Home Torsion, torsion length and finitely presented groups
Article Publicly Available

Torsion, torsion length and finitely presented groups

  • Maurice Chiodo EMAIL logo and Rishi Vyas
Published/Copyright: June 19, 2018

Abstract

We show that a construction by Aanderaa and Cohen used in their proof of the Higman Embedding Theorem preserves torsion length. We give a new construction showing that every finitely presented group is the quotient of some C(1/6) finitely presented group by the subgroup generated by its torsion elements. We use these results to show there is a finitely presented group with infinite torsion length which is C(1/6), and thus word-hyperbolic and virtually torsion-free.

1 Introduction

It is well known that the set of torsion elements in a group G, Tor(G), is not necessarily a subgroup. One can, of course, consider the subgroup Tor1(G)generated by the set of torsion elements in G: this subgroup is always normal in G.

The subgroup Tor1(G) has been studied in the literature, with a particular focus on its structure in the context of 1-relator groups. For example, suppose G is presented by a 1-relator presentation P with cyclically reduced relator Rk where R is not a proper power, and let r denote the image of R in G. Karrass, Magnus, and Solitar proved ([17, Theorem 3]) that r has order k and that every torsion element in G is a conjugate of some power of r; a more general statement can be found in [22, Theorem 6]. As immediate corollaries, we see that Tor1(G) is precisely the normal closure of r, and that G/Tor1(G) is torsion-free.

More generally, the manner in which Tor1(G) is impacted by the deficiencydef(G) of a finitely presented group G has also been investigated. The deficiency of G is the maximum value of m-n, where m and n are the number of generators and relators respectively as we range over all finite presentations of G. In [4, Corollary 3.6], Berrick and Hillman proved that if G is a finitely presentable group with def(G)>0, and Tor1(G) is either finitely generated or locally finite, then Tor1(G) is actually finite; again, in this situation G/Tor1(G) is torsion-free. They claim that the question of whether Tor1(G) is necessarily trivial under these hypotheses is open; using a result of Karras and Solitar [18, Main Theorem] one immediately sees that this triviality is indeed the case when G is presented by a 1-relator presentation.

In both cases described above, the quotient G/Tor1(G) is torsion-free. Unfortunately, this is not always the case. Consider, for example, the group C presented by the following presentation: x,y,zx3=e,y3=e,xy=z3; it can be shown that C is a finitely presented word-hyperbolic group ([10, Proposition 3.5]), but that C/Tor1(C)/3 ([10, Proposition 3.1]).

We can, however, iterate the process used to construct Tor1(G) to produce an ascending chain of normal subgroups Tor1(G)Tor2(G) of G. For finite n, we define Torn+1(G) via Torn+1(G)/Torn(G)=Tor1(G/Torn(G)); we define Torω(G):=nTorn(G). The ordinal for which this chain stabilises is called the torsion length of G and denoted by TorLen(G). It turns out that G/Torω(G) is the universal torsion-free quotient of G: it is torsion-free, and all other torsion-free quotients uniquely factor through it (see [9, Corollary 3.4]). Thus TorLen(G) is always bounded above by ω; this bound is attained when the chain mentioned above does not stabilise at any finite stage. Intuitively, TorLen(G) is the minimal number of times we need to ‘kill off’ torsion to get a torsion-free quotient of G.

The notion of torsion length first appeared, independently, in both [11] and our earlier work [10]. In [11], Cirio, D’Andrea, Pinzari and Rossi defined the torsion degree of a quantum group (here, quantum groups are C*-algebras equipped with a suitable comultiplication). The definition of torsion length aligns with torsion degree when a group is viewed as a quantum group via its associated C*-algebra. Further, they defined the notion of the “connected component at the identity” Q of a quantum group Q and remarked that for an ordinary group G (again viewed as a quantum group via its associated C*-algebra) this object corresponds to G/Torω(G) ([11, Example 3.17]). They also constructed a descending ordinal indexed family of quantum subgroups Gα “converging” to G; again, in the classical situation these objects correspond to the quotients G/Torα(G).

The quotient G/Torω(G) was first studied in [6], where Brodsky and Howie investigated this object (they use the notation G^) for various families of groups. A group is locally indicable if every non-trivial finitely generated subgroup admits a surjection onto : Brodsky and Howie showed that if a group has deficiency def(G)>0, then G/Torω(G) is locally indicable [6, Theorem 3.7]. They also showed that G/Torω(G) is locally indicable when G is 1-relator, or 2-relator with one relator having length 4, or 2-relator with one relator having length 5 and the other having length 8, or at most 5-relator with each relator having length 3. These results appear as [6, Theorems 1.1–1.4].

In [10] we began a preliminary investigation of torsion length. One of the main results of that work, which we generalise here in Theorem 6.13, was the following theorem:

Theorem ([10, Theorem 3.3]).

There exists a family of finitely presented groups {Pn}nN such that

  1. Pn+1/Tor1(Pn+1)Pn,

  2. TorLen(Pn)=n.

We then showed that a construction used to prove a classic embedding theorem of Higman, Neumann and Neumann (every countable group embeds into a 2-generator group) preserved torsion length. This fact, used with the theorem mentioned above, allowed us to arrive at the following result:

Theorem ([10, Theorem 3.10]).

There exists a 2-generator recursively presented group Q for which TorLen(Q)=ω.

This paper aims to extend [10, Theorem 3.10]. In Theorem 5.7, we prove the following:

Theorem 1.

There exists a finitely presented group F with TorLen(F)=ω.

We do this by showing that a particular construction used in a proof of the Higman Embedding Theorem preserves this invariant.

Let us be more precise. The Higman Embedding Theorem [16] states that a finitely generated, recursively presented group embeds into a finitely presented group. There are many proofs of this result, but these arguments share a common theme: they are all constructive. One must begin with a finite generating set for the group, and an algorithm that computes its relations, and then explicitly build a finitely presented group from this data. In this paper we pick a particular construction, due to Aanderaa and Cohen [1, 2] and presented in [12], examine it in detail, and conclude that the torsion length of the finitely presented group so constructed is the same as that of the recursively presented group that we started with.

