    where hiH and ϵi{1,1} for each 1im. For each 1im, let ˜hi,˜yiG with π(˜hi)=hi,π(˜yi)=yi. Then,

    π(g)=mi=1π(˜hi)1π(˜yi)ϵiπ(˜hi)=mi=1π(˜h1i˜yϵii˜hi)=π(mi=1˜h1i˜yϵii˜hi).

    Setting w:=mi=1˜h1i˜yϵii˜hi, it follows that g=uw for some uker(π). We have

    g=uw=u(mi=1˜h1i˜yϵii˜hi)=˜h11˜h1u(mi=1˜h1i˜yϵii˜hi)=˜h11˜h1u˜h11˜yϵ11˜h1mi=2˜h1i˜yϵii˜hi.

    With w0:=˜h1u˜h11˜yϵ11, we have

    π(w0)=π(˜h1u˜h11˜yϵ11)=π(˜h1)π(u)π(˜h1)1π(˜y1)ϵ1=π(˜h1)π(˜h1)1π(˜y1)ϵ1=π(˜y1)ϵ1.

    It follows that wϵ10π1(Y). So we have w0BGπ1(Y)(1). Now g=˜h11w0˜h1mi=2˜h1i˜yϵii˜hiBGπ1(Y)(n). Therefeore, π1(BHY(n))BGπ1(Y)(n).

    Next, we will introduce Definition 2.2 to prove Theorems 1.1 and 1.2. The significance of this definition will be explained later (see Remark 5.1).

    Definition 2.2. Let X be a subset of a group G. For any n0, define WX(n) to be the set of all elements of G that can be written as a product of exactly n conjugates of elements of X and their inverses.

    The next result follows easily from the above definitions.

    Lemma 2.2. Let G be a group, let XG, and let n,mN. Then,

    (i) WX(n)1=WX(n) and WX(n) is invariant under conjugation in G.

    (ii) WX(n)WX(m)=WX(n+m).

    (iii) WX(n)BX(n).

    Remark 2.1. Let G be a group, XG, and nN. In general, we do not have BX(n)WX(n). For example, let G:=(Z,+), and let X be the set of odd integers. It is not hard to observe that BX(2)=Z, while WX(2)=2Z. Also, BX(1)=X{0}, while WX(1)=X.

    Lemma 2.3. Let G be a group such that [G,G]=G and G/Z(G) is simple. Then every gGZ(G) normally generates G.

    Proof. Let gGZ(G). Set S:={g} and N:=S. By the definition, N is the smallest normal subgroup of G containing S. We have to show that G=N. Since gN and gZ(G), we have NZ(G). As NG, Lemma [24, Proposition 6.2(ⅲ)] implies that G=N.

    Next, let F be an arbitrary field. We will discuss the rational canonical form over F. Most of the material presented here comes from [11]. Let A be an n1×n2 matrix and B be an m1×m2 matrix. Then their direct sum, denoted by AB, is the (n1+m1)×(n2+m2) matrix of the form

    (A00B).

    Let A be a square matrix over F. The characteristic polynomial of A, denoted by CPA(x), is the polynomial defined by CPA(x)=det(xIA), where x is a variable. A monic polynomial over F is a single variable polynomial with coefficients in F, whose highest order coefficient is equal to one. Therefore, a monic polynomial has the form f(x)=xn+an1xn1++a1x+a0F[x]. The minimal polynomial of A, denoted by mA(x), is the unique monic polynomial f(x) of smallest degree such that f(A)=0. The companion matrix of a monic polynomial f(x)=xn+an1xn1++a1x+a0 over F is the matrix

    (0000a01000a10100a20000an20001an1).

    We denote the companion matrix of f(x) by Cf(x).

    A square matrix R over F is said to be a rational canonical form if R is a direct sum of companion matrices,

    R=Cf1(x)Cf2(x)Cfm(x)=(Cf1(x)000Cf2(x)0000Cfm(x)),

    for monic polynomials f1(x),...,fm(x) of degree at least one such that

    f1(x)|f2(x)||fm(x).

    The next result is [11, Section 12.2, Theorem 16].

    Theorem 2.1. Let AGLn(F).

    (i) The matrix A is GLn(F)-conjugate to a matrix R in rational canonical form.

    (ii) The rational canonical form R for A is unique.

    Let A be a square matrix over F. Then the matrix R from Theorem 2.1 is said to be the rational canonical form of A. If A is conjugate to a rational canonical form

    R=Cf1(x)Cf2(x)Cfm(x),

    where f1(x)|f2(x)||fm(x), then we say that the invariant factors of A are f1(x),f2(x),,fm(x). We call fm(x) the largest invariant factor of A.

    The next result gives a link between the characteristic polynomial of a matrix and its invariant factors. This result is quite helpful for determining the invariant factors, especially for matrices of small size.