The existence of a finitely presented group with infinite torsion length is then an immediate consequence of [10, Theorem 3.10]: take the recursively presented group constructed in [10, Theorem 3.10] and apply the Aanderaa–Cohen construction.

Section 6 of this paper is concerned with improving Theorem 5.7. The following result appears as Theorem 6.10.

Theorem 2.

There exists a finitely presented word-hyperbolic virtually special group W with TorLen(W)=ω. In particular, W is virtually torsion-free.

This is done using small cancellation theory. Of particular importance is the following theorem, whose content is contained in Proposition 6.4 and Theorem 6.7; this result is also of independent interest. Combined with Theorem 5.7, it proves Theorem 6.10.

Theorem 3.

Let P=x1,,xmr1,,rn be a finite presentation with all relators freely reduced, cyclically reduced, and distinct. For any kN, define the finite presentation Ptk:=x1,,xm,t(r1t)k,,(rnt)k,tk. Then Ptk presents a C(2/k) small cancellation group. Moreover, for k12, we have Ptk/Tor1(Ptk)P and so TorLen(Ptk)=TorLen(P)+1.

A part of the above theorem appeared in the work [7] of Bumagin and Wise; see Remark 6.11. In Section 7 we finish with a discussion of some open problems relating to torsion length and torsion subgroups.

1.1 Notation

A presentation P=XR is said to be a recursive presentation if X is a finite set and R is a recursive enumeration of relations; it is said to be a finite presentation if both X and R are finite. A group G is said to be finitely (respectively, recursively) presentable if it can be presented by a finite (respectively, recursive) presentation. If P,Q are group presentations, denote their free product presentation by P*Q: this is given by taking the disjoint union of their generators and relations. If g1,,gn are elements of a group G, we write g1,,gn for the subgroup in G generated by these elements and g1,,gnG for the normal closure of these elements in G. Let ω denote the smallest infinite ordinal. Let |X| denote the cardinality of a set X. If X is a set, let X-1 be a set of the same cardinality as, and disjoint from, X along with a fixed bijection *-1:XX-1. Write X* for the set of finite reduced words on XX-1.

2 Torn(G), HNN extensions and Britton’s lemma

Definition 2.1 ([9, Definition 3.1]).

Given a group G, define Torn(G) inductively as follows:

Tor0(G):={e},
Torn+1(G):={gGgTorn(G)Tor(G/Torn(G))}G,
Torω(G):=nTorn(G).

Observe that Tori+1(G)/Tori(G)=Tor1(G/Tori(G)).

Lemma 2.2 ([9, Corollary 3.4]).

The group G/Torω(G) is torsion-free. Moreover, if f:GH is a group homomorphism from G to a torsion-free group H, then Torω(G)ker(f), and so f factors through G/Torω(G).

Definition 2.3 ([10, Definition 2.5]).

The torsion length of G, TorLen(G), is the smallest ordinal n such that Torn(G)=Torω(G).

HNN extensions play a critical role in this paper; we briefly introduce them here.

Definition 2.4.

Let G be a group. Suppose there are isomorphisms φi:AiBi for 1in, where Ai and Bi are subgroups of G. Define the HNN extensionG*φ1,,φn with stable letters t1,,tn by

G*φ1,,φn:=(G*Fn)/{ti-1atiφi(a-1)aAi, 1in}G*Fn,

where {t1,,tn} is a free generating set of Fn. If φi=idAi for all 1in, we write G*A1,,An for G*φ1,,φn.

Definition 2.5.

Let G*φ1,,φn be an HNN extension with stable letters t1,,tn. Then a ti-pinch is a word of the form ti-1gti where gAi or tigti-1 where gBi. A word w is said to be reduced if no subword of w is a ti-pinch for any i.

Theorem 2.6 (Britton’s lemma, [21, Theorem 11.81]).

Let H=G*φ1,,φn be an HNN extension with stable letters t1,,tn, and let w be a word in H. If w=e in H, then w contains a ti-pinch as a subword, for some i.

Corollary 2.7.

If G*φ1,,φn is an HNN extension, then G embeds into G*φ1,,φn.

Given a group G, we write G;XR to denote (G*FX)/RG*FX, where R is any subset of G*FX.

3 Good subgroups of HNN extensions

The notion of a good subgroup was introduced in [12, Proposition 1.34], and named so in [23, Definition 2].

Definition 3.1.

Let H=G*φ1,,φn be an HNN extension. A good subgroup of G with respect to the HNN extension H is a subgroup KGH such that φi(KAi)=KBi for all 1in.

Lemma 3.2 ([12, Proposition 1.34]).

Let H:=G*φ1,,φn be an HNN extension of G with stable letters t1,,tn. Suppose KG is a good subgroup of G with respect to the HNN extension H, and let ψi:KAiKBi be the restriction of φi to KAi. Let K be the subgroup of H generated by K and t1,,tn. Then the natural map

νK:K*ψ1,,ψnK

is an isomorphism. Moreover, KG=K.

We now study good subgroups which are normal.

Definition 3.3.

Let H:=G*φ1,,φn be an HNN extension of G with stable letters t1,,tn. Let KG be a good subgroup of G with respect to the HNN extension H. Let φ¯i:Ai/(KAi)Bi/(KBi) be the induced isomorphism for each 1in. Define the following HNN extension with stable letters t¯1,,t¯n:

HK:=(G/K)*φ¯1,,φ¯n.

There is a surjective homomorphism

ϕK:HHK

which sends ggK for all gG, and tit¯i for all 1in.

Lemma 3.4.

Let G be a group, and H:=G*φ1,,φn an HNN extension of G with stable letters t1,,tn. Let KG. Then K is a good subgroup of G with respect to the HNN extension H if and only if KHG=K in H.

Proof.

() Assume that KHG=K in H. Take 1in and suppose xAiK. We know that φi(x)Bi: we need to verify φi(x)K. However, it is immediate that φi(x)=ti-1xtiKH, and thus that

φi(x)KHBi=KHGBi=KBi;

it follows that φi(AiK)BiK. The inclusion φi(AiK)BiK can be proved in a similar fashion.

() Suppose K is a good subgroup of G with respect to the HNN extension H, and take ϕK as in Definition 3.3. Then it is clear that

KKHGker(ϕK)GK;

the last inequality here is a consequence of Theorem 2.6. ∎

Lemma 3.5.