    Proposition 2.1. Let A be a square matrix over F.

    (i) The product of all the invariant factors of A is the characteristic polynomial of A.

    (ii) The minimal polynomial mA(x) divides the characteristic polynomial CPA(x).

    (iii) The minimal polynomial mA(x) is the largest invariant factor of A.

    Proof. See [11, Section 12.2, Proposition 20] for (ⅰ) and (ⅱ). (ⅲ) follows from [11, Section 12.2, Proposition 13].

    The finding of the characteristic and minimal polynomials determines all the invariant factors for 2×2 and 3×3 matrices. However, for n×n matrices with n4, the determination of the characteristic and minimal polynomials is generally insufficient for determining all of the invariant factors; see [11, Section 12.2, Examples (3) and (4)]. However, there are algorithms to obtain the invariant factors and convert a given n×n matrix to rational canonical form; see [11, p. 480] or [2]. We will be interested in 2×2 matrices, and so our main tool in determining invariant factors and rational canonical forms will be Proposition 2.1.

    Let us now recall the conjugacy classes of SL2(C). For all tC,sC, define

    D(s):=(s00s1),U(t):=(1t01),andI:=(1001).

    Any element of SL2(C) is conjugate to one of the matrices defined above; see [23, Section 1.2, Example 9]. More precisely, the following holds for gSL2(C).

    (ⅰ) If trace(g)=±2, then g is SL2(C)-conjugate to ±I or ±U(1).

    (ⅱ) If trace(g)±2, then g is SL2(C)-conjugate to D(t) for some tC with trace(g)=t+t1, and t is unique up to replacing t with t1.

    For {2,2}, there are only two SL2(C)-conjugacy classes of matrices of trace . In fact, ±U(1) and ±U(1) are conjugate to each other by (i00i).

    Remark 2.2. Observe that trace(g)=trace(g1) for all gSL2(C). Let g,hSL2(C){±I} and trace(g)=trace(h). Then, g is SL2(C)-conjugate to h. Therefore, gg1 in SL2(C).

    Proposition 2.2. Every non-scalar element gSL2(C) is GL2(C)-conjugate to the companion matrix h:=(011x), where x=trace(g). Moreover, C(x):=(011x) is SL2(C)-conjugate to C(x):=(011x).

    Proof. Let g be a non-scalar element of SL2(C). Then mg(x) has degree 2. Using Proposition 2.1, we deduce that mg(x) is the only invariant factor of g and that mg(x) is equal to CPg(x). So the rational canonical form of g is equal to the companion matrix of mg(x)=CPg(x). By a well-known formula for the characteristic polynomial of a 2×2 matrix, we have CPg(x)=x2trace(g)x+1. Its companion matrix and hence the rational canonical form of g are given by

    h=(011trace(g)).

    The first statement of the proposition now follows from Theorem 2.1.

    One can show that C(x) is SL2(C)-conjugate to C(x) by (xiii0).

    This section is based on [17, Section 3]. We will give a different approach based on the rational canonical form and Definition 2.2 to prove Theorems 1.1 and 1.2. Some results presented here will be needed in Section 5. We need first to introduce some notation and results:

    Let LSL2(C), and let us denote the subset {trace(A)|AL} of C by trace(L). Recall the notations C(x),C(x),D(s),U(t), and I from Section 2. We can prove the following results.

    Lemma 3.1. Suppose that LSL2(C) is conjugation invariant and L=L1. If trace(L)=C and ±C(2)L, then LSL2(C){±I}.

    Proof. Let ±IgSL2(C). We want to show that gL. It follows from the conjugacy classes of SL2(C) that we have two cases for g (see the previous section). The first case, if trace(g)=±2, then, g is SL2(C)-conjugate to ±C(2) (see Proposition 2.2). By hypothesis, we have that ±C(2)L and L is conjugation invariant. So gL. The second case, if trace(g)±2, then g is SL2(C)-conjugate to C(t+t1) for some tC and t is unique up to replacing t with t1 (see Proposition 2.2). Since trace(L)=C, there exists ˜gL such that trace(˜g)=trace(g). So, by Remark 2.2, we have ˜g is SL2(C)-conjugate to g. Since L is conjugation invariant, we have that gL. This completes the proof.

    Lemma 3.2. ([17, Lemma 3.6]) Let gSL2(C){±1} and let n1. Then IBg(n+1)IBg(n) or there exists hBg(n) such that h±I and trace(h)=trace(g).

    Corollary 3.1. ([17, Corollary 3.7]) Let gSL2(C){±1} Then IBg(2)trace(g)=0.

    Lemma 3.3. Let gSL2(C){±1} and let n1. Then IWg(n+1)IWg(n) or there exists hWg(n) such that h±I and trace(h)=trace(g).