Let H, K and ϕK be as in Definition 3.3. Then ker(ϕK)=KH.

Proof.

The containment KHker(ϕK) is immediate.

Let xker(ϕK). We induct on the total number of occurrences of ti or ti-1 over all i, where 1in, in the normal form of x in H: if x has none, then xK.

Assume that for some i, either ti or ti-1 appears at least once in x. By Britton’s lemma ϕK(x) has a subword of the form t¯i-1at¯i where aAi/(AiK) or t¯ibt¯i-1 where bBi/(BK). Thus x has a subword of the form ti-1ati where aAiK or tibt-1 where bBiK. Without loss of generality, we assume the former.

This subword ti-1at is of the form ti-1akti, where aAi and kK. But ti-1ati=bB for some bB. We can therefore write x as

λ1ti-1aktiλ2=λ1bti-1ktiλ2.

Observe that ti-1ktiKH, and thus that ti-1ktiλ2=λ2y, where yKH. We can therefore rewrite x=λ1bλ2y; from this we see that λ1bλ2ker(ϕK). By induction, we have that λ1bλ2KH. This tells us that xKH. ∎

Corollary 3.6.

Let H, K and HK be as in Definition 3.3. Then ϕK induces an isomorphism

ϕ¯K:H/KHHK.

4 The Higman embedding construction

The Higman Embedding Theorem states that a finitely generated, recursively presented group can be embedded in a finitely presented group. In this section we provide an overview of a proof of this result, introducing notation and constructions that will be used later in this paper.

4.1 Modular machines and their connection to Turing machines

Modular machines are an alternative way of formalising the notion of mechanical computation: they simulate Turing machines in a very natural way using integers rather than tapes. This can often be useful in group theoretic applications; for example, there is a proof of the Higman embedding theorem using modular machines (due to Aanderaa and Cohen and described in detail in Section 4.3) which is particularly transparent for the purposes of this paper.

Definition 4.1.

A modular machine consists of an integer m>1 and a finite set of quadruples each of the form (a,b,c,R) or (a,b,c,L), where m>a0 and m>b0 and m2>c0. We require that, for each such pair (a,b), there is at most one quadruple of of the form (a,b,*,*).

A modular machine configuration is an ordered pair (α,β)2. We write

(α,β)(α1,β1),

called a computational step of , if α=um+a and β=vm+b (with 0a,b<m) and there exists c such that either

  1. (a,b,c,R) and α1=um2+c and β1=v, or

  2. (a,b,c,L) and α1=u and β1=vm2+c.

Note that the action of on (α,β) depends only on the class of (α,β) modulo m. This is why we call a modular machine.

We write (α,β)*(α,β) if there exists a finite sequence

(α,β)=(α1,β1)(α2,β2)(αn,βn)=(α,β).

Such a sequence is called a computation of .

If, for α=um+a, β=vm+b (0a,b<m), no quadruple of begins with (a,b), then we say (α,β) is terminal. If (0,0) is terminal in , then we define the halting set of, denoted H0(), by

H0():={(α,β)(α,β)*(0,0)}.

The following result by Aanderaa and Cohen [1] (paraphrased), along with an analysis of its proof, shows that for each Turing machine T there is a modular machine (T) which simulates the action of T and conversely, that any modular machine can be simulated by a Turing machine. Thus these two notions of computation are equivalent. One can find a more detailed discussion of this material in [1].

Theorem 4.2 ([1, Theorem 2]).

Let T be a Turing machine. From T, we can construct a modular machine M(T) whose halting set H0(M(T)) is computationally equivalent to the halting set Ω(T) of T. Stated formally: Ω(T)mH0(M(T)).

4.2 Simulating a modular machine by a finitely presented group

We begin by describing how a modular machine can be simulated by a finitely presented group. This construction is then used in a proof of the Higman Embedding Theorem.

The idea is to follow the construction in [12, pp. 266–268]. This was derived from [1], where a detailed exposition of modular machines can be found. We felt, however, that the exposition in [23] was slightly clearer, so we replicate here the argument presented there (the differences are only slight).

  1. Define the group K:=x,y,t[x,y]=e.

  2. For all (r,s)2, define the word t(r,s):=y-sx-rtxrysK.

  3. Let T:={t(r,s)}(r,s)2K.

  4. Observe that T is free with basis {t(r,s)}(r,s)2.

  5. Observe that T=tK.

  6. For M>a0, N>b0, define

    Ka,bM,N:=t(a,b),xM,yNK,
    Ta,bM,N:={t(α,β)αamodM,βbmodN}TK.

  7. Let (i,j)2, and m,n. Observe that Tt(i,j),xm,yn is free with basis {t(r,s)rimodm,sjmodn}. In particular,

    TKa,bM,N=Ta,bM,N.
  8. Observe that the correspondence tt(a,b), xxM, yyN induces an isomorphism

    KKa,bM,N.

    This isomorphism sends t(u,v) to t(uM+a,vN+b) and thus induces an isomorphism

    TTa,bM,N.
  9. Let ={(ai,bi,ci,R)iI}{(aj,bj,cj,L)jJ} be a modular machine with modulus m.

  10. The maps in step (8) induce, for each iI and jJ, isomorphisms

    ϕi:Kai,bim,mKci,0m2,1,φj:Kaj,bjm,mK0,cj1,m2.
  11. Form the HNN extension

    K:=K*{ϕi}iI,{φj}jJ,

    with stable letters {ri}iI and {lj}jJ. Note that K is finitely presented.

  12. Define the subgroup T:=T,{ri}iI,{lj}jJK, where T is as in step (3).

  13. Define the set H0():={(α,β)(α,β)*(0,0)}.

  14. Define T:={t(α,β)(α,β)H0()}K.

  15. Define T:=T,{ri}iI,{lj}jJK.

  16. Observe that T=t,{ri}iI,{lj}jJ.

  17. Observe that t(α,β)T if and only if (α,β)H0().

  18. With the identity map θ:TT, form the HNN extension

    G:=K*θ

    with stable letter q.

  19. Observe that q-1t(α,β)q=t(α,β) in G if and only if (α,β)H0().

Taking with nonrecursive halting set H0() gives a finitely presented group G with undecidable word problem.