    Proof. We follow arguments found in the proof of Lemma 3.2. Let ±IgSL2(C). If IWg(n+1)Wg(n), then I=hm for some hWg(n) and some ±ImWg(1). This implies that h=m1±I. Thus, trace(h)=trace(m1)=trace(m) (see Remark 2.2). Since mWg(1), we have trace(m)=trace(k1g±1k) for some kSL2(C). Hence, trace(h)=trace(k1g±1k)=trace(g). Conversely, if hWg(n) such that h±I and trace(h)=trace(g), then h is conjugate to g±1. But Wg(n)=Wg(n)1. Hence, g1Wg(n) and I=g1gWg(n)Wg(1)=Wg(n+1).

    Corollary 3.2. Let gSL2(C){±I}. Then IWg(2)trace(g)=0.

    Proof. Let ±IgSL2(C). If IWg(2), then there exist m,kWg(1) such that mk=IWg(2). This implies that k=m1 and that trace(k)=trace(m1)=trace(m). In fact, we have m=h1g±1h and k=h1g±1h for some h and h in SL2(C). Since the trace is invariant under conjugation and trace(g)=trace(g1) in SL2(C), we have trace(m)=trace(g)=trace(k). But k=m1. So trace(g)=trace(g). It follows that trace(g)=0. Conversely, suppose that trace(g)=0. We want to show that IWg(2). Since trace(g)=0, we have g is SL2(C)-conjugate to C(i+i1) (see Proposition 2.2). So, I=C(i+i1)C(i+i1)Wg(2).

    Lemma 3.4. Suppose that gSL2(C) is conjugate either to ±C(2) or to C(t+t1) for some t0,±1. Then Wg(2)SL2(C){I}.

    Proof. Let a,b,vC. Assume that X=(v+v1a1a0)Wg(1) and Y=(v+v1b1b0)Wg(1). Let A=(a1a2a3a4)SL2(C). We want to show that

    (k110)=A1XAYWg(1)Wg(1)=Wg(2), where kC. (3.1)

    In particular, we want to show that trace(Wg(2))=C.

    Apply (3.1) with

    a1=a2=(k1)(k2),a3=1(k1),a4=1,a=v(k2)(k1)2,andb=v,

    where k{1,2} to see that

    trace(Wg(2))=C{1,2}.

    But

    I=XX1Wg(2)

    and

    (1110)=(1110)1(v+v1v1v0)(1110)(v+v1v1v0)Wg(2).

    Thus trace(Wg(2))=C.

    We now want to show that ±C(2)Wg(2). Clearly, we have C(2)Wg(2) since 2trace(Wg(2)) and (k110)I. It remains to show that C(2)Wg(2). Assume first that g=C(t+t1). Apply (3.1) with

    a1=1(t21),a2=a3=1,a4=0,v=t,a=(t21)t,andb=t1,

    where t0,±1 obtains C(2)Wg(2). Next, assume that g=±C(2). Then

    C(2)(v+v1v2bbv20)(v+v1b1b0)Wg(2)(v=±1).

    We showed that trace(Wg(2))=C and ±C(2)Wg(2). Lemma 3.1 implies that Wg(2)SL2(C){I}.

    Corollary 3.3. Suppose that gSL2(C) is conjugate either to ±C(2) or to C(t+t1) for some t0,±1. Then Bg(2)SL2(C){I}.

    Proof. It follows from Lemma 3.4 that Wg(2)SL2(C){I}. Thus Bg(2)SL2(C){I} by Lemma 2.2(ⅲ).

    Corollary 3.4. Suppose that gSL2(C) is conjugate either to ±C(2) or to C(t+t1) for some t0,±1. Then Wg(2+n)=SL2(C) for all n1.

    Proof. Let ±IgSL2(C). If trace(g)2 or trace(g)2 then up to conjugacy and taking inverses we assume that either g=±C(2) or g=C(t+t1) for some t0,±1.

    First, we show that Wg(3)=SL2(C). It follows from Lemma 3.4 and Corollary 3.2 that Wg(2)=SL2(C){I}. In other words, let IhSL2(C). Then there exist g1,g2Wg(1) such that

    h=g1g2Wg(1)Wg(1)=Wg(2).

    Using Lemma 3.4 and Corollary 3.2 again, we have

    g1=˜g1˜˜g2Wg(1)Wg(1)=Wg(2).

    So,

    h=˜g1˜˜g1g2Wg(3).

    It remains to show that IWg(3). This also follows from Lemma 3.4 and Corollary 3.2 since if Wg(2)=SL2(C){I} then g1Wg(2) and therefore gg1=IWg(3). So we have Wg(3)=SL2(C).

    We show next that Wg(4)=SL2(C). Since Wg(3)=SL2(C), there are g1,g2,g3Wg(1) such that

    ˜h=g1g2g3Wg(3),

    for all ˜hG. But

    g1=˜g1˜˜g2Wg(1)Wg(1).