For our purposes, a useful consequence of the above construction is that we can simulate any modular machine by a finitely generated group: see step (19).

4.3 The Higman Embedding Theorem

We now give an overview of the construction used in a particular proof of the Higman Embedding Theorem, taken directly from [12, pp. 279–281]. We note that this proof originally comes from [2].

  1. Let C=c1,,cnS be a finitely generated recursively presented group, where S corresponds to the set H0() of a modular machine ; see step (7) below. Denote the modulus of by m. We assume that S covers all the trivial words in the group.

  2. Re-write every word in C as a word in the free monoid on {c1,,c2n} with ci-1 replaced by cn+i.

  3. To each word w=cikcik-1ci0 associate an m-ary representation

    α=j=0kijmj.
  4. Define I:={αα represents a word}. That is,

    α=j=0kβjmj,

    where 1βj2n.

  5. For αI, define wα(c) to be the word formed from α.

  6. For αI, write wα(b), wα(bc) for the words obtained from wα(c) by replacing ci with bi and bici, respectively (where {bi}i=12n is a new set of symbols).

  7. Observe that, for all αI, we have

    wα(c)S(α,0)H0(M).
  8. Recall the group K from step (11) of Section 4.2. Define

    U:={t,{ri}iI,{lj}jJ};

    U is a subset of K.

  9. Define tα:=t(α,0)K.

  10. Form the free product

    H1:=K*(C×b1,,bn-)*d-,

    and set bn+i:=bi-1 for 1in.

  11. Observe that the sets {tααI} and {tαwα(b)dαI} each form a free basis for the subgroups they respectively generate in H1. The correspondence tαtαwα(b)d extends to an isomorphism ψ between these subgroups.

  12. Form the HNN extension

    H2:=H1*ψ

    with stable letter p.

  13. Define the subgroup

    A:=t,x,d,b1,,bn,pH2.
  14. For 1i2n, define the subgroup

    Ai:=ti,xm,bid,b1,,bn,pH2.
  15. Observe that for all i, A is isomorphic to Ai via the map ψi induced by the correspondence sending tti, xxm, dbid, bjbj for all  1jn, and pp.

  16. Observe that t,x,d,b1,,bn and ti,xm,bid,b1,,bn are both good in H1 with respect to the HNN extension H2. Therefore A, and the Ai for 1i2n, are all HNN extensions.

  17. Define the subgroup

    A+:=U,d,b1,,bn,pH2.
  18. Define the subgroup

    A-:=U,d,b1c1,,bncn,pH2.
  19. Observe that U,d,b1,,bn is good in H1 with respect to the HNN extension H2. Therefore A+ is an HNN extension.

  20. Observe that A+ is isomorphic to A- via the map ψ+:A+A- induced by the correspondence sending uu for all uU, dd, bjbjcj for all 1jn, and pp.

  21. With the isomorphisms defined above, define the HNN extension

    H3:=H2*ψ1,,ψ2n,ψ+

    with stable letters a1,,a2n and k.

  22. Observe that H3 is finitely presented, and CH3.

5 Properties of the embedding construction

In this section, the groups C,H1,H2 and H3 will be as in Section 4.3.

Lemma 5.1.

Let X be a subset of C. Then:

  1. XH1 is good in H1 with respect to the HNN extension H2.

  2. XH2 is good in H2 with respect to the HNN extension H3.

Proof.

We claim that the following is true:

XH1{tααI}={e},
XH1{tαwα(b)dαI}={e}.

To see this, consider the map λ:H1H1 induced by the identity maps on K, b1,,bn-, and d-, and the trivial map on C.

The map λ, restricted to KM*({e}×b1,,bn-)*d-, is injective, and thus injective on both {tααI} and {tαwα(b)dαI}. However, XH1 is contained in ker(λ).

This proves the first part of the lemma; we now move to the second.

Take the map λ defined above. It is clear that λ extends to a map λ¯:H2H2, sending pp. Again, XH2ker(λ¯). As before, we see that the restriction of λ¯ to KM*({e}×b1,,bn-)*d-,p is injective. It follows that

XH2A=XH2Ai={e}

for all 1i2n.

Finally, consider the inclusions

ι-:A-H2,ι+:A+H2.

Step (20) of Section 4.3 tells us that the restriction of λ to A- is injective with image A+, and thus induces an isomorphism λ:A-A+; λ is inverse to the map ψ+ defined in (15) of Section 4.3. We see that λ¯ι+λ=λ¯ι-:A-H2:

It is clear that λ¯ι+ is injective, and thus that λ¯ι- is as well. Since XH2 is contained in ker(λ¯), we see that XH2A-=XH2A+={e}. This proves the last part of the lemma. ∎

Before we proceed, we need the following observation. It is proved in the same way that [10, Corollary 2.9] is, by using the torsion theorem for HNN extensions.

Lemma 5.2.

Let G be a group, and φ:HK an isomorphism between subgroups H,KG. Let G*φ be the associated HNN extension. Then

Tor1(G*φ)=Tor1(G)G*φ=Tor(G)G*φ.

Lemma 5.3.

For all m0, the following hold:

  1. Torm(H1)=Torm(C)H1.

  2. Torm(H1)C=Torm(C).

Proof.

By [10, Proposition 2.10], we know that

Torm(H1)=Torm(K)Torm(C×b1,,bn-)Torm(d-)H1.

However, K, b1,,bn- and d- are all torsion-free. It follows that

Torm(H1)=Torm(C)H1

for all m. This proves part (1). For the second part, observe that there is a map μ:H1C induced by the trivial map on K and d- and the standard projection to C on C×b1,,bn-. The map μ restricts to the identity on C and sends Torm(H1) to Torm(C). The result follows. ∎

Lemma 5.4.

For i=1,2, and for all m0, the following hold:

  1. Torm(Hi+1)=Torm(Hi)Hi+1.

  2. Torm(Hi) is good in Hi with respect to the HNN extension Hi+1.

Proof.

We prove this by induction on m. The result is obvious for m=0.

We now come to the inductive step. Let i{1,2}. Assume the statement is true for m. The induction hypothesis tells us that Torm(Hi+1)=Torm(Hi)Hi+1 and that Torm(Hi) is good in Hi with respect to the HNN extension Hi+1. Thus, by Lemma 3.4,

Torm(Hi)Hi+1Hi=Torm(Hi).