    So,

    ˜h=˜g1˜˜g2g2g3Wg(4).

    This shows that Wg(4)=SL2(C). Similarly, one can show that Wg(n+1)=SL2(C) for all n4. This completes the proof.

    Proof of Theorem 1.1. Let gSL2(C)Z(SL2(C)). It follows from Lemma 2.3 that g normally generates SL2(C). We have Bg(1)Bh(1) for any hSL2(C) with trace(g)trace(h). Then Bg(1)SL2(C). Now g±1 is conjugate either to ±C(2) or to C(t+t1) for some tC with t±1. It follows from Corollaries 3.1 and 3.3 that Bg(2)=SL2(C){I}. Then g1Bg(2) and gg1=IBg(3)=SL2(C). Therefore, Δ1(SL2(C))=3. Moreover, Δ(SL2(C))=3 (see [17, Lemma 2.11]).

    Let G=SL2(C) and H=PSL2(C). Let π:GH be the natural projection. If gG, then it follows from Lemma 2.1(ⅴ) that Bπ(g)(n)=π(Bg(n)) for any nN. Since G is finitely normally generated, we have that H is finitely normally generated. Using similar arguments as in the proof of Lemma 2.1(ⅵ), one can show that π1(Bπ(g)(n))=Bg(n)Bg(n) for all nN.

    Proof of Theorem 1.2. Set h:=π(g)PSL2(C), where ±IgSL2(C). Since PSL2(C) has infinitely many conjugacy classes, we have that Bh(1)PSL2(C). It follows from Corollary 3.3 that Bg(2)SL2(C){I}. As Bh(2)=π(Bg(2)), we have π(SL2(C){I})=Bh(2). Thus Δ1(PSL2(C))=2. Moreover, Δ(PSL2(C))=2 (see [17, Lemma 2.11]).

    The key to the proof of Theorem 1.7 is the following lemma.

    Lemma 4.1. Let n1 be a natural number and let G1,,Gn be groups. Let G=G1××Gn, and let S be a normally generating subset of G. Let 1in and

    πi:GGi,(g1,,gn)gi.

    Then {πi(s)|sS} normally generates Gi.

    Proof. Let gGi and let us identify g with (1G1,,1Gi1,g,1Gi+1,,1Gn). Then g can be written as a product of conjugates of elements of S and their inverses. This implies that g (considered as an element of Gi) can be written as a product of conjugates of elements of {πi(s)|sS}{πi(s)1|sS}. So Gi is normally generated by {πi(s)|sS}.

    Proof of Theorem 1.7. We follow arguments found in the proof of Theorem 1.6. For each 1in, let

    πi:GGi,(g1,,gn)gi,

    be the projection of G onto Gi. Let S be an element of Γ(G). Let 1in. Since S normally generates G, we have that {πi(s)|sS} normally generates Gi by Lemma 4.1. As Z(Gi)Gi, it follows that there is some siS with πi(si)Z(Gi). We claim that there exists some yiGi with [yi,πi(si)]Z(Gi). Otherwise, we would have that [yi,πi(si)]Z(Gi) for all yiGi. But this would imply that πi(si)Z(Gi) is central in Gi/Z(Gi), which is not possible since Gi/Z(Gi) is nonabelian simple. Now take some yiGi with [yi,πi(si)]Z(Gi) and define xi:=[yi,πi(si)]. By Lemma 2.3, xi normally generates Gi. Considering xi and yi as elements of G, we have xi=[yi,si]BS(2). So we have xiS2. As xi normally generates Gi for all 1in, we have that {x1,,xn} normally generates G. Since xiS2 for all 1in, it follows that

    GS2G{x1,,xn}.

    It is easy to see that

    G{x1,,xn}ni=1Gixini=1Δ(Gi).

    It follows that

    GS2ni=1Δ(Gi).

    Since S was assumed to be an arbitrary element of Γ(G), it follows that G is uniformly bounded.

    The keys to the proofs of Theorems 1.4 and 1.5 are Definition 2.2 and the following proposition.

    Proposition 5.1. Let n1 be a natural number. Let G1,,Gn, be groups and let Ni be a proper normal subgroup of Gi for each i{1,,n}. Suppose that, for 1in,Gi is normally generated by each giGiNi. Suppose, moreover, that Gi/Ni,1in, is not cyclic. Let G=G1××Gn. Then Δ(G)=Δn(G).

    Proof. Since Γn(G)Γ(G), we have Δn(G)Δ(G). So it suffices to show that Δ(G)Δn(G). Let SΓ(G). Let 1in. Since S normally generates G and since Ni is a proper normal subgroup of Gi, there is some xi=(gi,1,,gi,n)S, such that gi,iNi. Define T:={x±11,,x±1n}. We need to show that T normally generates G. By hypothesis, Gi/Ni is not cyclic. In particular, GiNi,gi,i. Take some hiGiNi,gi,i. Since gi,iNi, we have by hypothesis that Gi is normally generated by gi,i. So hi can be written as a product of conjugates of elements of {gi,i,g1i,i}. Hence, there exist a natural number k1, elements a1,,akGi, and ϵ1,,ϵk{1,1}, such that

    hi=kj=1a1jgϵji,iaj.