Combining these facts, we see that Torm(Hi+1)Hi=Torm(Hi). As a consequence, the inclusion HiHi+1 induces an embedding

Hi/Torm(Hi)Hi+1/Torm(Hi+1);

via this, we identify Hi/Torm(Hi) as a subgroup of Hi+1/Torm(Hi+1).

Using Lemma 3.5 and Corollary 3.6, we see that

Hi+1/Torm(Hi+1)=Hi+1/Torm(Hi)Hi+1
Hi/Torm(Hi);stableirelationsi,

where stablei, relationsi are, respectively, the stable letters and relations of the HNN construction of Hi+1 from Hi. It then follows from Lemma 5.2 that

Tor1(Hi+1/Torm(Hi+1))=Tor1(Hi/Torm(Hi))Hi+1/Torm(Hi+1).

The preimage of Tor1(Hi+1/Torm(Hi+1)) in Hi+1 is Torm+1(Hi+1); the preimage of Tor1(Hi/Torm(Hi)Hi+1/Torm(Hi+1) in Hi+1 is Torm+1(Hi)Hi+1. Thus Torm+1(Hi+1)=Torm+1(Hi)Hi+1, and (1) is proved for the case m+1.

We have just proved that part (1) is true for m+1; combining this fact with Lemma 5.3 (1), we see that

Torm+1(Hi)=Torm+1(Hi-1)Hi
==Torm+1(H1)Hi=Torm+1(C)Hi.

Lemma 5.1 then tells us that Torm+1(Hi) is good in Hi with respect to the HNN extension Hi+1. ∎

The next corollary now follows from Lemmas 5.3, 5.4 and 3.4:

Corollary 5.5.

For i=1,2,3, and for all m0, the following hold:

  1. Torm(Hi)=Torm(C)Hi.

  2. Torm(Hi)C=Torm(C).

Theorem 5.6.

There is a uniform construction that, on input of a recursive presentation of a group C, outputs a finite presentation of a group H in which C embeds, with TorLen(C)=TorLen(H).

Proof.

This is an immediate consequence of Corollary 5.5, taking H=H3. As

Torm(C)H3=Torm(H3)

for all m, Torm(C)=Torm+1(C) implies that Torm(H3)=Torm+1(H3).

Conversely, since

Torm(H3)C=Torm(C)

for all m, Torm(H3)=Torm+1(H3) implies that Torm(C)=Torm+1(C).

In conclusion, Torm(H3)=Torm+1(H3) if and only if Torm(C)=Torm+1(C) for any m. Thus the sequences Torj(H3) and Torj(C) stabilise at precisely the same value of j (if at all), and so TorLen(H3)=TorLen(C). ∎

Theorem 5.7.

There exists a finitely presented group F with TorLen(F)=ω.

Proof.

In [10, Theorem 3.10], we proved that there is a 2-generator, recursively presented group with infinite torsion length. We now apply Theorem 5.6. ∎

An interesting exercise would be to construct an explicit finite presentation of such a group. Theoretically, this could be done by carefully following the constructions given above. The presentation that arises as the output of such a process, however, would undoubtedly be very complicated. A more straightforward presentation, perhaps giving a group that arises elsewhere in the literature, would be interesting.

6 A word-hyperbolic virtually special example

We now show various ways of constructing finitely presented virtually special groups with infinite torsion length. We thank Henry Wilton for initially suggesting that this is possible and pointing out an alternate way to prove it.

Definition 6.1.

Let Γ be an undirected graph on finite vertex set labeled 1,,n, and edge set E. The right-angled Artin group (RAAG), A(Γ), associated to Γ is the group with presentation

x1,,xn[xi,xj] for all {i,j}E.

A group G is said to be special if it is a subgroup of some RAAG. More generally, a group G is said to be virtually special if it contains a finite index subgroup which is special.

Every RAAG on n generators can be seen as an HNN extension of a RAAG on n-1 generators; it follows that every RAAG is torsion-free, and thus that virtually special groups are virtually torsion-free.

For the remainder of this section, if P=XR is a group presentation we denote by P¯ the group presented by P, and if wX* is a word in the generators of P then we denote by w¯ the element of P¯ represented by w.

Definition 6.2.

Let P=XR be a presentation, where each rR is freely reduced and cyclically reduced (as a word in X*), and where R is symmetrised (i.e., closed under taking cyclic permutations and inverses).

A nonempty freely reduced word wX* is called a piece with respect to P if there exist two distinct elements r1,r2R that have w as maximal common initial segment.

Let 0<λ<1. Then P is said to satisfy the C(λ)small cancellation condition if whenever w is a piece with respect to P and w is a subword of some rR, then |u|<λ|r| as words.

A group is called a C(λ) group if it can be presented by a presentation satisfying the C(λ) small cancellation condition.

If P=XR is a presentation of a group G where R is not symmetrised, we can take the symmetrised closure Rsym of R, where Rsym consists of all cyclic permutations of words in R and R-1 (with repetitions removed). Then Rsym is symmetrised and Psym=XRsym is also a presentation of G. In a slight abuse of notation, we call the presentation P=XR symmetrised if R is symmetrised. Observe that if R is finite, then so is Rsym.

The following theorem is a consequence of the substantial results of Agol [3] and Wise [24].

Theorem 6.3.

Let P=XR be a finite presentation satisfying the C(1/6) small cancellation condition. Then P¯ is both word-hyperbolic and virtually special.

Proof.

Finitely presented C(1/6) groups are known to be word-hyperbolic (see [13, 14]). One then uses [24, Theorem 1.2] and [3, Theorem 1.1] to show that P¯ is virtually special. ∎

Proposition 6.4.

Let P=XR={r1,r2,} be a presentation, with all words in R freely reduced, cyclically reduced, and distinct. For any kN, define the presentation

Ptk:=X,ttk,(r1t)k,(r2t)k,.

Then (Ptk)sym is symmetrised and satisfies the C(2/k) small cancellation condition. If P is finite, then Ptk¯ is word-hyperbolic and virtually special for all k12.

Proof.