    Set ConjG(T):={g1tg|gG,tT}. For each 1jk, let

    ¯aj:=(1G1,,1Gi1,aj,1Gi+1,,1Gn).

    Also, set

    li=kj=1ϵj.

    Then we have

    ((gi,1)li,,(gi,i1)li,hi,(gi,i+1)li,,(gi,n)li)=kj=1¯aj1xϵji¯ajConjG(T). (5.1)

    We also have

    xlii=((gi,1)li,,(gi,n)li)ConjG(T). (5.2)

    Multiplying (5.1) with (5.2), we obtain

    (1G1,,1Gi1,hi(gi,i)li,1Gi+1,,1Gn)ConjG(T). (5.3)

    Since hiNi,gi,i, we have ˜hi:=hi(gi,i)liNi,gi,i and hence ˜hiNi. By hypothesis, Gi is normally generated by ˜hi. Hence, each element of Gi can be written as a product of conjugates of {˜hi,˜h1i}. Now, since ConjG(T)G, (5.3) implies that

    (1G1,,1Gi1,ui,1Gi+1,,1Gn)ConjG(T), (5.4)

    for any uiGi. It follows that ConjG(T)=G. In other words, G is normally generated by T and hence also normally generated by T0:={x1,,xn}. Since T0S, we have GSGT0. Thus,

    Δ(G)=sup{GS| SΓ(G)}sup{GT0| T0Γn(G)}=Δn(G).

    Since SΓ(G) was arbitrarily chosen, it follows that Δ(G)Δn(G).

    Proof of Theorem 1.4. We have Δ(Gn)=Δn(Gn) by Proposition 5.1 and Δ(Gn)2n by [18, Lemma 2.11(b)] and Theorem 1.2. Therefore, it suffices to show that Δ(Gn)2n.

    Let XΓ(Gn). We are going to show that Gn=BX(2n). This clearly implies that GnX2n, and as X was arbitrarily chosen, it follows that Δ(Gn)2n.

    Define surjective group homomorphisms

    πi:GnG,(g1,,gn)gi,

    for all 1in. Since X normally generates Gn, there exists some xi=(gi,1,,gi,n)X with πi(xi)I, for each 1in. In other words, we have that gi,iI for all 1in.

    Let (w1,w2,,wn)Gn. By Theorem 1.2, π1(x1) normally generates G, and we have that w1π1(x1)2. It follows from Lemma 2.1(ⅴ) that π1:GnG maps BX(2) to Bπ1(X)(2). Hence, there exists vBX(2) with π1(v)=w1. Clearly, v=(w1,v2,v3,,vn) for some v2,v3,,vnG. Therefore,

    (w1,w2,,wn)=v(I,v12w2,v13w3,,v1nwn).

    We will show that (I,v12w2,I,,I)BX(2). By Lemma 3.4, we can write v12w2 as a product of exactly two conjugates of g2,2 and g12,2. Also, by Remark 2.2, g2,2 and g12,2 are conjugate. Therefore, (I,v12w2,I,,I) can be written as a product of the form

    (I,v12w2,I,,I)=(I,h,I,,I)1x2(I,h,I,,I)(I,˜h,I,,I)1x12(I,˜h,I,,I),

    for some h and ˜hG. Hence, (I,v12w2,I,,I)Wx2(2)WX(2)BX(2).

    Similarly, one can see that (I,I,,I,v1iwi,I,,I,I)Bxi(2), where 3in. This implies that (I,v12w2,v13w3,,v1nwn)BX(2n2). As vBX(2), it follows that

    (w1,w2,,wn)BX(2)BX(2n2)=BX(2n).

    So Δ(Gn)2n.

    The next lemma will be needed to prove Theorem 1.5.

    Lemma 5.1. Let G be a group. Let n1 be a natural number, and let Gn=G××G. Let x=(g1,,gn)Gn. Suppose that gg1 for all gG. Then,

    Wx(m)=(Wg1(m)××Wgn(m))for allm0.

    Proof. Pick an element yWx(m). Then there exist u1,,umGn and ϵ1,,ϵm{1,1} such that y=mi=1u1ixϵiui. For 1im, let ui,1,,ui,nG such that ui=(ui,1,,ui,n). Then

    y=mi=1(ui,1,,ui,n)1(g1,,gn)ϵi(ui,1,,ui,n)=mi=1(u1i,1,,u1i,n)(gϵi1,,gϵin)(ui,1,,ui,n).