Let S={tk,(r1t)k,(r2t)k,}. We first need to check that every sSsym is freely and cyclically reduced. But this follows from the fact that R is freely and cyclically reduced, along with the strategic placements of the t’s. By definition, Ssym is symmetrised. We now show that Ssym satisfies the C(2/k) small cancellation condition.

Step (1). Any cyclic permutation of (rit)k shares a piece with tk of length at most one (and no piece with t-k). Similarly, any cyclic permutation of (rit)-k shares a piece with t-k of length at most one (and no piece with tk). Such pieces have length at most 1/k of either word.

Step (2). Consider shared pieces of cyclic permutations of pairs of words of the form (rit)k and (rjt)k. If ri=ab and rj=cd, where a,b,c,d are words, then we are left with considering words of the form bt(rit)k-1a and dt(rjt)k-1c respectively. As rirj, the initial overlap of these can be at most btdt, followed by some overlap of ri and rj (of length at most min{|ri|,|rj|}, as rirj and t is acting as an end marker). So this initial overlap can have length at most min{2|ri|,2|rj|}+1 which is less than 2/k of the length of either word.

Step (3). By repeating the arguments as in step 2, we can show that cyclic permutations of pairs of words of the form (rit)-k and (rjt)-k overlap at most 2/k of the length of either word.

Step (4). We now consider shared pieces of cyclic permutations of pairs of words of the form (rit)k and (rjt)-k=(t-1rj-1)k. If ri=ab and rj=cd, where a, b, c, d are words, then the words we are considering must be of the form bt(rit)k-1a and c-1(t-1rj-1)(k-1)t-1d-1, respectively. An initial overlap cannot involve t or t-1, and thus has length at most min{|ri|,|rj|}; this is less than 1/k of the length of either word.

It follows that (Ptk)sym satisfies the C(2/k) small cancellation condition; in the case where P is finite, we appeal to Theorem 6.3 to finish the proof. ∎

The following standard result was first proved in [13]; see [22, Theorem 6] for an explicit statement of the result.

Lemma 6.5 ([13, Theorem VIII]).

Let P=XR satisfy the C(1/6) small cancellation condition. Then an element gP¯ has order n>1 if and only if there is a relator rR of the form r=sn in X*, with sX*, such that g is conjugate to s¯ in P¯.

Lemma 6.6.

Let P=XR={r1,r2,} be a presentation, with all words in R freely reduced, cyclically reduced, and distinct. Let Ptk be as before, with k12. Then

Tor1(Ptk¯)=t¯,r1t¯,r2t¯,Ptk¯.

Proof.

(In this proof, we take all normal closures to be in Ptk¯.) Clearly the elements of {t¯,r1t¯,r2t¯,} are all torsion elements in Ptk¯, and so we have that

t¯,r1t¯,r2t¯,Tor1(Ptk¯).

To show the converse, it suffices to show that

Tor(Ptk¯)t¯,r1t¯,r2t¯,.

So take some torsion element gTor(Ptk¯) with o(g)=n. By Proposition 6.4, (Ptk)sym satisfies the C(1/6) small cancellation condition; thus, by Lemma 6.5, g is conjugate to some s¯ with sn=r for some relator r of (Ptk)sym. If r=tk or t-k, then s is a power of t and so g is conjugated into t¯,r1t¯,r2t¯,.

Otherwise, sn is equal to some cyclic permutation of some (rit)k or (rit)-k; it is enough to just consider the first case. Then there is some cyclic permutation q of s such that qn=(rit)k as words in X*.

What word q can we have which satisfies (rit)k=qn? If |q|<|ri|, then q contains no t and thus qn contains no t; a contradiction. If |q|=|rit|, then q=rit and so g is conjugate to q¯=rit¯ which clearly lies in t¯,r1t¯,r2t¯,. If |q|>|rit|, then q=(rit)za, where ri=ab is a decomposition of ri. Thus qn=((rit)za)n, and this can only be equal to (rit)k if a=. In this case q=(rit)z for some z, and so g is conjugate to q¯=(rit)z¯ which lies in t¯,r1t¯,r2t¯,. Thus

Tor(Ptk¯)t¯,r1t¯,r2t¯,.

Theorem 6.7.

Let P=XR be a finite presentation with all words in R freely reduced, cyclically reduced, and distinct. Then, for any k12, Ptk¯ is word-hyperbolic, virtually special, and satisfies

Ptk¯/Tor1(Ptk¯)P¯.

Thus, in this case, TorLen(Ptk¯)=TorLen(P¯)+1.

Proof.

This follows immediately from Proposition 6.4 and Lemma 6.6. ∎

Remark 6.8.

One may ask why the introduction of the extra generator t is necessary when constructing Ptk. It is indeed true that, given a finite presentation P=x1,,xmr1,,rn, the finite presentation

Q=x1,,xmr1k1,,rnkn

(where k1,,kn1) presents a group Q¯ with

TorLen(Q¯)-1TorLen(P¯)TorLen(Q¯).

The reader can easily verify this. However, it is not necessarily the case that Q¯/Tor1(Q¯)P¯. As an example, we can consider P=x,y,zx,y3,xy=z3 and Q=x,y,zx3,y3,xy=z3. It is clear that P is just a presentation for the cyclic group with nine elements, C9. On the other hand, by [10, Proposition 3.1], Q¯/Tor1(Q¯)C3.

Remark 6.9.

In [15, Corollary 1.4] it has been shown that every finitely generated C(1/6) group is acylindrically hyperbolic (this notion was first defined in [19]). Using this, the assumption that P is finite can be relaxed in Theorem 6.7 if we allow ourselves a slightly weaker conclusion. If we continue to assume that X is finite while no longer requiring R to be so, then [15, Corollary 1.4] implies – assuming the notation of Theorem 6.7 – that Ptk¯ is acylindrically hyperbolic. It still follows from Lemma 6.6 that

Ptk¯/Tor1(Ptk¯)P¯.

We thank the anonymous referee for bringing this to our attention.

Theorem 6.10.

There exists a finitely presented word-hyperbolic virtually special group W with TorLen(W)=ω. In particular, W is virtually torsion-free.

Proof.

Let P be a finite presentation of a group with infinite torsion length; such things exist, by Theorem 5.7. Then, by Theorem 6.7, Pt12¯ is hyperbolic and virtually special, and TorLen(Pt12¯)=TorLen(P¯)+1=ω. Take W=Pt12¯. ∎

Remark 6.11.

The main construction in [7] can be used to obtain a similar result to Theorem 6.7 above. Given a finite presentation P=AR, we see in [7, p. 141, equation (4)] an explicit finite presentation of a C(1/6) group G and NG such that G/NP¯, and moreover that P¯ is isomorphic to Out(N) ([7, Theorem 11]). Further analysis shows that N=Tor1(G), normally generated by two elements. However, both the finite presentation of G in [7] and its manner of construction seem to be substantially more complicated than the finite presentation Pt12 constructed above

Remark 6.12.

In [10, Lemma 2.3] we showed that Tori(H)Tori(G) whenever HG. However, this does not extend to bounding torsion length of subgroups, even for finitely presented groups. Using the fact that there are finitely presented groups of any torsion length ([10, Theorem 3.3]), including ω (Theorem 5.7), along with the Adian–Rabin construction ([8, Theorem 2.4]), one can show that given any finitely presented group H, and any ordinal 1nω, there is a finitely presented group of torsion length n into which H embeds.

We finish this section with an alternate construction for, and strengthening of, the result obtained as [10, Theorem 3.3].

Theorem 6.13.

Define the sequence of finite presentations

P0:=--,P1:=t1t112,P2:=t1,t2(t112t2)12,t212,

and, in general,

Pn:=(Pn-1)tn12=t1,,tn((t112t2)12)tn)12,,(tn-112tn)12,tn12.

Then P¯n is a C(1/6) (and therefore word-hyperbolic and virtually special) group, P¯n/Tor1(P¯n)P¯n-1, and TorLen(P¯n)=n.

Proof.

This follows immediately from Theorem 6.7. ∎

7 Quotients

We are interested in the universal torsion-free quotients of finitely presented groups. We begin with the following observation.

Proposition 7.1.

Let G be a finitely presented group with infinite torsion length (see Theorem 6.10). Then G/Torω(G) is finitely generated and recursively presented, but not finitely presented.

Proof.

Assume G/Torω(G) is finitely presented. Then Torω(G) must be the normal closure of finitely many elements of G; say Torω(G)=g1,,gnG. But then each gi lies in some Tormi(G), and as the Torj(G) form a nested sequence, all the gi lie in TorM(G) for M=max{mi}. Thus Torω(G)=TorM(G), and so G has finite torsion length; a contradiction. ∎

With this in mind, we ask the following question:

Question 1.

Is there a finitely presented group G for which G/Tor1(G) is recursively presented but not finitely presented?

Note that if such a group were to exist, then using the Adian–Rabin construction ([8, Theorem 2.4]) one could construct a group G such that any sequence drawing from “finitely presented” and “not finitely presented” is realised by looking at the sequence G/Tor1(G),G/Tor2(G), .

In the case of word-hyperbolic groups, however, it is always true that the quotient G/Tor1(G) is finitely presented, as we now show; moreover, in this context a finite presentation for G/Tor1(G) can be algorithmically constructed. We begin with a result of Papasoglu.

Theorem 7.2 ([20]).

There is a partial algorithm that, on input of a finite presentation P, halts if and only if P¯ is a word-hyperbolic group. Moreover, when this algorithm does halt, it outputs a hyperbolicity constant δ for P.

For a finitely generated group G with finite generating set X, we define the ball of radius r about the identity, BX(e,r), to be the set of elements

BX(e,r):={gGthere exists wX* with |w|r and w¯=g in G}.

The following standard lemma will be of use; the proof of [5, III.Γ, Theorem 3.2] provides an argument to verify it:

Lemma 7.3.

Let G be a finitely presented word-hyperbolic group with hyperbolicity constant δ. Then any finite subgroup HG is conjugate in G to some subgroup in the (4δ+2)-ball around the origin. That is, there exists some gG such that g-1HgB(e,4δ+2).

Theorem 7.4.

Let P=XR be a finite presentation of a word-hyperbolic group G with hyperbolicity constant δ. Let SX,δ be the finite set

SX,δ:={gTor(G)gBX(e,4δ+2)}

Then Tor(G)G=SX,δG. Moreover, from P and δ we can algorithmically construct the set SX,δ.

Proof.

Let g be a torsion element in G. Then, by Lemma 7.3, g is conjugate to a subgroup in the ball BX(e,4δ+2). Thus Tor(G)G=SX,δG, and so the first statement is proved.

Now, using the uniform solution to the word problem for hyperbolic groups (see [5, III.Γ, Theorems 2.4–2.6]), we can identify a set of words (of length at most r) together representing all elements in SX,δ as follows: enumerate all words of length at most r in X*; call these w1,,wk. For each wi, compute minimal-length words for wi2,wi3, and so on until either some wim¯ lies outside the ball BX(e,4δ+2) or is trivial. If, for wi, the former occurs first, then discard wi. If, for wi, the latter occurs first, then add wi to our set. At the end of this process, we will have formed the set SX,δ, algorithmically from P and δ. ∎

Using Theorems 6.7, 7.2 and 7.4, we immediately see the following:

Corollary 7.5.

Let G be a finitely presented word-hyperbolic group. Then the quotient G/Tor1(G) is finitely presented. Moreover, given a finite presentation P for the group G, we can algorithmically construct from it a finite presentation for G/Tor1(G). Finally, any finitely presented group Q can be obtained as QG/Tor1(G) for some C(1/6) (and therefore word-hyperbolic) group G.

Remark 7.6.

Indeed, every finitely generated group H can be obtained as HG/Tor1(G) for some C(1/6) (and hence acylindrically hyperbolic) group G: see Remark 6.9. We thank the anonymous referee for pointing this out.


Communicated by George Willis


Award Identifier / Grant number: 170/12

Award Identifier / Grant number: 253/13

Award Identifier / Grant number: 659102

Award Identifier / Grant number: FN PP00P2-144681/1

Funding statement: The first author was partially supported by the Swiss National Science Foundation grant FN PP00P2-144681/1. This work is part of a project that has received funding from the European Union’s Horizon 2020 research and innovation programme under the Marie Skłodowska-Curie grant agreement No. 659102. The second author was partially supported by Israel Science Foundation grants 170/12 and 253/13, the Center for Advanced Studies in Mathematics at Ben-Gurion University of the Negev, and by the Israel Council for Higher Education.

Acknowledgements

We thank Henry Wilton, Mark Hagen, Ben Barrett and Alan Logan for their assistance in preparing Sections 6 and 7, and Andrew Glass, Jack Button, and the anonymous referee for their comments and suggestions.

References

[1] S. Aanderaa and D. Cohen, Modular machines, the word problem for finitely presented groups and Collins’ theorem, Word Problems II (Oxford 1976), Stud. Log. Found. Math. 95, North-Holland, Amsterdam (1980), 1–16. 10.1016/S0049-237X(08)71327-2Search in Google Scholar

[2] S. Aanderaa and D. Cohen, Modular machines and the Higman–Clapham–Valiev embedding theorem, Word Problems II (Oxford 1976), Stud. Log. Found. Math. 95, North-Holland, Amsterdam (1980), 7–28. 10.1016/S0049-237X(08)71328-4Search in Google Scholar

[3] I. Agol, The virtual Haken conjecture, Doc. Math. 18 (2013), 1045–1087. 10.4171/dm/421Search in Google Scholar

[4] A. Berrick and J. Hillman, Perfect and acyclic subgroups of finitely presentable groups, J. Lond. Math. Soc. 68 (2003), no. 3, 683–698. 10.1112/S0024610703004587Search in Google Scholar

[5] M. Bridson and A. Haefliger, Metric Spaces of Non-Positive Curvature, Springer, Berlin, 1999. 10.1007/978-3-662-12494-9Search in Google Scholar

[6] S. D. Brodsky and J. Howie, The universal torsion-free image of a group, Israel J. Math. 98 (1997), 209–228. 10.1007/BF02937335Search in Google Scholar

[7] I. Bumagin and D. T. Wise, Every group is an outer automorphism group of a finitely generated group, J. Pure Appl. Algebra 200 (2005), no. 1, 137–147. 10.1016/j.jpaa.2004.12.033Search in Google Scholar

[8] M. Chiodo, Finding non-trivial elements and splittings in groups, J. Algebra 331 (2011), 271–284. 10.1016/j.jalgebra.2010.12.013Search in Google Scholar

[9] M. Chiodo, On torsion in finitely presented groups, Groups Complex. Cryptol. 6 (2014), no. 1, 1–8. 10.1515/gcc-2014-0001Search in Google Scholar

[10] M. Chiodo and R. Vyas, A note on torsion length, Comm. Algebra 43 (2015), no. 11, 4825–4835. 10.1080/00927872.2014.974249Search in Google Scholar

[11] L. Cirio, A. D’Andrea, C. Pinzari and S. Rossi, Connected components of compact matrix quantum groups and finiteness conditions, J. Funct. Anal. 267 (2014), no. 9, 3154–3204. 10.1016/j.jfa.2014.08.022Search in Google Scholar

[12] D. Cohen, Combinatorial Group Theory: A Topological Approach, Cambridge University Press, Cambridge, 1989. 10.1017/CBO9780511565878Search in Google Scholar

[13] M. Greendlinger, On Dehn’s algorithms for the conjugacy and word problems, with applications, Comm. Pure Appl. Math. 13 (1960), 641–677. 10.1002/cpa.3160130406Search in Google Scholar

[14] M. Gromov, Hyperbolic groups, Essays in Group Theory, Math. Sci. Res. Inst. Publ. 8, Springer, New York (1987), 75–263. 10.1007/978-1-4613-9586-7_3Search in Google Scholar

[15] D. Gruber and A. Sistro, Infinitely presented graphical small cancellation groups are acylindrically hyperbolic, preprint (2016), https://arxiv.org/abs/1408.4488v3. 10.5802/aif.3215Search in Google Scholar

[16] G. Higman, Subgroups of finitely presented groups, Proc. Roy. Soc. Lond. Ser. A 262 (1961), 455–475. 10.1098/rspa.1961.0132Search in Google Scholar

[17] A. Karrass, W. Magnus and D. Solitar, Elements of finite order in groups with a single defining relation, Comm. Pure Appl. Math. 13 (1960), 57–66. 10.1002/cpa.3160130107Search in Google Scholar

[18] A. Karrass and D. Solitar, One relator groups having a finitely presented normal subgroup, Proc. Amer. Math. Soc. 69 (1978), no. 2, 219–222. 10.1090/S0002-9939-1978-0466323-3Search in Google Scholar

[19] D. Osin, Acylindrically hyperbolic groups, Trans. Amer. Math. Soc. 368 (2016), no. 2, 851–888. 10.1090/tran/6343Search in Google Scholar

[20] P. Papasoglu, An algorithm detecting hyperbolicity, Geometric and Computational Perspectives on Infinite Groups (Minneapolis 1994), DIMACS Ser. Discrete Math. Theoret. Comput. Sci. 25, American Mathematical Society, Providence (1996), 193–200. 10.1090/dimacs/025/10Search in Google Scholar

[21] J. Rotman, An Introduction to the Theory of Groups, 4th ed., Springer, New York, 1995. 10.1007/978-1-4612-4176-8Search in Google Scholar

[22] P. Schupp, A survey of small cancellation theory, Word Problems, North-Holland, Amsterdam (1973), 569–589. 10.1016/S0049-237X(08)71921-9Search in Google Scholar

[23] S. Simpson, A slick proof of the unsolvability of the word problem for finitely presented groups, preprint (2005), www.personal.psu.edu/t20/logic/seminar/050517.pdf. Search in Google Scholar

[24] D. Wise, Cubulating small cancellation groups, Geom. Funct. Anal. 14 (2004), no. 1, 150–214. 10.1007/s00039-004-0454-ySearch in Google Scholar

Received: 2017-09-13
Revised: 2018-05-04
Published Online: 2018-06-19
Published in Print: 2018-09-01

© 2018 Walter de Gruyter GmbH, Berlin/Boston

Downloaded on 21.9.2025 from https://www.degruyterbrill.com/document/doi/10.1515/jgth-2018-0022/html
Scroll to top button