    Hence y(Wg1(m)××Wgn(m)). Take an element z=(z1,,zn) of (Wg1(m)××Wgn(m)). Then, for each 1in, we have

    zi=h1i,1gϵi,1ihi,1h1i,mgϵi,mihi,m,

    for some hi,1,,hi,mG and some ϵi,1,,ϵi,m{1,1}. Since gig1i by hypothesis, we may assume that ϵi,1,,ϵi,m=1. So we have

    z=(mi=1h11,ig1h1,i,,mi=1h1n,ignhn,i)=mi=1(h1,i,,hn,i)1(g1,,gn)(h1,i,,hn,i).

    Thus, z can be written as a product of exactly m conjugates of x=(g1,,gn). So zWx(m).

    Corollary 5.1. Let n1 be a natural number. Let x=(g1,,gn)(SL2(C))n. We have

    Wx(m)=(Wg1(m)××Wgn(m))for allm0.

    Proof. This follows directly from Remark 2.2 and Lemma 5.1.

    Remark 5.1. Let n1 be a natural number, G be a group, and x=(g1,,gn)Gn. Also, let m1 be a natural number. In general, we do not have

    Bx(m)=(Bg1(m)××Bgn(m)).

    This is demonstrated by the following example.

    Set

    g:=U(1)=(1101)SL2(C),

    and

    x:=(g,g)(SL2(C))2.

    In view of the conjugacy classes of SL2(C) and Definition 2.1, we have

    Bx(1)={(I,I)}{(M1,M2)|M1,M2SL2(C){I}andtrace(M1)=trace(M2)=2},

    and

    Bg(1)={I}{M|MSL2(C){I}andtrace(M)=2}.

    Now, we have (I,g)(Bg(1)×Bg(1)), but (I,g)Bx(1). Hence,

    Bx(1)(Bg(1)×Bg(1)).

    However, we have seen in Corollary 5.1 that

    Wx(1)=(Wg(1)×Wg(1)).

    This explains the importance of Definition 2.2.

    Proof of Theorem 1.5. We have Δ(Gn)=Δn(Gn) by Proposition 5.1 and Δ(Gn)3n by [18, Lemma 2.11(b)] and Theorem 1.1. Therefore, it suffices to show that Δ(Gn)3n.

    Let XΓ(Gn). We are going to show that Gn=BX(3n). This clearly implies that GnX3n, and as X was arbitrarily chosen, it follows that Δ(Gn)3n.

    Define surjective group homomorphisms

    πi:GnG,(g1,,gn)gi,

    for all 1in. Since X normally generates Gn, there exists some xi=(gi,1,,gi,n)X with πi(xi)±I for each 1in. In other words, we have that gi,i±I for all 1in.

    In order to complete the proof, we need the following three claims:

    Claim 5.1. Let 2jn. Then G=Wg2,j(3)Wgn,j(3).

    Proof. Let 2kn. As a consequence of Corollary 3.4, we have Wgk,j(3)=G or {I} or {I}. Also, Wgj,j(3)=G. It follows that G=Wg2,j(3)Wgn,j(3).

    Claim 5.2. We have IWg2,1(3)Wgn,1(3) or IWg2,1(3)Wgn,1(3).

    Proof. Suppose that gj,1±I for some 2jn. Then we have Wgj,1(3)=G by Corollary 3.4. It easily follows that Wg2,1(3)Wgn,1(3)=G, whence the claim holds.

    Assume now that gj,1=±I for all 2jn. Then it is clear that Wgj,1(3)={I} or {I} for all 2jn. Hence Wg2,1(3)Wgn,1(3)={I} or {I}, and the claim follows.

    Claim 5.3. Wx2(3)Wxn(3)=(Wg2,1(3)Wgn,1(3))××(Wg2,n(3)Wgn,n(3)).

    Proof. By Corollary 5.1, we have

    Wx2(3)Wxn(3)=(Wg2,1(3)××Wg2,n(3))(Wgn,1(3)××Wgn,n(3))=(Wg2,1(3)Wgn,1(3))××(Wg2,n(3)Wgn,n(3)).

    Now let (w1,w2,,wn)Gn. Our goal is to show that (w1,w2,,wn)BX(3n). By Theorem 1.1, π1(x1) normally generates G, and we have that Gπ1(x1)3. So we have w1π1(x1)3 and w1π1(x1)3. Hence, w1 and w1 are elements of Bπ1(X)(3).

    By Claim 5.2, we have IWg2,1(3)Wgn,1(3) or IWg2,1(3)Wgn,1(3). We now consider both cases.

    Case 1. IWg2,1(3)Wgn,1(3).

    By Lemma 2.1(ⅴ), there is some vBX(3) with π1(v)=w1. Clearly, v=(w1,v2,,vn) for some v2,,vnG. By Claims 5.1 and 5.3, we have

    (I,v12w2,v13w3,,v1nwn)Wx2(3)Wxn(3)BX(3n3).

    As vBX(3), it follows that

    (w1,w2,,wn)=v(I,v12w2,v13w3,,v1nwn)BX(3n).

    Case 2. IWg2,1(3)Wgn,1(3).

    By Lemma 2.1(ⅴ), there is some vBX(3) with π1(v)=w1. Clearly, v=(w1,v2,,vn) for some v2,,vnG. By Claims 5.1 and 5.3, we have

    (I,v12w2,v13w3,,v1nwn)Wx2(3)Wxn(3)BX(3n3).

    As vBX(3), it follows that

    (w1,w2,,wn)=v(I,v12w2,v13w3,,v1nwn)BX(3n).

    This completes the proof of Theorem 1.5.

    In general, calculating \Delta(G) for a group G is not easy, and the goal of this paper was to compute this value for some finitely normally generated algebraic groups over \mathbb{C}. We found the exact values of \Delta((SL_2(\mathbb{C}))^{n}) and \Delta((PSL_2(\mathbb{C}))^{n}) for every n\in \mathbb{N}. Kȩdra et al. in [18] showed that every finitely normally generated linear algebraic group G over an algebraically closed field is uniformly bounded and

    \begin{equation} \Delta(G)\leq 4\, \mathrm{dim}(G)+\Delta (G/G^{0}), \end{equation} (6.1)

    where G^{0} is the identity component of G. Thus we have improved (6.1) for the particular cases of (SL_2(\mathbb{C}))^{n} and (PSL_2(\mathbb{C}))^{n}. More precisely, we showed that the bounds in (6.1) are far from sharp. Our main results showed that for any n\in \mathbb{N}, we have \Delta((SL_2(\mathbb{C}))^{n}) = 3n (see Theorem 1.5) and \Delta((PSL_2(\mathbb{C}))^{n}) = 2n (see Theorem 1.4), whereas formula (6.1) implies that \Delta((SL_2(\mathbb{C}))^{n})\leq 12n and \Delta((PSL_2(\mathbb{C}))^{n})\leq 12n.

    We remark that \Delta(SL_2(\mathbb{C})) and \Delta(PSL_2(\mathbb{C})) have been found in [17, Theorem 3.3] and [17, Theorem 3.4]), respectively. However, we used a different approach based on the rational canonical form and Definition 2.2 to study them. This new method was needed to prove Theorems 1.4 and 1.5. Moreover, this approach could be applicable to study \Delta((SL_2(\mathbb{F}))^{n}) and \Delta((PSL_2(\mathbb{F}))^{n}), where \mathbb{F} is an arbitrary field. For example, we have that \Delta(SL_2(\mathbb{R})) = 4 (see [17, Theorem 3.1]) and \Delta(PSL_2(\mathbb{R})) = 3 (see [17, Theorem 3.2]). If we copy the same arguments as in Theorems 1.4 and 1.5, then we conjecture that \Delta((SL_2(\mathbb{R}))^{n}) = 4n and \Delta((PSL_2(\mathbb{R}))^{n}) = 3n for every n\in \mathbb{N}. We leave this in a future paper and compare that with [19, Theorem 1.4].

    This paper is based on a part of the author's PhD thesis, written at the University of Aberdeen under the supervision of Professor Benjamin Martin and Dr. Ehud Meir. The author is deeply grateful to them for their invaluable advice and support.

    The author declares no conflicts of interest.



    [1] F. Aseeri, J. Kaspczyk, The conjugacy diameters of non-abelian finite p-groups with cyclic maximal subgroups, AIMS Mathematics, 9 (2024), 10734–10755. http://dx.doi.org/10.3934/math.2024524 doi: 10.3934/math.2024524
    [2] A. Ballester-Bolinches, R. Esteban-Romero, V. Pérez-Calabuig, A note on the rational canonical form of an endomorphism of a vector space of finite dimension, Oper. Matrices, 12 (2018), 823–836. http://dx.doi.org/10.7153/oam-2018-12-49 doi: 10.7153/oam-2018-12-49
    [3] M. Brandenbursky, J. Kȩdra, On the autonomous metric on the group of area-preserving diffeomorphisms of the 2-disc, Algebr. Geom. Topol., 13 (2013), 795–816. http://dx.doi.org/10.2140/agt.2013.13.795 doi: 10.2140/agt.2013.13.795
    [4] M. Brandenbursky, Bi-invariant metrics and quasi-morphisms on groups of Hamiltonian diffeomorphisms of surfaces, Int. J. Math., 26 (2015), 1550066. http://dx.doi.org/10.1142/S0129167X15500664 doi: 10.1142/S0129167X15500664
    [5] M. Brandenbursky, Ś. Gal, J. Kȩdra, M. Marcinkowski, The cancellation norm and the geometry of bi-invariant word metrics, Glasgow Math. J., 58 (2016), 153–176. http://dx.doi.org/10.1017/S0017089515000129 doi: 10.1017/S0017089515000129
    [6] M. Brandenbursky, J. Kȩdra, E. Shelukhin, On the autonomous norm on the group of Hamiltonian diffeomorphisms of the torus, Commun. Contemp. Math., 20 (2018), 1750042. http://dx.doi.org/10.1142/S0219199717500420 doi: 10.1142/S0219199717500420
    [7] M. Brandenbursky, J. Kȩdra, Fragmentation norm and relative quasimorphisms, Proc. Amer. Math. Soc., 150 (2022), 4519–4531. http://dx.doi.org/10.1090/PROC/14683 doi: 10.1090/PROC/14683
    [8] D. Burago, S. Ivanov, L. Polterovich, Conjugation-invariant norms on groups of geometric origin, Adv. Stud. Pure Math., 2008 (2008), 221–250. http://dx.doi.org/10.2969/aspm/05210221 doi: 10.2969/aspm/05210221
    [9] D. Calegari, Stable commutator length is rational in free groups, J. Amer. Math. Soc., 22 (2009), 941–961. http://dx.doi.org/10.1090/S0894-0347-09-00634-1 doi: 10.1090/S0894-0347-09-00634-1
    [10] D. Calegari, D. Zhuang, Stable W-length, In: Topology and geometry in dimension three: triangulations, invariants, and geometric structures, Providence: American Mathematical Society, 2011,145–169.
    [11] D. Dummit, R. Foote, Abstract algebra, 3 Eds., Hoboken: John Wiley & Sons, Inc., 2004.
    [12] Ś. Gal, J. Kȩdra, On bi-invariant word metrics, J. Topol. Anal., 3 (2011), 161–175. http://dx.doi.org/10.1142/S1793525311000556
    [13] Ś. Gal, J. Gismatullin, Uniform simplicity of groups with proximal action, Trans. Amer. Math. Soc. Ser. B, 4 (2017), 110–130. http://dx.doi.org/10.1090/btran/18 doi: 10.1090/btran/18
    [14] Ś. Gal, J. Kȩdra, A. Trost, Finite index subgroups in Chevalley groups are bounded: an addendum to "on bi-invariant word metrics", J. Topol. Anal., in press. http://dx.doi.org/10.1142/S1793525323500115
    [15] J. Gismatullin, Boundedly simple groups of automorphisms of trees, J. Algebra, 392 (2013), 226–243. http://dx.doi.org/10.1016/j.jalgebra.2013.06.023 doi: 10.1016/j.jalgebra.2013.06.023
    [16] J. Humphreys, Linear algebraic groups, 1 Ed., Nwe York: Springer-Verlag, 1975. http://dx.doi.org/10.1007/978-1-4684-9443-3
    [17] J. Kȩdra, A. Libman, B. Martin, Strong and uniform boundedness of groups, arXiv: 1808.01815. http://dx.doi.org/10.48550/arXiv.1808.01815
    [18] J. Kȩdra, A. Libman, B. Martin, Strong and uniform boundedness of groups, J. Topol. Anal., 15 (2023), 707–739. http://dx.doi.org/10.1142/S1793525321500497 doi: 10.1142/S1793525321500497
    [19] J. Kȩdra, A. Libman, B. Martin, Uniform boundedness for algebraic and lie groups, arXiv: 2022.13885. http://dx.doi.org/10.48550/arXiv.2202.13885
    [20] A. Libman, C. Tarry, Conjugation diameter of the symmetric groups, Involve, 13 (2020), 655–672. http://dx.doi.org/10.2140/involve.2020.13.655 doi: 10.2140/involve.2020.13.655
    [21] D. McDuff, D. Salamon, Introduction to symplectic topology, 3 Eds, Oxford: Oxford University Press, 2017. http://dx.doi.org/10.1093/oso/9780198794899.001.0001
    [22] A. Muranov, Finitely generated infinite simple groups of infinite square width and vanishing stable commutator length, J. Topol. Anal., 2 (2010), 341–384. http://dx.doi.org/10.1142/S1793525310000380 doi: 10.1142/S1793525310000380
    [23] W. Rossmann, Lie groups: an introduction through linear groups, Oxford: Oxford University Press, 2002. http://dx.doi.org/10.1093/oso/9780198596837.001.0001
    [24] M. Suzuki, Group theory II, Berlin: Springer, 1986.
    [25] T. Tsuboi, On the uniform simplicity of diffeomorphism groups, In: Differential geometry, Singapore: World Scientific Publishing, 2009, 43–55. http://dx.doi.org/10.1142/9789814261173_0004
  • Reader Comments
  • © 2024 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(782) PDF downloads(80) Cited by(0)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog