This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

On the dynamics of endomorphisms of the direct product of two free groups

André Carvalho
Center for Mathematics and Applications (NOVA Math)
NOVA SST 2829–516 Caparica Portugal [email protected]
Abstract

We prove that Brinkmann’s problems are decidable for endomorphisms of Fn×FmF_{n}\times F_{m}: given (x,y),(z,w)Fn×Fm(x,y),(z,w)\in F_{n}\times F_{m} and ΦEnd(Fn×Fm)\Phi\in\text{End}(F_{n}\times F_{m}), it is decidable whether there is some kk\in\mathbb{N} such that (x,y)Φk=(z,w)(x,y)\Phi^{k}=(z,w) (or (x,y)Φk(z,w)(x,y)\Phi^{k}\sim(z,w)). We also prove decidability of a two-sided version of Brinkmann’s conjugacy problem for injective endomorphisms which, from the work of Logan, yields a solution to the conjugacy problem in ascending HNN-extensions of Fn×FmF_{n}\times F_{m}. Finally, we study the dynamics of automorphisms of Fn×FmF_{n}\times F_{m} at the infinity, proving that every point in a continuous extension of an automorphism to the completion is either periodic or wandering, implying that the dynamics is asymptotically periodic, as occurs in the free and free-abelian times free cases.

Introduction

The dynamical study of automorphisms and endomorphisms of groups has been a topic of interest for many in the past 30 years. In this paper, we will consider two different questions about the dynamics of endomorphisms of Fn×FmF_{n}\times F_{m}, with n,m2n,m\geq 2, the direct product of two nonabelian free groups. We remark that the study of endomorphisms of free-abelian times free groups and their dynamics is developed in [20, 8, 15].

The direct product of two free groups is known to be a good source of undecidability results, mostly due to Mihailova’s construction [27]. However, endomorphisms of these groups have been classified in [10], and positive decidability results were obtained: the Whitehead problem was proven to be decidable, fixed and periodic subgroups were proven to be computable (in case they are finitely generated) and in [12] it is proved that the triviality of the intersection of the subgroup fixed by a monomorphism with the subgroup fixed by an endomorphism of Fn×FmF_{n}\times F_{m} can be decided, showing that, in some sense, subgroups fixed by endomorphisms of Fn×FmF_{n}\times F_{m} are quite special among arbitrary subgroups. Some dynamical results about the behaviour around infinite fixed points are obtained in [10]

In the first part, we will solve Brinkamnn’s problems for endomorphisms of Fn×FmF_{n}\times F_{m}. Brinkmann’s (equality) problem on GG, BrP(GG), consists on deciding, on input two elements x,yGx,y\in G and an endomorphism ϕEnd(G)\phi\in\text{End}(G), whether yy belongs to the ϕ\phi-orbit of xx, i.e., whether there is some kk\in\mathbb{N} such that xϕk=yx\phi^{k}=y. Similarly, Brinkmann’s conjugacy problem on GG, BrCP(GG), consists on deciding whether there is some kk\in\mathbb{N} such that xϕkyx\phi^{k}\sim y. Brinkmann proved in [7] that these problems were decidable for automorphisms of the free group. This turned out to be particularly important in proving decidability of the conjugacy problem for free-by-cyclic groups in [3]. In [26], Logan solved several variations of this problem for general endomorphisms and used them to solve the conjugacy problem in ascending HNN-extensions of the free group, generalizing the work in [3]. In fact, it is proved in [15] that Logan’s results imply decidability of Brinkmann’s problems for endomorphisms (not necessarily injective) of the free group. Kannan and Lipton had already solved in [24] the problem of deciding whether, given an n×nn\times n matrix QQ of rational numbers and two vectors of rational numbers x,ynx,y\in\mathbb{Q}^{n}, there is a natural number ii\in\mathbb{N} such that xQi=yxQ^{i}=y, which is more general than Brinkmann’s Problem for free-abelian groups. In [11], the link between Brinkmann’s conjugacy problem and the conjugacy problem in cyclic extensions of the group was extended to generalized versions of the problems and in [13], the author studied a quantification of Brinkmann’s problem in the context of virtually free groups.

We prove that both problems are decidable for endomorphisms of Fn×FmF_{n}\times F_{m}.

Theorem 3.1.

There exists an algorithm taking as input integers n,m>1n,m>1, ΦEnd(Fn×Fm)\Phi\in\text{End}(F_{n}\times F_{m}) and two elements (x,y),(z,w)Fn×Fm(x,y),(z,w)\in F_{n}\times F_{m} that outputs kk\in\mathbb{N} such that (x,y)Φk=(z,w)(x,y)\Phi^{k}=(z,w) if such a kk exists and outputs NO otherwise.

Theorem 3.2.

There exists an algorithm taking as input integers n,m>1n,m>1, ΦEnd(Fn×Fm)\Phi\in\text{End}(F_{n}\times F_{m}) and two elements (x,y),(z,w)Fn×Fm(x,y),(z,w)\in F_{n}\times F_{m} that outputs kk\in\mathbb{N} such that (x,y)Φk(z,w)(x,y)\Phi^{k}\sim(z,w) if such a kk exists and outputs NO otherwise.

Logan proved the decidability of the conjugacy problem for ascending HNN-extensions of free groups by reducing this problem to a two-sided version of Brinkmann’s problem and the twisted conjugacy problem for injective endomorphisms of FnF_{n}. We remark that it is shown in [30] that almost all one relator groups with 33 or more generators are subgroups of ascending HNN-extensions of free groups and that the conjugacy problem is open for general one-relator groups. We solve this two-sided version of Brinkmann’s conjugacy problem as well as the twisted conjugacy problem for injective endomorphisms of Fn×FmF_{n}\times F_{m} and, using Logan’s method, this yields decidability of the conjugacy problem for ascending HNN-extensions of Fn×FmF_{n}\times F_{m}.

Corollary 4.4.

The conjugacy problem is solvable for ascending HNN-extensions of Fn×FmF_{n}\times F_{m}, for n,m>1n,m>1.

In the last part we study the dynamics at the infinity of automorphisms of Fn×FmF_{n}\times F_{m}.

For automorphisms of free groups, infinite fixed points were discussed by Bestvina and Handel in [2] and Gaboriau, Jaeger, Levitt and Lustig in [19]. The dynamics of free groups automorphisms is proved to be asymptotically periodic in [25] and the same was obtained for free-abelian times free groups in [8]. In [16], Cassaigne and Silva study the dynamics of infinite fixed points for monoids defined by special confluent rewriting systems (which contain free groups as a particular case). This was also achieved by Silva in [32] for virtually injective endomorphisms of virtually free groups and in [8] and [10] this was done for free-abelian times free groups and the direct product of two free groups, respectively.

We denote by FnF_{n} the free group of rank nn and its alphabet by X={x1,,xn}X=\{x_{1},\ldots,x_{n}\}. Given two words uu and vv on a free group, we write uvu\wedge v to denote the longest common prefix of uu and vv. The prefix metric on a free group is defined by

d(u,v)={2|uv| if uv0 otherwise.d(u,v)=\begin{cases}2^{-|u\wedge v|}\text{ if $u\neq v$}\\ 0\text{ otherwise}\end{cases}.

The prefix metric on a free group is in fact an ultrametric and its completion (F^n,d^)(\widehat{F}_{n},\widehat{d}) is a compact space which can be described as the set of all finite and infinite reduced words on the alphabet XX1X\cup X^{-1}. We will denote by Fn\partial F_{n} the set consisting of only the infinite words and call it the boundary of FnF_{n}.

We consider Fn×FmF_{n}\times F_{m} endowed with the product metric given by taking the prefix metric in each factor. This is also an ultrametric and Fn×Fm^\widehat{F_{n}\times F_{m}} is homeomorphic to Fn^×Fm^\widehat{F_{n}}\times\widehat{F_{m}} by uniqueness of the completion (Theorem 24.4 in [34]).

When seen as a CAT(0) cube complex, or alternatively, as a median algebra, this coincides with the Roller compactification (see [29, 18, 22, 23, 6]). Indeed, the Roller boundary and the Gromov boundary coincide in the free group and the behaviour of the Roller compactification when taking direct products is the same as the one of the completion of metric spaces, i.e., denoting by X¯\bar{X} the Roller compactification of XX, we have that

X¯=i=1mX¯1××X¯m.\bar{X}=\bigcup_{i=1}^{m}\bar{X}_{1}\times\ldots\times\bar{X}_{m}.

It is well known by a general topology result [21, Section XIV.6] that every uniformly continuous mapping φ\varphi between metric spaces admits a unique continuous extension φ^\widehat{\varphi} to the completion. The converse is obviously true in this case: if a mapping between metric spaces admits a continuous extension to the completion, since the completion is compact, then the extension must be uniformly continuous, and so does the restriction to the original mapping. Uniformly continuous endomorphisms with respect to this metric are described in [10] and it turns out that every automorphism of Fn×FmF_{n}\times F_{m} is indeed uniformly continuous. Characterizing and studying some properties of uniformly continuous endomorphisms has been done before for other classes of groups (see for example [17, 31, 32, 1, 8]).

An interesting property of this metric (and so, of this boundary) is that the uniformly continuous endomorphisms of Fn×FmF_{n}\times F_{m} for this metric dd are precisely the coarse-median preserving endomorphisms for the product coarse median obtained by taking the median operator induced by the coarse-median operators given by hyperbolicity of FnF_{n} and FmF_{m}:

Corollary 5.2.

Coarse-median preserving endomorphisms of Fn×FmF_{n}\times F_{m} are precisely the uniformly continuous ones.

Coarse-median preservation turns out to be a useful tool to obtain interesting properties of automorphisms (see [23]), including finiteness results on the fixed subgroup of an automorphism. We remark that uniformly continuous preserving endomorphisms also coincide with the uniformly continuous ones for free-abelian times free groups.

In the study of dynamical systems, the notion of ω\omega-limit set plays a crucial role. Given a metric space XX, a continuous function f:XXf:X\to X, and a point xXx\in X, the ω\omega-limit set ω(x,f)\omega(x,f) of xx consists of the accumulation points of the sequence of points in the orbit of xx. Understanding the ω\omega-limits gives us a grasp on the behaviour of the system in the long term. If the space XX is compact, then ω\omega-limit sets are nonempty, compact and ff-invariant. Informally, a point is said to be wandering if it has some neighborhood such that, from some point on, its points leave the neighborhood and don’t come back. Obviously, a wandering point cannot belong to an ω\omega-limit set.

In [25], the authors proved that in the case where ff is the extension of a free group automorphism to the completion, then, for every point xF^nx\in\widehat{F}_{n}, ω(x,f)\omega(x,f) is a periodic orbit. In [8], something stronger was proven for automorphisms of m×Fn\mathbb{Z}^{m}\times F_{n}: for a uniformly continuous automorphism, every point in the completion must be either periodic or wandering. This shows that non-periodic points, when iterated long enough wander away from some neighborhood carrying the neighborhood with them. In particular, since ω\omega-limit sets are nonempty, they must be periodic orbits.

In this paper, we prove the same result for automorphisms of Fn×FmF_{n}\times F_{m}:

Theorem 5.3.

Let ΦAut(Fn×Fm)\Phi\in\text{Aut}(F_{n}\times F_{m}), dd be the product metric given by taking the prefix metric in each factor and let Φ^\widehat{\Phi} be its continuous extension to the completion with respect to dd. Then every point in Fn×Fm^\widehat{F_{n}\times F_{m}} is either wandering or periodic.

It follows in particular that the dynamics of automorphisms of Fn×FmF_{n}\times F_{m} is asymptotically periodic.

Preliminaries

The purpose of this section is to introduce the classification of endomorphisms of Fn×FmF_{n}\times F_{m} obtained in [10], notation and some known results on the dynamics of endomorphisms of free and free-abelian groups that will be used later.

Endomorphisms of Fn×FmF_{n}\times F_{m}

For (i,j)[n]×[m](i,j)\in[n]\times[m], we define λi:Fn\lambda_{i}:F_{n}\to\mathbb{Z} as the endomorphism given by akδika_{k}\mapsto\delta_{ik} and τj:Fm\tau_{j}:F_{m}\to\mathbb{Z} given by bkδjkb_{k}\mapsto\delta_{jk}, where δij\delta_{ij} is the Kronecker symbol.

For xFn,yFmx\in F_{n},y\in F_{m} and integers pi,qi,rj,sjp_{i},q_{i},r_{j},s_{j}\in\mathbb{Z}, we denote i[n]λi(x)pi\sum\limits_{i\in[n]}\lambda_{i}(x)p_{i} by xPx^{P}; j[m]τj(y)rj\sum\limits_{j\in[m]}\tau_{j}(y)r_{j} by yRy^{R}; i[n]λi(x)qi\sum\limits_{i\in[n]}\lambda_{i}(x)q_{i} by xQx^{Q} and j[n]τj(y)si\sum\limits_{j\in[n]}\tau_{j}(y)s_{i} by ySy^{S}. We also define P={pii[n]}P=\{p_{i}\in\mathbb{Z}\mid i\in[n]\}, Q={qii[n]}Q=\{q_{i}\in\mathbb{Z}\mid i\in[n]\}, R={rjj[m]}R=\{r_{j}\in\mathbb{Z}\mid j\in[m]\} and S={sjj[m]}S=\{s_{j}\in\mathbb{Z}\mid j\in[m]\}. We will keep this notation throughout the paper.

In [10], the author classified endomorphisms of the direct product of two finitely generated free groups Fn×FmF_{n}\times F_{m}, with m,n>1m,n>1 in seven different types:

  1. (I)

    (x,y)(uxP+yR,vxQ+yS)(x,y)\mapsto\left(u^{x^{P}+y^{R}},v^{x^{Q}+y^{S}}\right), for some 1uFn1\neq u\in F_{n}, 1vFm1\neq v\in F_{m} and integers pi,qi,rj,sjp_{i},q_{i},r_{j},s_{j}\in\mathbb{Z} for (i,j)[n]×[m](i,j)\in[n]\times[m], such that P,Q,R,S{0}P,Q,R,S\neq\{0\}.

  2. (II)

    (x,y)(yϕ,vxQ+yS)(x,y)\mapsto\left(y\phi,v^{x^{Q}+y^{S}}\right), for some nontrivial homomorphism ϕ:FmFn\phi:F_{m}\to F_{n}, 1vFm1\neq v\in F_{m} and integers qi,sjq_{i},s_{j}\in\mathbb{Z} for (i,j)[n]×[m](i,j)\in[n]\times[m], such that Q,S{0}Q,S\neq\{0\}.

  3. (III)

    (x,y)(uxP+yR,yϕ),(x,y)\mapsto\left(u^{x^{P}+y^{R}},y\phi\right), for some nontrivial endomorphism ϕEnd(Fm)\phi\in\text{End}(F_{m}), 1uFn1\neq u\in F_{n}, and integers pi,rjp_{i},r_{j}\in\mathbb{Z} for (i,j)[n]×[m](i,j)\in[n]\times[m], such that P,R{0}P,R\neq\{0\}.

  4. (IV)

    (x,y)(yϕ,yψ)(x,y)\mapsto(y\phi,y\psi), for some nontrivial homomorphism ϕ:FmFn\phi:F_{m}\to F_{n} and nontrivial endomorphism ψEnd(Fm)\psi\in\text{End}(F_{m}).

  5. (V)

    (x,y)(1,vxQ+yS)(x,y)\mapsto\left(1,v^{x^{Q}+y^{S}}\right), for some 1vFm1\neq v\in F_{m}, and integers qi,sjq_{i},s_{j}\in\mathbb{Z} for (i,j)[n]×[m](i,j)\in[n]\times[m], such that Q,S{0}Q,S\neq\{0\}.

  6. (VI)

    (x,y)(xϕ,yψ)(x,y)\mapsto(x\phi,y\psi), for some endomorphisms ϕEnd(Fn)\phi\in\text{End}(F_{n}), ψEnd(Fm)\psi\in\text{End}(F_{m}).

  7. (VII)

    (x,y)(yψ,xϕ)(x,y)\mapsto(y\psi,x\phi), for homomorphisms ϕ:FnFm\phi:F_{n}\to F_{m} and ψ:FmFn\psi:F_{m}\to F_{n}.

From [10, Proposition 3.2], injective endomorphisms correspond to endomorphisms of type VI or VII such that the component mappings ϕ\phi and ψ\psi are injective. We denote by Mon(G)\mbox{\rm Mon}(G) the monoid of monomorphisms of a group GG. In [10, Corollary 3.3], it is shown that automorphisms of Fn×FmF_{n}\times F_{m} are the type VI endomorphisms with bijective component endomorphisms ϕ\phi and ψ\psi and if n=mn=m, there are also automorphisms of type VII, given by the type VII endomorphisms with bijective component homomorphisms ϕ\phi and ψ\psi. Hence, groups of the form Fn×FnF_{n}\times F_{n} have more automorphisms than groups of the form Fn×FmF_{n}\times F_{m} with nmn\neq m. We will usually denote endomorphisms (and homomorphisms) of free groups by ϕ,ψ\phi,\psi and endomorphisms of Fn×FmF_{n}\times F_{m} by Φ,Ψ\Phi,\Psi.

Given an endomorphism by the image of the generators of Fn×FmF_{n}\times F_{m}, its type is decidable. Indeed, consider an endomorphism φ:Fn×FmFn×Fm\varphi:F_{n}\times F_{m}\to F_{n}\times F_{m} defined by (ai,1)(xi,yi)(a_{i},1)\mapsto(x_{i},y_{i}) and (1,bj)(zj,wj)(1,b_{j})\mapsto(z_{j},w_{j}) for i[n]i\in[n] and j[m].j\in[m]. We define X={xii[n]}X=\{x_{i}\mid i\in[n]\}, Y={yii[n]}Y=\{y_{i}\mid i\in[n]\}, Z={zjj[m]}Z=\{z_{j}\mid j\in[m]\} and W={wjj[m]}W=\{w_{j}\mid j\in[m]\}. We say that these sets are trivial if they are singletons containing only the empty word and nontrivial otherwise. As seen in [10], matching the numbering above the endomorphisms are classified as follows:

  1. (I)

    All sets X,Y,ZX,Y,Z and WW are nontrivial

  2. (II)

    XX is the only trivial set

  3. (III)

    YY is the only trivial set

  4. (IV)

    XX and YY are the only trivial sets

  5. (V)

    XX and ZZ are the only trivial sets

  6. (VI)

    YY and ZZ are trivial sets

  7. (VII)

    XX and WW are trivial sets

Therefore, when we take an endomorphism as input (meaning that we are given images of the generators), we will often assume that its type is known.

Dynamics of endomorphisms

2.2.1 Orbits

We now review some concepts that will be used when studying the dynamics of endomorphisms.

Brinkmann’s problems were proven to be decidable for automorphisms of the free group by Brinkmann in [7]. In [26], Logan extended these resuls to injective endomorphisms, and using a result from [26], it is shown in [15] that these problems are decidable for arbitrary endomorphisms of the free group. Kannan and Lipton proved (something more general than) the decidability of Brinkmann’s problem for free-abelian groups.

Theorem 2.1.

Given a matrix Mn×n()M\in\mathcal{M}_{n\times n}(\mathbb{Q}) and two vectors x,ynx,y\in\mathbb{Q}^{n} it is decidable whether there exists some kk\in\mathbb{N} such that xMk=yxM^{k}=y.

Remark 2.2.

We will often use the above theorem for affine transformations. It is seen in [15] that this is not a problem, as an affine transformation can be seen as a restriction of a linear one.

Let GG be a group, x,yGx,y\in G and ϕEnd(G)\phi\in\text{End}(G). Then, the set of ϕ\phi-logarithms (resp. ϕ~\widetilde{\phi}-logarithms) of yy in base xx is ϕ-logx(y)={k0xϕk=y}\phi\text{-}\log_{x}(y)=\{k\geq 0\mid x\phi^{k}=y\} (resp. ϕ~-logx(y)={k0xϕky}\tilde{\phi}\text{-}\log_{x}(y)=\{k\geq 0\mid x\phi^{k}\sim y\}). An element xGx\in G for which there is some k>0k>0 such that xϕk=xx\phi^{k}=x (resp. xϕkxx\phi^{k}\sim x) is said to be ϕ\phi-periodic (resp. ϕ~\tilde{\phi}-periodic). The smallest kk satisfying the condition is called the ϕ\phi-period (resp. ϕ~\tilde{\phi}-period).

Given a finite orbit Orbϕ(x)\text{Orb}_{\phi}(x), we say that Orbϕ(x)Per(ϕ)\text{Orb}_{\phi}(x)\cap\text{Per}(\phi) is the periodic part of the orbit and Orbϕ(x)Per(ϕ)\text{Orb}_{\phi}(x)\setminus\text{Per}(\phi) is the straight part of the orbit.

{\cdots}x{x}xϕ{x\phi}xϕ2{x\phi^{2}}{\cdots}xϕr{x\phi^{r}}xϕr+2{x\phi^{r+2}}xϕr+1{x\phi^{r+1}}
Figure 1: A finite orbit

In Figure 1, the straight part of the orbit corresponds to {x,xϕ,,xϕr1}\{x,x\phi,\ldots,x\phi^{r-1}\} and the periodic part of the orbit to {xϕkkr}={xϕr,,xϕr+p1}\{x\phi^{k}\mid k\geq r\}=\{x\phi^{r},\ldots,x\phi^{r+p-1}\}, where pp is the period of xϕrx\phi^{r}. We will consider the same notions up to conjugacy and the notation should be clear from context: for example, a finite ϕ~\tilde{\phi}-orbit is one containing only finitely many conjugacy classes. Naturally, ϕ\phi-logarithms and ϕ~\tilde{\phi}-logarithms have the form p+qp+q\mathbb{N} for natural numbers pp and qq and are computable, as long as Brinkmann’s problems are decidable.

2.2.2 Dynamics at the infinity

When we consider a hyperbolic group GG endowed with a visual metric dd and take its completion, we obtained the Gromov completion of GG. As said above, the endomorphisms of GG admitting a continuous extension to the completion are precisely the uniformly continuous ones (with respect to dd). It is proved in [9] that uniformly continuous endomorphisms of a hyperbolic group satisfy a Hölder condition. An endomorphism ϕ\phi satisfies a Hölder condition with constants (K,r)(K,r) if

d(xϕ,yϕ)Kd(x,y)r.d(x\phi,y\phi)\leq Kd(x,y)^{r}.

If ϕ\phi satisfies a Hölder condition, then so does its continuous extension to the completion, by [1].

We now present some standard dynamical definitions that will be useful to the classification of infinite points. A point xGx\in G is said to be a φ\varphi-wandering point if there is a neighbourhood UU of xx and a positive integer NN such that for all n>Nn>N, we have that UφnU=.U\varphi^{n}\cap U=\emptyset. A point xGx\in G is said to be a φ\varphi-recurrent point if, for every neighbourhood UU of xx, there exists n>0n>0 such that xφnUx\varphi^{n}\in U. Now, let ff be a homeomorphism of a compact space KK. Given yKy\in K, the ω\omega-limit set ω(y,f)\omega(y,f), or simply ω(y)\omega(y), is the set of limit points of the sequence fn(y)f^{n}(y) as n+n\to+\infty. We say that the dynamics is asymptotically periodic if every ω\omega-limit set is a periodic orbit (see [25]). A point is recurrent if it belongs to its own ω\omega-limit set.

We will study these concepts for Φ^\widehat{\Phi}, the continuous extension of an endomorphism to the completion Fn×Fm^\widehat{F_{n}\times F_{m}} when the group is endowed with the product metric given by taking the prefix metric in each direct factor. As noted above, this is the Roller completion of the group.

We have that

periodicrecurrentnonwandering.\displaystyle\text{periodic}\implies\text{recurrent}\implies\text{nonwandering}. (1)

It is proved in [8, Proposition 5.10] that, if ϕ\phi is a uniformly continuous endomorphism of a free group, every nonwandering point of ϕ^\widehat{\phi} is recurrent. In fact, the proof only uses the fact that ϕ^\widehat{\phi} satisfies a Hölder condition. So the following more general statement, which can be found in [14, Proposition 8.1.29], holds:

Proposition 2.3.

Let (X,d)(X,d) be a metric space and ϕ:(X,d)(X,d)\phi:(X,d)\to(X,d) be a Hölder mapping. Then every nonwandering point is recurrent.

If the dynamics is asymptotically periodic, since recurrent points belong to their own ω\omega-limit set and these are periodic orbits, then recurrent points are periodic. So if the dynamics is asymptotically periodic and the mapping is Hölder, all implications in (1) are in fact equivalences. Conversely, if all implications in (1) are equivalences, since every point in a ω\omega-limit set must be nonwandering, it must be periodic and we have that the dynamics is asymptotically periodic. This dichotomy wandering/periodic holds for automorphisms of free-abelian times free groups (see [8]) and some (defined in a precise way) uniformly continuous endomorphisms of free-abelian times free groups. It is not known to the author if this holds for general uniformly continuous endomorphisms (injective and not surjective) of free groups.

Brinkmann’s Problems

The goal of this section is to prove the following theorems.

Theorem 3.1.

There exists an algorithm taking as input integers n,m>1n,m>1, ΦEnd(Fn×Fm)\Phi\in\text{End}(F_{n}\times F_{m}) and two elements (x,y),(z,w)Fn×Fm(x,y),(z,w)\in F_{n}\times F_{m} that outputs kk\in\mathbb{N} such that (x,y)Φk=(z,w)(x,y)\Phi^{k}=(z,w) if such a kk exists and outputs NO otherwise.

Theorem 3.2.

There exists an algorithm taking as input integers n,m>1n,m>1, ΦEnd(Fn×Fm)\Phi\in\text{End}(F_{n}\times F_{m}) and two elements (x,y),(z,w)Fn×Fm(x,y),(z,w)\in F_{n}\times F_{m} that outputs kk\in\mathbb{N} such that (x,y)Φk(z,w)(x,y)\Phi^{k}\sim(z,w) if such a kk exists and outputs NO otherwise.

We now present two technical lemmas that will be useful later.

Lemma 3.3.

There exists an algorithm that, on input two reduced words u,vFnu,v\in F_{n}, decides whether uu is conjugate to some power of vv and, in case it is, outputs the unique value of kk\in\mathbb{N} such that uvku\sim v^{k}.

Proof. Two words uu and vv are conjugate if and only if the cyclic reduced core of uu, u~\tilde{u} is a cyclic permutation of the cyclic reduced core of vv, v~\tilde{v}. In particular, their cyclically reduced cores must have the same length. So, we compute u~\tilde{u}, v~\tilde{v} and check if k=|u~||v~|k=\frac{|\tilde{u}|}{|\tilde{v}|} is an integer. If it is not, then there is no power kk such that uvku\sim v^{k} and if it is, then it is our only candidate. Hence, it only remains to check whether uvku\sim v^{k} or not. ∎

The equality version of this lemma can be seen to hold in the same way.

Lemma 3.4.

There exists an algorithm that, on input two reduced words u.,vFnu.,v\in F_{n}, decides whether uu is equal to to some power of vv in FnF_{n} and, in case it is, outputs the unique value of kk\in\mathbb{N} such that u=vku=v^{k}.

Since the word problem and the conjugacy problem are decidable in Fn×FmF_{n}\times F_{m}, we will prove that we can decide the existence of a positive kk and that is equivalent to deciding if there is a nonnegative value of kk such that (x,y)Φk=(z,w)(x,y)\Phi^{k}=(z,w) (or (x,y)Φk(z,w)(x,y)\Phi^{k}\sim(z,w), in the conjugacy case).

Type I

In this case, we have that (x,y)Φ=(uxP+yR,vxQ+yS)(x,y)\Phi=(u^{x^{P}+y^{R}},v^{x^{Q}+y^{S}}), for some words u,v1u,v\neq 1. It is shown in [10, Subsection 5.1] that, defining sequences in \mathbb{Z} by

{a1(x,y)=xP+yRb1(x,y)=xQ+ySan+1(x,y)=an(x,y)uP+bn(x,y)vRbn+1(x,y)=an(x,y)uQ+bn(x,y)vS,\displaystyle\begin{cases}a_{1}(x,y)=x^{P}+y^{R}\\ b_{1}(x,y)=x^{Q}+y^{S}\\ a_{n+1}(x,y)=a_{n}(x,y)u^{P}+b_{n}(x,y)v^{R}\\ b_{n+1}(x,y)=a_{n}(x,y)u^{Q}+b_{n}(x,y)v^{S}\end{cases},

we have that, for every k>0k>0, (x,y)Φk=(uak(x,y),vbk(x,y))(x,y)\Phi^{k}=(u^{a_{k}(x,y)},v^{b_{k}(x,y)}). We want to decide, given (x,y),(z,w)Fn×Fm(x,y),(z,w)\in F_{n}\times F_{m}, whether there is some kk\in\mathbb{N} such that (x,y)Φk=(z,w)(x,y)\Phi^{k}=(z,w). Using Lemma 3.4 to decide if zz is a power of uu and ww is a power of vv. If in any of the cases the answer is no, then we will not have a positive solution to our problem; if both answer yes, we compute exponents (a,b)2(a,b)\in\mathbb{N}^{2} such that (z,w)=(ua,bv)(z,w)=(u^{a},b^{v}).

So, for all nn\in\mathbb{N},

[anbn]=[uPvRuQvS]n1[xP+yRxQ+yS],\begin{bmatrix}a_{n}\\ b_{n}\end{bmatrix}=\begin{bmatrix}u^{P}&v^{R}\\ u^{Q}&v^{S}\end{bmatrix}^{n-1}\begin{bmatrix}x^{P}+y^{R}\\ x^{Q}+y^{S}\end{bmatrix},

hence, putting MΦ=[uPvRuQvS]M_{\Phi}=\begin{bmatrix}u^{P}&v^{R}\\ u^{Q}&v^{S}\end{bmatrix}, the problem can now be translated as the problem of deciding whether there is some kk\in\mathbb{N} such that

MΦk[xP+yRxQ+yS]=[ab],M_{\Phi}^{k}\begin{bmatrix}x^{P}+y^{R}\\ x^{Q}+y^{S}\end{bmatrix}=\begin{bmatrix}a\\ b\end{bmatrix},

which can be done by [24].

In the conjugacy case, we start by checking if k=0k=0 is a solution to our problem by solving the conjugacy problem. If not, using Lemma 3.3, we check if zz is conjugate to a power of uu and ww is conjugate to a power of vv. If in any of the cases the answer is no, then we will not have a positive solution to our problem; if both answer yes, we compute the unique exponents (a,b)2(a,b)\in\mathbb{N}^{2} such that (z,w)(ua,bv)(z,w)\sim(u^{a},b^{v}). Then we simply check if there is some kk such that (x,y)Φk=(ua,vb)(x,y)\Phi^{k}=(u^{a},v^{b}) using the equality algorithm above.

Type II

In this case, we have that (x,y)Φ=(yϕ,vxQ+yS)(x,y)\Phi=(y\phi,v^{x^{Q}+y^{S}}), for some word v1v\neq 1 and homomorphism ϕ:FmFn\phi:F_{m}\to F_{n}. It is shown in [10, Subsection 5.2] that, defining a sequence in \mathbb{Z} by

{a1(x,y)=xQ+ySa2(x,y)=a1(x,y)vS+yϕQan(x,y)=an1(x,y)vS+an2(x,y)(vϕ)Q\displaystyle\begin{cases}a_{1}(x,y)=x^{Q}+y^{S}\\ a_{2}(x,y)=a_{1}(x,y)v^{S}+y\phi^{Q}\\ a_{n}(x,y)=a_{n-1}(x,y)v^{S}+a_{n-2}(x,y)(v\phi)^{Q}\\ \end{cases}

we have that, for every k>1k>1, (x,y)Φk=(vak1(x,y)ϕ,vak(x,y))(x,y)\Phi^{k}=(v^{a_{k-1}(x,y)}\phi,v^{a_{k}(x,y)}). We start by checking if k=1k=1 is a solution. Then, by Lemma 3.4 we test if zz is a power of vϕv\phi and if ww is a power of vv. If one of them is not, then there is no solution kk. If both are, we compute a,ba,b\in\mathbb{N} such that (z,w)=(vaϕ,vb)(z,w)=(v^{a}\phi,v^{b}). Our problem now becomes finding k>0k>0 such that [ba]=[ak+1(x,y)ak(x,y)]\begin{bmatrix}b&a\end{bmatrix}=\begin{bmatrix}a_{k+1}(x,y)&a_{k}(x,y)\end{bmatrix}. Similarly to the type I case, we use Kannan-Lipton’s algorithm to decide the existence of such a kk. Consider the matrix MΦ=[vS(vϕ)Q10]M_{\Phi}=\begin{bmatrix}v^{S}&(v\phi)^{Q}\\ 1&0\end{bmatrix}. Clearly, for k>0k>0,

MΦ[ak+1(x,y)ak(x,y)]=[ak+2(x,y)ak+1(x,y)],M_{\Phi}\begin{bmatrix}a_{k+1}(x,y)\\ a_{k}(x,y)\end{bmatrix}=\begin{bmatrix}a_{k+2}(x,y)\\ a_{k+1}(x,y)\end{bmatrix},

and so

MΦk[a2(x,y)a1(x,y)]=[ak+2(x,y)ak+1(x,y)].M_{\Phi}^{k}\begin{bmatrix}a_{2}(x,y)\\ a_{1}(x,y)\end{bmatrix}=\begin{bmatrix}a_{k+2}(x,y)\\ a_{k+1}(x,y)\end{bmatrix}.

By [24], we can decide if there is some kk such that

[ak+2(x,y)ak+1(x,y)]=MΦk[a2(x,y)a1(x,y)]=[ba],\begin{bmatrix}a_{k+2}(x,y)\\ a_{k+1}(x,y)\end{bmatrix}=M_{\Phi}^{k}\begin{bmatrix}a_{2}(x,y)\\ a_{1}(x,y)\end{bmatrix}=\begin{bmatrix}b\\ a\end{bmatrix},

and we are done.

The variation up to conjugacy is similar to the previous case: we first decide if zz is conjugate to some power of vϕv\phi and ww is conjugate to some power of vv and then apply the algorithm for equality.

Type III

We have that (x,y)(uxP+yR,yϕ)(x,y)\mapsto\left(u^{x^{P}+y^{R}},y\phi\right), for some word u1u\neq 1 and endomorphism ϕEnd(Fm)\phi\in\text{End}(F_{m}). It is seen in [10, Subsection 5.3] that

(ua,y)Φk=(ua(uP)k+t=0k1(yϕt)R(uP)kt1,yϕk).(u^{a},y)\Phi^{k}=\left(u^{a(u^{P})^{k}+\sum\limits_{t=0}^{k-1}(y\phi^{t})^{R}(u^{P})^{k-t-1}},y\phi^{k}\right).

We start by solving BrP(FnF_{n}) to compute p0,p1p_{0},p_{1}\in\mathbb{N} such that ϕ-logy(w)=p0+p1\phi\text{-}\log_{y}(w)=p_{0}+p_{1}\mathbb{N}. If p1=0p_{1}=0, then p0p_{0} is our only candidate, so the only thing to check is if (x,y)Φp0=(z,w)(x,y)\Phi^{p_{0}}=(z,w). If p10p_{1}\neq 0, ww is a ϕ\phi-periodic point with period p1p_{1}. We check if zz is a power of uu. If it is not, then we are done, since there is no possible kk; if it is, we compute aa\in\mathbb{N} such that z=uaz=u^{a}.

Now consider the following affine transformation of \mathbb{Z}

Θ:\displaystyle\Theta\colon\mathbb{Z} \displaystyle\longrightarrow\mathbb{Z}
c\displaystyle c c(uP)p1+t=0p11(wϕt)R(uP)p1t1,\displaystyle\longmapsto c(u^{P})^{p_{1}}+\sum_{t=0}^{p_{1}-1}(w\phi^{t})^{R}(u^{P})^{p_{1}-t-1},

and denote by (an)n(a_{n})_{n} the sequence such that (x,y)Φn=(uan,yϕn)(x,y)\Phi^{n}=(u^{a_{n}},y\phi^{n}). Clearly, given kk\in\mathbb{N}, we can compute aka_{k}. We claim that, for kp0+p1k\in p_{0}+p_{1}\mathbb{N}, writing k=p0+p1rk=p_{0}+p_{1}r, we have that

(x,y)Φk=(uap0Θr,yϕk).\displaystyle(x,y)\Phi^{k}=(u^{a_{p_{0}}\Theta^{r}},y\phi^{k}). (2)

We prove it by induction over rr. Clearly, if r=0r=0, then (x,y)Φk=(x,y)Φp0=(uap0,yϕk).(x,y)\Phi^{k}=(x,y)\Phi^{p_{0}}=(u^{a_{p_{0}}},y\phi^{k}). Now assume that the claim holds for rr up to some n0n\geq 0. Notice that w=yϕp0w=y\phi^{p_{0}} is periodic with period p1p_{1}, and so, for all n,tn,t\in\mathbb{N}, yϕp0+np1+t=wϕt.y\phi^{p_{0}+np_{1}+t}=w\phi^{t}. Hence,

(x,y)Φp0+(n+1)p1\displaystyle(x,y)\Phi^{p_{0}+(n+1)p_{1}} =(x,y)Φp0+np1Φp1\displaystyle=(x,y)\Phi^{p_{0}+np_{1}}\Phi^{p_{1}}
=(uap0Θn,yϕp0+np1)Φp1\displaystyle=(u^{a_{p_{0}}\Theta^{n}},y\phi^{p_{0}+np_{1}})\Phi^{p_{1}}
=(u(ap0Θn)(uP)p1+t=0p11(yϕp0+np1+t)R(uP)p1t1,yϕp0+(n+1)p1)\displaystyle=\left(u^{(a_{p_{0}}\Theta^{n})(u^{P})^{p_{1}}+\sum\limits_{t=0}^{p_{1}-1}(y\phi^{p_{0}+np_{1}+t})^{R}(u^{P})^{p_{1}-t-1}},y\phi^{p_{0}+(n+1)p_{1}}\right)
=(u(ap0Θn)(uP)p1+t=0p11(wϕt)R(uP)p1t1,yϕp0+(n+1)p1)\displaystyle=\left(u^{(a_{p_{0}}\Theta^{n})(u^{P})^{p_{1}}+\sum\limits_{t=0}^{p_{1}-1}(w\phi^{t})^{R}(u^{P})^{p_{1}-t-1}},y\phi^{p_{0}+(n+1)p_{1}}\right)
=(uap0Θn+1,yϕk).\displaystyle=\left(u^{a_{p_{0}}\Theta^{n+1}},y\phi^{k}\right).

Since our only candidate solutions are of the form p0+p1p_{0}+p_{1}\mathbb{N}, we can proceed as follows: we check manually the existence of a solution for kk up to p0p_{0}, using the word problem; then, if we obtain no positive answer, we verify the existence of some rr\in\mathbb{N} such that ap0Θr=aa_{p_{0}}\Theta^{r}=a, which can be done by Remark 2.2.

When considering the conjugation variant, we proceed analogously. Using BrCP(FnF_{n}), we compute p0,p1p_{0},p_{1}\in\mathbb{N} such that ϕ~-logy(w)=p0+p1\tilde{\phi}\text{-}\log_{y}(w)=p_{0}+p_{1}\mathbb{N}. We then check if zz is conjugate to some power of uu and if so, compute aa such that uazu^{a}\sim z. Using the same argument as above, we check if there is some rr such that the first component (x,y)Φp0+rp1(x,y)\Phi^{p_{0}+rp_{1}} is uau^{a}. In fact, even though ww is no longer a periodic point, it is enough that ww is ϕ~\tilde{\phi}-periodic since, for words u,vFmu,v\in F_{m} such that uvu\sim v, i.e., for which there is some zFmz\in F_{m} such that u=z1vzu=z^{-1}vz, we have that

uR=j[m]τj(u)rj=j[m]τj(z1vz)rj=j[m]τj(v)rj=vR.u^{R}=\sum\limits_{j\in[m]}\tau_{j}(u)r_{j}=\sum\limits_{j\in[m]}\tau_{j}(z^{-1}vz)r_{j}=\sum\limits_{j\in[m]}\tau_{j}(v)r_{j}=v^{R}.

This means that for all n,tn,t\in\mathbb{N}, (yϕp0+np1+t)R=(wϕt)R(y\phi^{p_{0}+np_{1}+t})^{R}=(w\phi^{t})^{R} and the same reduction done above can be done here and the claim (2) holds in this case too. Hence, we reduce our problem to finding some rr\in\mathbb{N} such that ap0Θr=aa_{p_{0}}\Theta^{r}=a, which can be done by Remark 2.2.

Type IV

We have that (x,y)(yϕ,yψ)(x,y)\mapsto(y\phi,y\psi), for some homomorphism ϕ:FmFn\phi:F_{m}\to F_{n} and endomorphism ψEnd(Fm)\psi\in\text{End}(F_{m}), and so, as seen in [10, Subsection 5.4], (x,y)Φk=(yψk1ϕ,yψk),(x,y)\Phi^{k}=(y\psi^{k-1}\phi,y\psi^{k}), for k>0k>0. Using BrP(FnF_{n}), we start by computing ψ-logy(w)\psi\text{-}\log_{y}(w). If it is empty, then there are no solutions kk. If not, ψ-logy(w)=p0+p1\psi\text{-}\log_{y}(w)=p_{0}+p_{1}\mathbb{N} for some computable p0,p1p_{0},p_{1}\in\mathbb{N}. We check if k=p0k=p_{0} is a solution or not. We have that w=yψp0w=y\psi^{p_{0}} is a ψ\psi-periodic point with period p1p_{1}. For k=p0+p1rk=p_{0}+p_{1}r, with r1r\geq 1, we have that (x,y)Φk=(yψp0+p1r1ϕ,w)(x,y)\Phi^{k}=(y\psi^{p_{0}+p_{1}r-1}\phi,w). But for all r1r\geq 1, yψp0+p1r1=yψp0+p11y\psi^{p_{0}+p_{1}r-1}=y\psi^{p_{0}+p_{1}-1}. So we simply check if z=yψp0+p11ϕz=y\psi^{p_{0}+p_{1}-1}\phi. If it is, then p0+p1>0p_{0}+p_{1}\mathbb{N}_{>0} is the set of solutions; if not, then there are no solutions to our problem.

We can handle the conjugacy version of the problem in the same way, computing ψ~-logy(w)\tilde{\psi}\text{-}\log_{y}(w) instead of ψ-logy(w)\psi\text{-}\log_{y}(w). In this case, we have that, for all r1r\geq 1, yψp0+p1r1yψp0+p11y\psi^{p_{0}+p_{1}r-1}\sim y\psi^{p_{0}+p_{1}-1} and so yψp0+p1r1ϕyψp0+p11ϕy\psi^{p_{0}+p_{1}r-1}\phi\sim y\psi^{p_{0}+p_{1}-1}\phi and the set of solutions is p0+p1>0p_{0}+p_{1}\mathbb{N}_{>0} if zyψp0+p11ϕz\sim y\psi^{p_{0}+p_{1}-1}\phi and empty otherwise.

Type V

We have that (x,y)(1,vxQ+yS)(x,y)\mapsto\left(1,v^{x^{Q}+y^{S}}\right), for some 1vFm1\neq v\in F_{m}, for some word vFmv\in F_{m}, and so, as seen in [10, Subsection 5.5], (x,y)φk=(1,v(xQ+yS)(vS)k1)(x,y)\varphi^{k}=\left(1,v^{(x^{Q}+y^{S})(v^{S})^{k-1}}\right), for k>0k>0. So we start by checking if z=1z=1 and if ww is a power of vv. If any of these does not hold, then there is no positive solution kk. If both hold, we compute aa\in\mathbb{Z} such that w=vaw=v^{a}. Clearly, kk is a solution if and only if (vS)k1=axQ+yS(v^{S})^{k-1}=\frac{a}{x^{Q}+y^{S}}, and its existence is decidable. The conjugacy problem is analogous: we compute the unique aa^{\prime}\in\mathbb{Z} such that wvaw\sim v^{a^{\prime}} and check if there is some kk such that (vS)k1=axQ+yS(v^{S})^{k-1}=\frac{a^{\prime}}{x^{Q}+y^{S}}.

Type VI

We have that (x,y)(xϕ,yψ)(x,y)\mapsto(x\phi,y\psi), for some endomorphisms ϕEnd(Fn)\phi\in\text{End}(F_{n}), ψEnd(Fm)\psi\in\text{End}(F_{m}), and so t (x,y)φk=(xϕk,yψk)(x,y)\varphi^{k}=(x\phi^{k},y\psi^{k}), for k>0k>0. So this case is also simple: we compute p0,p1,q0,q1p_{0},p_{1},q_{0},q_{1}\in\mathbb{N} such that ϕ-logx(z)=p0+p1\phi\text{-}\log_{x}(z)=p_{0}+p_{1}\mathbb{N} and ψ-logy(w)=q0+q1\psi\text{-}\log_{y}(w)=q_{0}+q_{1}\mathbb{N} (if there are no p0p_{0} or q0q_{0}, then there are no solutions to our problem). Our set of solutions is the intersection ϕ-logx(z)ψ-logy(w)\phi\text{-}\log_{x}(z)\cap\psi\text{-}\log_{y}(w), which we can check if it is empty or not. For the conjugacy version of the problem we proceed in the same way, computing ϕ~-logx(z)\tilde{\phi}\text{-}\log_{x}(z) and ψ~-logy(w)\tilde{\psi}\text{-}\log_{y}(w) instead.

Type VII

We have that (x,y)(yψ,xϕ)(x,y)\mapsto(y\psi,x\phi), for homomorphisms ϕ:FnFm\phi:F_{n}\to F_{m} and ψ:FmFn\psi:F_{m}\to F_{n}, and so, as seen in [10, Subsection 5.7], for k0k\geq 0, (x,y)φ2k=(x(ϕψ)k,y(ψϕ)k)(x,y)\varphi^{2k}=(x(\phi\psi)^{k},y(\psi\phi)^{k}) and (x,y)φ2k+1=(y(ψϕ)kψ,x(ϕψ)kϕ)(x,y)\varphi^{2k+1}=(y(\psi\phi)^{k}\psi,x(\phi\psi)^{k}\phi). So we will describe two algorithms: one to decide the existence of an even kk, and another one to decide the existence of an odd kk. The first one is simple, since it can be seen as an application of the previous case, defining Ψ\Psi to be the type VI endomorphism mapping (x,y)(x,y) to (xϕψ,yψϕ)(x\phi\psi,y\psi\phi). Now we describe the second one. We start by checking if there is some even rr\in\mathbb{N} such that (x,y)Φr=(z,w)Φ(x,y)\Phi^{r}=(z,w)\Phi. If there is none, then there is no solution kk; otherwise compute such a minimal rr. It is proved in [10, Subsection 5.7] that the subgroup of periodic points of Φ\Phi is given by Per(Φ)=Per(ϕψ)×Per(ψϕ)\text{Per}(\Phi)=\text{Per}(\phi\psi)\times\text{Per}(\psi\phi). In particular, it is computable and we can check if a given element of Fn×FmF_{n}\times F_{m} is periodic or not. We verify if (wψ,zϕ)=(z,w)Φ(w\psi,z\phi)=(z,w)\Phi is periodic or not. If it is not, then our only candidate is k=r1k=r-1. Indeed, if there was some other k>rk>r, then we would have that

((z,w)Φ)Φkr+1=(x,y)ΦrΦkr+1=(x,y)Φk+1=(z,w)Φ\left((z,w)\Phi\right)\Phi^{k-r+1}=(x,y)\Phi^{r}\Phi^{k-r+1}=(x,y)\Phi^{k+1}=(z,w)\Phi

and if there was some other k<rk<r we would have that

(z,w)ΦΦrk=(x,y)ΦkΦrk+1=(x,y)Φr+1=(z,w)Φ.(z,w)\Phi\Phi^{r-k}=(x,y)\Phi^{k}\Phi^{r-k+1}=(x,y)\Phi^{r+1}=(z,w)\Phi.

So, assume that (z,w)Φ(z,w)\Phi is periodic with period ss. In this case, there are only finitely many checks to do since the orbit of (x,y)(x,y) is finite, as illustrated in the following picture:

{\cdots}(x,y){(x,y)}(x,y)Φ{(x,y)\Phi}{\cdots}(x,y)Φr=(z,w)Φ{(x,y)\Phi^{r}=(z,w)\Phi}(z,w)Φ3{(z,w)\Phi^{3}}(z,w)Φ2{(z,w)\Phi^{2}}
Figure 2: The orbit of (x,y)(x,y).

The conjugacy case is similar. Checking the existence of an even kk can be translated as an instance of the algorithm for Type VI endomorphisms in the same way as above. Then we verify if there is some even rr such that (x,y)Φr(z,w)Φ(x,y)\Phi^{r}\sim(z,w)\Phi. If not, then there is no solution kk, and if there is one, we compute it. We can check if (z,w)Φ(z,w)\Phi is Φ~\tilde{\Phi}-periodic, because that happens if and only if there is some s2s\in 2\mathbb{N} such that (z,w)ΦΦ2Φs(z,w)Φ(z,w)\Phi\Phi^{2}\Phi^{s}\sim(z,w)\Phi. Indeed, if there is such an ss, then (z,w)ΦΦs+2(z,w)Φ(z,w)\Phi\Phi^{s+2}\sim(z,w)\Phi and (z,w)Φ(z,w)\Phi is Φ~\tilde{\Phi}-periodic and conversely, if (z,w)Φ(z,w)\Phi is Φ~\tilde{\Phi}-periodic, then there is some p>0p\in\mathbb{N}_{>0} such that (z,w)ΦΦp(z,w)Φ(z,w)\Phi\Phi^{p}\sim(z,w)\Phi and so

(z,w)ΦΦ2Φ2p2=(z,w)ΦΦ2p(z,w)ΦΦp(z,w)Φ.(z,w)\Phi\Phi^{2}\Phi^{2p-2}=(z,w)\Phi\Phi^{2p}\sim(z,w)\Phi\Phi^{p}\sim(z,w)\Phi.

As done in the equality case, if (z,w)Φ(z,w)\Phi is not Φ~\tilde{\Phi}-periodic, there is only one possible solution and if it is, our problem amounts to finitely many checks as the orbit of (x,y)(x,y) has only finitely many conjugacy classes.

The conjugacy problem for ascending HNN-extensions of Fn×FmF_{n}\times F_{m}

Let GG be a group and Φ\Phi be an injective endomorphism of GG. The ascending HNN-extension of GG induced by Φ\Phi is the group with relative presentation

GΦ=G,xx1gx=Φ(g),gGG\ast_{\Phi}\,=\,\langle G,x\mid x^{-1}gx=\Phi(g),g\in G\rangle (3)

and with normal forms given by xigxjx^{i}gx^{-j}, with i,j0i,j\in\mathbb{N}_{0} and gG.g\in G.

In [3], it was proved that the conjugacy problem was solvable in free-by-cyclic groups, by reducing this problem to the solution of BrCP(FnF_{n}) and TCP(FnF_{n}) for automorphisms. In [26], Logan proved that the conjugacy problem was decidable for ascending HNN-extensions of a free group, extending the work in [3]. To do so, Logan solved a variation of BrCP, which we call two-sided Brinkmann’s conjugacy problem and TCP(FnF_{n}) for injective endomorphisms of the free group. The two-sided Brinkmann conjugacy problem in GG with respect to 𝒯End(G)\mathcal{T}\subseteq\text{End}(G), denoted by 2BrCP(G)\text{2BrCP}(G) is the problem of deciding, on input an endomorphism Φ𝒯\Phi\in\mathcal{T} and elements x,yGx,y\in G, whether there is some pair (r,s)2(r,s)\in\mathbb{N}^{2} such that xΦryΦsx\Phi^{r}\sim y\Phi^{s}.

We will show that TCP(Fn×FmF_{n}\times F_{m}) and 2BrCP(Fn×Fm)\text{2BrCP}(F_{n}\times F_{m}) are decidable with respect to monomorphisms, which, using Logan’s techniques implies that the conjugacy problem is decidable for ascending HNN-extensions of the direct product of two free groups. Notice that the free-abelian times free case was already dealt with in [15]. We recall that it is proved in [10, Proposition 3.2] that injective endomorphisms of Fn×FmF_{n}\times F_{m} are precisely the types VI and VII endomorphisms having injective component homomorphisms ϕ\phi and ψ\psi. We start by proving a lemma that will be useful in proving decidability of 2BrCP(Fn×Fm)\text{2BrCP}(F_{n}\times F_{m}) in the injective case.

Lemma 4.1.

There exists an algorithm taking as input an integer n>1n>1, an injective endomorphism ϕMon(Fn)\phi\in\mbox{\rm Mon}(F_{n}), and an element xFnx\in F_{n} that decides if the number of conjugacy classes in the orbit of xx is finite or not. Equivalently, it is decidable whether there is a pair (p,q)2(p,q)\in\mathbb{N}^{2} such that xϕpxϕqx\phi^{p}\sim x\phi^{q} with p>qp>q.

Proof. First, we verify if xx is ϕ~\tilde{\phi}-periodic using [26, Lemma 4.1] to decide if there is some kk\in\mathbb{N} such that xϕϕkxx\phi\phi^{k}\sim x. If xx is ϕ~\tilde{\phi}-periodic, then the orbit of xx contains only finitely many conjugacy classes and we are done. So, we may assume that xx is not ϕ~\tilde{\phi}-periodic.

Suppose that there are (p,q)2(p,q)\in\mathbb{N}^{2} such that xϕpxϕqx\phi^{p}\sim x\phi^{q} with p>qp>q and let pp be the minimal positive integer for which there is some q<pq<p such that xϕpxϕqx\phi^{p}\sim x\phi^{q}. Then, there is some u𝔽nu\in\mathbb{F}_{n} such that xϕp=u1xϕqux\phi^{p}=u^{-1}x\phi^{q}u. Notice that q>0q>0, since otherwise we would have that xx is ϕ~\tilde{\phi}-periodic. If uIm(ϕ)u\in\text{Im}(\phi), i.e., u=uϕu=u^{\prime}\phi for some uFnu^{\prime}\in F_{n}, then xϕp=(u1ϕ)(xϕq)(uϕ)x\phi^{p}=(u^{\prime-1}\phi)(x\phi^{q})(u^{\prime}\phi), which, by injectivity implies that xϕp1=u1xϕq1ux\phi^{p-1}=u^{\prime-1}x\phi^{q-1}u^{\prime} and this contradicts the minimality of pp. Thus, for the minimal pp, we know that xϕp=u1xϕqux\phi^{p}=u^{-1}x\phi^{q}u, for some uFnIm(ϕ)u\in F_{n}\setminus\text{Im}(\phi).

Therefore, the number of conjugacy classes in the orbit of xx is finite if and only if there is some pair (p,q)2(p,q)\in\mathbb{N}^{2} such that p>q>0p>q>0 and some uFnIm(ϕ)u\in F_{n}\setminus\text{Im}(\phi) such that xϕp=u1xϕqux\phi^{p}=u^{-1}x\phi^{q}u. Verifying this condition can be done by applying [26, Lemma 4.3] to verify if there is some pair (p,q)2(p,q)\in\mathbb{N}^{2} with pq0p\geq q\geq 0 such that there is some element uFnIm(ϕ)u\in F_{n}\setminus\text{Im}(\phi) satisfying (xϕ)ϕp=u1(xϕq)u(x\phi)\phi^{p}=u^{-1}(x\phi^{q})u. ∎

Theorem 4.2.

There exists an algorithm taking as input two integers n,m>1n,m>1, two elements (x,y),(z,w)Fn×Fm(x,y),(z,w)\in F_{n}\times F_{m} and an injective endomorphism ΦMon(Fn×Fm)\Phi\in\mbox{\rm Mon}(F_{n}\times F_{m}) that outputs a pair (r,s)2(r,s)\in\mathbb{N}^{2} such that (x,y)Φr(z,w)Φs(x,y)\Phi^{r}\sim(z,w)\Phi^{s} if such a pair exists and outputs NO otherwise.

Proof. Suppose that Φ\Phi is an injective endomorphism of type VI. Then, (x,y)Φ=(xϕ,yψ)(x,y)\Phi=(x\phi,y\psi) for some ϕMon(Fn)\phi\in\mbox{\rm Mon}(F_{n}) and ψMon(Fm)\psi\in\mbox{\rm Mon}(F_{m}). Naturally,

(x,y)Φr(z,w)Φs(xϕr,yψr)(zϕs,wψs){xϕrzϕsyψrwψs.(x,y)\Phi^{r}\sim(z,w)\Phi^{s}\iff(x\phi^{r},y\psi^{r})\sim(z\phi^{s},w\psi^{s})\iff\begin{cases}x\phi^{r}\sim z\phi^{s}\\ y\psi^{r}\sim w\psi^{s}\end{cases}.

We start by using [26, Proposition 4.4] to check whether there is a pair (r1,s1)(r_{1},s_{1}) such that xϕr1zϕs1x\phi^{r_{1}}\sim z\phi^{s_{1}}. If there is none, then we answer NO. If there is such a pair, we compute the minimal r1r_{1} for which there is some kk\in\mathbb{N} such that xϕr1zϕkx\phi^{r_{1}}\sim z\phi^{k}. This is possible by [26, Lemma 4.1]: we keep increasing the candidate values r1r_{1} until the algorithm in [26, Lemma 4.1] answers YES. When it does, we compute the minimal s1s_{1} such that xϕr1zϕs1x\phi^{r_{1}}\sim z\phi^{s_{1}}, which can be done by iteratively solving the conjugacy problem. Then we verify if the orbit of zz has finitely many conjugacy classes or not, using Lemma 4.1. If there are only finitely many conjugacy classes, we compute the minimal qzq_{z}\in\mathbb{N} such that zϕqzz\phi^{q_{z}} is ϕ~\tilde{\phi}-periodic and pzp_{z}\in\mathbb{N}, the ϕ~\tilde{\phi}-period of zϕqzz\phi^{q_{z}}. If there are infinitely many conjugacy classes, we say that qz=q_{z}=\infty. We claim that the set 𝒮\mathcal{S} of all pairs (p,q)2(p,q)\in\mathbb{N}^{2} such that xϕp=zϕqx\phi^{p}=z\phi^{q} is precisely

{(r1,s1)+(1,1) if qz=(r1,s1)+(1,1)+(0,pz)+(pz,0) if qzs1{(r1+i,s1+i)0i<qzs1}((r1+qzs1,qz)+(1,1)+(0,pz)+(pz,0)) if qz>s1\begin{cases}(r_{1},s_{1})+(1,1)\mathbb{N}&\text{ if $q_{z}=\infty$}\\ (r_{1},s_{1})+(1,1)\mathbb{N}+(0,p_{z})\mathbb{N}+(p_{z},0)\mathbb{N}&\text{ if $q_{z}\leq s_{1}$}\\ \{(r_{1}+i,s_{1}+i)\mid 0\leq i<q_{z}-s_{1}\}\cup\left((r_{1}+q_{z}-s_{1},q_{z})+(1,1)\mathbb{N}+(0,p_{z})\mathbb{N}+(p_{z},0)\mathbb{N}\right)&\text{ if $q_{z}>s_{1}$}\end{cases}

Suppose that qz=q_{z}=\infty. Then, clearly (r1,s1)+(1,1)𝒮(r_{1},s_{1})+(1,1)\mathbb{N}\subseteq\mathcal{S}. So, suppose that xϕpzϕqx\phi^{p}\sim z\phi^{q}. Then, by minimality of r1r_{1}, we have that pr1p\geq r_{1}. Clearly,

zϕqxϕp=xϕr1ϕpr1zϕs1ϕpr1=zϕs1+pr1.\displaystyle z\phi^{q}\sim x\phi^{p}=x\phi^{r_{1}}\phi^{p-r_{1}}\sim z\phi^{s_{1}}\phi^{p-r_{1}}=z\phi^{s_{1}+p-r_{1}}. (4)

Since the orbit of zz has infinitely many conjugacy classes, then it must be the case that q=s1+pr1q=s_{1}+p-r_{1}, i.e., qs1=pr1q-s_{1}=p-r_{1}, and so, (p,q)=(r1,s1)+(pr1)(1,1)(r1,s1)+(1,1)(p,q)=(r_{1},s_{1})+(p-r_{1})(1,1)\in(r_{1},s_{1})+(1,1)\mathbb{N}.

Now assume that qzs1q_{z}\leq s_{1}. Then zϕs1z\phi^{s_{1}} has ϕ~\tilde{\phi}-period pzp_{z}, and so xϕr1x\phi^{r_{1}} must also have ϕ~\tilde{\phi}-period pzp_{z}. Indeed, xϕr1ϕkxϕr1x\phi^{r_{1}}\phi^{k}\sim x\phi^{r_{1}} if and only if zϕs1ϕkxϕr1zϕs1z\phi^{s_{1}}\phi^{k}\sim x\phi^{r_{1}}\sim z\phi^{s_{1}}. Thus, it is clear that (r1,s1)+(1,1)+(0,pz)+(pz,0)𝒮(r_{1},s_{1})+(1,1)\mathbb{N}+(0,p_{z})\mathbb{N}+(p_{z},0)\mathbb{N}\subseteq\mathcal{S}. Now suppose that xϕpzϕqx\phi^{p}\sim z\phi^{q}. Again, by minimality of r1r_{1}, we have that pr1p\geq r_{1}, which implies that xϕpx\phi^{p} (and so, zϕqz\phi^{q}) is in the ϕ~\tilde{\phi}-periodic part of the orbit, and has period pzp_{z}. Proceeding as above, we get that zϕqzϕs1+pr1z\phi^{q}\sim z\phi^{s_{1}+p-r_{1}}. If q>s1+pr1q>s_{1}+p-r_{1}, there is some rr\in\mathbb{N} such that qps1+r1=rpzq-p-s_{1}+r_{1}=rp_{z}, which means that q=s1+(pr1)+rpzq=s_{1}+(p-r_{1})+rp_{z}, thus

(p,q)=(r1,s1)+(pr1)(1,1)+r(0,pz)(r1,s1)+(1,1)+(0,pz)+(pz,0).(p,q)=(r_{1},s_{1})+(p-r_{1})(1,1)+r(0,p_{z})\in(r_{1},s_{1})+(1,1)\mathbb{N}+(0,p_{z})\mathbb{N}+(p_{z},0)\mathbb{N}.

If, on the other hand qs1+pr1q\leq s_{1}+p-r_{1}, there is some rr\in\mathbb{N} such that s1+pr1q=rpzs_{1}+p-r_{1}-q=rp_{z}, and so p=r1+(qs1)+rpzp=r_{1}+(q-s_{1})+rp_{z}, thus

(p,q)=(r1,s1)+(qs1)(1,1)+r(pz,0)(r1,s1)+(1,1)+(0,pz)+(pz,0).(p,q)=(r_{1},s_{1})+(q-s_{1})(1,1)+r(p_{z},0)\in(r_{1},s_{1})+(1,1)\mathbb{N}+(0,p_{z})\mathbb{N}+(p_{z},0)\mathbb{N}.

Now, suppose that qz>s1q_{z}>s_{1}. Obviously {(r1+i,s1+i)0i<qzs1}𝒮\{(r_{1}+i,s_{1}+i)\mid 0\leq i<q_{z}-s_{1}\}\subseteq\mathcal{S}. Again, since xϕr1+qzs1zϕqzx\phi^{r_{1}+q_{z}-s_{1}}\sim z\phi^{q_{z}}, then xϕr1+qzs1x\phi^{r_{1}+q_{z}-s_{1}} has ϕ~\tilde{\phi}-period pzp_{z}, and so it is clear that (r1+qzs1,qz)+(1,1)+(0,pz)+(pz,0)𝒮(r_{1}+q_{z}-s_{1},q_{z})+(1,1)\mathbb{N}+(0,p_{z})\mathbb{N}+(p_{z},0)\mathbb{N}\subseteq\mathcal{S}. Now suppose that xϕpzϕqx\phi^{p}\sim z\phi^{q}. By minimality of r1r_{1}, we must have that pr1p\geq r_{1}. Suppose that p<r1+qzs1p<r_{1}+q_{z}-s_{1}. Then s1+pr1<qzs_{1}+p-r_{1}<q_{z}, which means that zϕs1+pr1z\phi^{s_{1}+p-r_{1}} is not ϕ~\tilde{\phi}-periodic. As in (4), zϕqzϕs1+pr1z\phi^{q}\sim z\phi^{s_{1}+p-r_{1}} and so, it must be that q=s1+pr1q=s_{1}+p-r_{1} and (p,q)=(r1,s1)+(pr1,qs1)=(r1,s1)+(pr1,pr1)(p,q)=(r_{1},s_{1})+(p-r_{1},q-s_{1})=(r_{1},s_{1})+(p-r_{1},p-r_{1}). Since we are assuming that pr1<qzs1p-r_{1}<q_{z}-s_{1}, it follows that (p,q){(r1+i,s1+i)0i<qzs1}.(p,q)\in\{(r_{1}+i,s_{1}+i)\mid 0\leq i<q_{z}-s_{1}\}. If, on the other hand, pr1+qzs1p\geq r_{1}+q_{z}-s_{1},

xϕpxϕr1ϕpr1zϕs1ϕpr1,x\phi^{p}\sim x\phi^{r_{1}}\phi^{p-r_{1}}\sim z\phi^{s_{1}}\phi^{p-r_{1}},

with pr1qzs1p-r_{1}\geq q_{z}-s_{1} and so xϕpx\phi^{p} (and so, zϕqz\phi^{q}) lies in the ϕ~\tilde{\phi}-periodic part of the orbit, and has period pzp_{z}. So we proceed as in the previous case. As in (4), zϕqzϕs1+pr1z\phi^{q}\sim z\phi^{s_{1}+p-r_{1}}. If q>s1+pr1q>s_{1}+p-r_{1}, there is some rr\in\mathbb{N} such that qps1+r1=rpzq-p-s_{1}+r_{1}=rp_{z}, which means that q=s1+(pr1)+rpzq=s_{1}+(p-r_{1})+rp_{z}, thus

(p,q)=(r1+qzs1,qz)+(pr1+s1qz)(1,1)+r(0,pz)(r1,s1)+(1,1)+(0,pz)+(pz,0).(p,q)=(r_{1}+q_{z}-s_{1},q_{z})+(p-r_{1}+s_{1}-q_{z})(1,1)+r(0,p_{z})\in(r_{1},s_{1})+(1,1)\mathbb{N}+(0,p_{z})\mathbb{N}+(p_{z},0)\mathbb{N}.

If, on the other hand qs1+pr1q\leq s_{1}+p-r_{1}, there is some rr\in\mathbb{N} such that s1+pr1q=rpzs_{1}+p-r_{1}-q=rp_{z}, and so p=r1+(qs1)+rpzp=r_{1}+(q-s_{1})+rp_{z}, thus

(p,q)=(r1+qzs1,qz)+(qqz)(1,1)+r(pz,0)(r1,s1)+(1,1)+(0,pz)+(pz,0).(p,q)=(r_{1}+q_{z}-s_{1},q_{z})+(q-q_{z})(1,1)+r(p_{z},0)\in(r_{1},s_{1})+(1,1)\mathbb{N}+(0,p_{z})\mathbb{N}+(p_{z},0)\mathbb{N}.

Proceeding analogously, we compute the set 𝒮2\mathcal{S}^{\prime}\subseteq\mathbb{N}^{2} of pairs (p,q)2(p,q)\in\mathbb{N}^{2} such that yϕp=wϕqy\phi^{p}=w\phi^{q}. These are semilinear sets, so their intersection can be computed (in particular its emptyness can be decided) and the solution set for our problem is precisely 𝒮𝒮\mathcal{S}\cap\mathcal{S}^{\prime}.

Now suppose that Φ\Phi is a type VII injective endomorphism. Then (x,y)Φ=(yψ,xϕ)(x,y)\Phi=(y\psi,x\phi) for some injective homomorphisms ϕ:FnFm\phi:F_{n}\to F_{m} and ψ:FmFn\psi:F_{m}\to F_{n} and, for k0k\geq 0, (x,y)φ2k=(x(ϕψ)k,y(ψϕ)k)(x,y)\varphi^{2k}=(x(\phi\psi)^{k},y(\psi\phi)^{k}) and (x,y)φ2k+1=(y(ψϕ)kψ,x(ϕψ)kϕ)(x,y)\varphi^{2k+1}=(y(\psi\phi)^{k}\psi,x(\phi\psi)^{k}\phi). We prove that we can decide if there is a pair (r,s)2(r,s)\in\mathbb{N}^{2} by showing that it is decidable whether there is such a pair (r,s)(r,s) in 2×22\mathbb{N}\times 2\mathbb{N}, in 2×(2+1)2\mathbb{N}\times(2\mathbb{N}+1), in (2+1)×2(2\mathbb{N}+1)\times 2\mathbb{N} and in (2+1)×(2+1)(2\mathbb{N}+1)\times(2\mathbb{N}+1).

The first case is immediate, since we are deciding if there are some r,s2r^{\prime},s^{\prime}\in\mathbb{N}^{2} such that (x(ϕψ)r,y(ψϕ)r)(z(ϕψ)s,w(ψϕ)s)(x(\phi\psi)^{r^{\prime}},y(\psi\phi)^{r^{\prime}})\sim(z(\phi\psi)^{s^{\prime}},w(\psi\phi)^{s^{\prime}}), which can be reduced to the type VI case, by defining the defining the type VI monomorphism given by Ψ:(x,y)(xϕψ,yψϕ)\Psi:(x,y)\mapsto(x\phi\psi,y\psi\phi). The second and third cases are the same since one can be obtained from the other by swapping the roles of (x,y)(x,y) and (z,w)(z,w). We will now see that we can decide whether there are some r,sr^{\prime},s^{\prime}\in\mathbb{N} such that (x,y)Φ2r+1(z,w)Φ2s(x,y)\Phi^{2r^{\prime}+1}\sim(z,w)\Phi^{2s^{\prime}}. But

(x,y)Φ2r+1(z,w)Φ2s\displaystyle(x,y)\Phi^{2r^{\prime}+1}\sim(z,w)\Phi^{2s^{\prime}}
\displaystyle\iff\; (y(ψϕ)rψ,x(ϕψ)rϕ)(z(ϕψ)s,w(ψϕ)s)\displaystyle(y(\psi\phi)^{r^{\prime}}\psi,x(\phi\psi)^{r^{\prime}}\phi)\sim(z(\phi\psi)^{s^{\prime}},w(\psi\phi)^{s^{\prime}})
\displaystyle\iff\; ((yψ)(ϕψ)r,(xϕ)(ψϕ)r)(z(ϕψ)s,w(ψϕ)s)\displaystyle((y\psi)(\phi\psi)^{r^{\prime}},(x\phi)(\psi\phi)^{r^{\prime}})\sim(z(\phi\psi)^{s^{\prime}},w(\psi\phi)^{s^{\prime}})

which is again an instance of the case for the type VI injective endomorphism Ψ\Psi defined above with input (yψ,xϕ)(y\psi,x\phi) and (z,w)(z,w). We can use the same trick to solve the remaining case of deciding if there is a solution (r,s)(2+1)×(2+1)(r,s)\in(2\mathbb{N}+1)\times(2\mathbb{N}+1). We want to decide whether there are some r,sr^{\prime},s^{\prime}\in\mathbb{N} such that (x,y)Φ2r+1(z,w)Φ2s+1(x,y)\Phi^{2r^{\prime}+1}\sim(z,w)\Phi^{2s^{\prime}+1}. But

(x,y)Φ2r+1(z,w)Φ2s+1\displaystyle(x,y)\Phi^{2r^{\prime}+1}\sim(z,w)\Phi^{2s^{\prime}+1}
\displaystyle\iff\; (y(ψϕ)rψ,x(ϕψ)rϕ)(w(ψϕ)sψ,z(ϕψ)sϕ)\displaystyle(y(\psi\phi)^{r^{\prime}}\psi,x(\phi\psi)^{r^{\prime}}\phi)\sim(w(\psi\phi)^{s^{\prime}}\psi,z(\phi\psi)^{s^{\prime}}\phi)
\displaystyle\iff\; ((yψ)(ϕψ)r,(xϕ)(ψϕ)r)((wψ)(ϕψ)s,(zϕ)(ψϕ)s)\displaystyle((y\psi)(\phi\psi)^{r^{\prime}},(x\phi)(\psi\phi)^{r^{\prime}})\sim((w\psi)(\phi\psi)^{s^{\prime}},(z\phi)(\psi\phi)^{s^{\prime}})

which is again an instance of the case for the type VI injective endomorphism Ψ\Psi defined above with input (yψ,xϕ)(y\psi,x\phi) and (wψ,zϕ)(w\psi,z\phi). ∎

Now we can prove decidability of TCP(Fn×FmF_{n}\times F_{m}) for monomorphisms of Fn×FmF_{n}\times F_{m}.

Proposition 4.3.

There exists an algorithm taking as input two integers n,m>1n,m>1, two elements (x,y),(z,w)Fn×Fm(x,y),(z,w)\in F_{n}\times F_{m} and an injective endomorphism ΦMon(Fn×Fm)\Phi\in\mbox{\rm Mon}(F_{n}\times F_{m}) that outputs an element (u,v)Fn×Fm(u,v)\in F_{n}\times F_{m} such that (x,y)=(u1,v1)Φ(z,w)(u,v)(x,y)=(u^{-1},v^{-1})\Phi(z,w)(u,v) if such an element exists and outputs NO otherwise.

Proof. Suppose that Φ\Phi is a type VI monomorphism. Then, (x,y)Φ=(xϕ,yψ)(x,y)\Phi=(x\phi,y\psi) for some ϕMon(Fn)\phi\in\mbox{\rm Mon}(F_{n}) and ψMon(Fm)\psi\in\mbox{\rm Mon}(F_{m}). We have that

(x,y)=(u1,v1)Φ(z,w)(u,v){x=(u1ϕ)zuy=(v1ψ)wv.(x,y)=(u^{-1},v^{-1})\Phi(z,w)(u,v)\iff\begin{cases}x=(u^{-1}\phi)zu\\ y=(v^{-1}\psi)wv\end{cases}.

So, there exists an element (u,v)Fn×Fm(u,v)\in F_{n}\times F_{m} as desired if and only if TCP(FnF_{n}) answers YES on input (x,z,ϕ)(x,z,\phi) and TCP(FmF_{m}) answers YES on input (y,w,ψ)(y,w,\psi). That is decidable by [26, Lemma 2.2] or [33, Theorem 2.4].

Now assume that Φ\Phi is a monomorphism of type VII. Then (x,y)Φ=(yψ,xϕ)(x,y)\Phi=(y\psi,x\phi), for injective homomorphisms ϕ:FnFm\phi:F_{n}\to F_{m} and ψ:FmFn\psi:F_{m}\to F_{n}. Applying this description of Φ\Phi, we obtain that

(x,y)=(u1,v1)Φ(z,w)(u,v){x=(v1ψ)zuy=(u1ϕ)wv{u1=x1(v1ψ)zy=(u1ϕ)wv\displaystyle(x,y)=(u^{-1},v^{-1})\Phi(z,w)(u,v)\iff\begin{cases}x=(v^{-1}\psi)zu\\ y=(u^{-1}\phi)wv\end{cases}\iff\begin{cases}u^{-1}=x^{-1}(v^{-1}\psi)z\\ y=(u^{-1}\phi)wv\end{cases}
\displaystyle\iff {u1=x1(v1ψ)zy=(x1ϕ)(v1ψϕ)(zϕ)wv{u=z1(vψ)x(xϕ)y=(v1ψϕ)(zϕ)wv.\displaystyle\begin{cases}u^{-1}=x^{-1}(v^{-1}\psi)z\\ y=(x^{-1}\phi)(v^{-1}\psi\phi)(z\phi)wv\end{cases}\iff\begin{cases}u=z^{-1}(v\psi)x\\ (x\phi)y=(v^{-1}\psi\phi)(z\phi)wv\end{cases}.

So, if (x,y)(x,y) and (z,w)(z,w) are Φ\Phi-twisted conjugate with conjugator (u,v)(u,v), then (xϕ)y(x\phi)y and (zϕ)w(z\phi)w are ψϕ\psi\phi-twisted conjugate with conjugator vv. Conversely, if (zϕ)w(z\phi)w and (xϕ)y(x\phi)y are ψϕ\psi\phi-twisted conjugate with some conjugator vv, then (x,y)(x,y) and (z,w)(z,w) are Φ\Phi-twisted conjugate with conjugator (z1(vψ)x,v)(z^{-1}(v\psi)x,v). Since it can be decided whether (zϕ)w(z\phi)w and (xϕ)y(x\phi)y are ψϕ\psi\phi-twisted conjugate, by solving TCP(FmF_{m}), we are done. ∎

By the results in [26], the two results above yield the following corollary.

Corollary 4.4.

The conjugacy problem is solvable for ascending HNN-extensions of Fn×FmF_{n}\times F_{m}, for n,m>1n,m>1.

Dynamics at the infinity

A metric space (X,d)(X,d) is said to be a median space if, for all x,y,zXx,y,z\in X, there is some unique point μ(x,y,z)X\mu(x,y,z)\in X, known as the median of x,y,zx,y,z, such that d(x,y)=d(x,μ(x,y,z))+d(μ(x,y,z),y);d(x,y)=d(x,\mu(x,y,z))+d(\mu(x,y,z),y); d(y,z)=d(y,μ(x,y,z))+d(μ(x,y,z),z);d(y,z)=d(y,\mu(x,y,z))+d(\mu(x,y,z),z); and d(z,x)=d(z,μ(x,y,z))+d(μ(x,y,z),x)d(z,x)=d(z,\mu(x,y,z))+d(\mu(x,y,z),x). We call μ:X3X\mu:X^{3}\to X the median operator of the median space XX.

Coarse median spaces were introduced by Bowditch in [5]. Following the equivalent definition given in [28], we say that, given a metric space XX, a coarse median on XX is a ternary operation μ:X3X\mu:X^{3}\to X satisfying the following:

there exists a constant C0C\geq 0 such that, for all a,b,c,xXa,b,c,x\in X, we have that

  1. 1.

    μ(a,a,b)=a and μ(a,b,c)=μ(b,c,a)=μ(b,a,c);\mu(a,a,b)=a\text{ and }\mu(a,b,c)=\mu(b,c,a)=\mu(b,a,c);

  2. 2.

    d(μ(μ(a,x,b),x,c),μ(a,x,μ(b,x,c)))C;d(\mu(\mu(a,x,b),x,c),\mu(a,x,\mu(b,x,c)))\leq C;

  3. 3.

    d(μ(a,b,c),μ(x,b,c))Cd(a,x)+C.d(\mu(a,b,c),\mu(x,b,c))\leq Cd(a,x)+C.

Given a group GG, a word metric on GG measures the distance of the shortest path in the Cayley graph of GG with respect to some set of generators, i.e., for two elements g,hGg,h\in G, we have that d(g,h)d(g,h) is the length of the shortest word whose letters come from the generating set representing g1hg^{-1}h. Following the definitions in [23], two coarse medians μ1,μ2:X3X\mu_{1},\mu_{2}:X^{3}\to X are said to be at bounded distance if there exists some constant CC such that d(μ1(x,y,z),μ2(x,y,z))Cd(\mu_{1}(x,y,z),\mu_{2}(x,y,z))\leq C for all x,y,zXx,y,z\in X, and a coarse median structure on XX is an equivalence class [μ][\mu] of coarse medians pairwise at bounded distance. When XX is a metric space and [μ][\mu] is a coarse median structure on XX, we say that (X,[μ])(X,[\mu]) is a coarse median space. Following Fioravanti’s definition, a coarse median group is a pair (G,[μ])(G,[\mu]), where GG is a finitely generated group with a word metric dd and [μ][\mu] is a GG-invariant coarse median structure on GG, meaning that for each gGg\in G, there is a constant C(g)C(g) such that d(gμ(g1,g2,g3),μ(gg1,gg2,gg3)C(g)d(g\mu(g_{1},g_{2},g_{3}),\mu(gg_{1},gg_{2},gg_{3})\leq C(g), for all g1,g2,g3Gg_{1},g_{2},g_{3}\in G. The author in [23] also remarks that this definition is stronger than the original definition from [5], that did not require GG-invariance. Despite being better suited for this work, it is not quasi-isometry-invariant nor commensurability-invariant, unlike Bowditch’s version.

Now, a KK-hyperbolic group is such that there is some K0K\geq 0 for which every geodesic triangle has a KK-center, i.e., a point that, up to a bounded distance, depends only on the vertices, and is KK-close to every edge of the triangle. Given three points, the operator that associates the three points to the KK-center of a geodesic triangle they define is coarse median. In fact, by [28, Theorem 4.2] it is the only coarse-median structure that we can endow XX with. Given two groups G1G_{1} and G2G_{2} endowed with coarse median operators μ1\mu_{1} and μ2\mu_{2}, then it is easy to check that the operator μ:(G1×G2)3G1×G2\mu:(G_{1}\times G_{2})^{3}\to G_{1}\times G_{2} defined by μ((x1,x2),(y1,y2),(z1,z2))=(μ1(x1,y1,z1),μ2(x2,y2,z2))\mu((x_{1},x_{2}),(y_{1},y_{2}),(z_{1},z_{2}))=(\mu_{1}(x_{1},y_{1},z_{1}),\mu_{2}(x_{2},y_{2},z_{2})) is coarse median. We refer to μ\mu as the product coarse median operator and this is the coarse median we will endow Fn×FmF_{n}\times F_{m} with.

Given a coarse median group (G,[μ])(G,[\mu]), an automorphism φAut(G)\varphi\in\text{Aut}(G) is said to be coarse-median preserving if it fixes [μ][\mu], i.e., when there is some constant C0C\geq 0 such that for all gig_{i}^{\prime}s, we have that

d(μ(g1,g2,g3)φ,μ(g1φ,g2φ,g3φ))C,d(\mu(g_{1},g_{2},g_{3})\varphi,\mu(g_{1}\varphi,g_{2}\varphi,g_{3}\varphi))\leq C,

with respect to some word metric dd. This can naturally be defined for general endomorphisms, not necessarily bijective, as done in [8, 9], and even to homomorphisms between different groups: we say that a homomorphism φ:(G1,μ1)(G2,μ2)\varphi:(G_{1},\mu_{1})\to(G_{2},\mu_{2}) is coarse-median preserving if there is some C0C\geq 0 such that for all x,y,zG1x,y,z\in G_{1},

d(μ2(xφ,yφ,zφ),μ1(x,y,z)φ)Cd(\mu_{2}(x\varphi,y\varphi,z\varphi),\mu_{1}(x,y,z)\varphi)\leq C

for some word metric dd of G2G_{2}.

Recall the notation from Section 2, namely the definition of sets X,Y,ZX,\,Y,\,Z and WW.

It is shown in [9, Theorem 4.11] that an endomorphism ϕ\phi of a hyperbolic group GG is coarse-median preserving if and only if the bounded reduction property holds for ϕ\phi. In particular, if GG is free, ϕ\phi is coarse-median preserving if and only if it is either trivial or injective. Following the same strategy as in [9, Theorem 4.11], we can see that the result holds for homomorphisms between free groups of different ranks.

Proposition 5.1.

Let φ:FnFm\varphi:F_{n}\to F_{m} be a homomorphism. Then φ\varphi is coarse-median preserving if and only if it is either trivial or injective.

Proof. Suppose that φ\varphi is not trivial nor injective. Then there are some x,yFnx,y\in F_{n} such that xφ=1x\varphi=1 and yφ1y\varphi\neq 1. For all CC\in\mathbb{N}, we have that

d(μ1(x,yCxyC,yC)φ,μ2(xφ,(yCxyC)φ,yCφ))\displaystyle d(\mu_{1}(x,y^{C}xy^{-C},y^{C})\varphi,\mu_{2}(x\varphi,(y^{C}xy^{-C})\varphi,y^{C}\varphi))
=\displaystyle=\; d(yCφ,μ2(1,1,yCφ))\displaystyle d(y^{C}\varphi,\mu_{2}(1,1,y^{C}\varphi))
=\displaystyle=\; d(1,yCφ)\displaystyle d(1,y^{C}\varphi)
\displaystyle\geq\; C.\displaystyle C.

Since CC can be arbitrarily large, φ\varphi is not coarse-median preserving.

Clearly, if φ\varphi is trivial, it is coarse-median preserving, so the only thing left to prove is that injectivity implies coarse-median preservation. Suppose that φ\varphi is injective. By [9, Proposition 5.3], for all M0M\geq 0, there is some N0N\geq 0 such that

(u|v)M(uφ|vφ)N,(u|v)\leq M\implies(u\varphi|v\varphi)\leq N,

for all u,vFnu,v\in F_{n}. Following the proofs of [9, Proposition 4.2 and Theorem 4.6] step by step, we get that there is some NN\in\mathbb{N} such that, for all x,yFnx,y\in F_{n}, the image of the geodesic [x,y][x,y] is at a Hausdorff distance at most NN from the geodesic [xφ,yφ][x\varphi,y\varphi]. Let x,y,zFnx,y,z\in F_{n}. Then μ1(x,y,z)\mu_{1}(x,y,z) belongs to [x,y][x,y], [y,z][y,z] and [x,z][x,z], and so μ1(x,y,z)φ\mu_{1}(x,y,z)\varphi is at a distance at most NN from [xφ,yφ][x\varphi,y\varphi], [yφ,zφ][y\varphi,z\varphi] and [xφ,zφ][x\varphi,z\varphi], thus μ1(x,y,z)φ\mu_{1}(x,y,z)\varphi is an NN-center of the triangle [[xφ,yφ,zφ]][[x\varphi,y\varphi,z\varphi]]. Since μ2(xφ,yφ,zφ)\mu_{2}(x\varphi,y\varphi,z\varphi) is also an NN-center of [[xφ,yφ,zφ]][[x\varphi,y\varphi,z\varphi]], by [4, Lemma 3.1.5], it is at a bounded distance from μ1(x,y,z)φ\mu_{1}(x,y,z)\varphi and this bound depends only on NN.

Therefore, φ\varphi is coarse-median preserving. ∎

Now we establish an equivalence between preservation of a coarse median and uniform continuity in Fn×FmF_{n}\times F_{m}, similar to what happens in free and free-abelian times free groups.

Theorem 5.2.

Coarse-median preserving endomorphisms of Fn×FmF_{n}\times F_{m} are precisely the uniformly continuous ones.

Proof. Let ΦEnd(Fn×Fm)\Phi\in\text{End}(F_{n}\times F_{m}) be a coarse-median preserving endomorphism with constant CC.

We start by proving that 1XX={1}1\in X\implies X=\{1\}. Suppose there are some i,j[n]i,j\in[n] such that (ai,1)Φ=(1,yi)(a_{i},1)\Phi=(1,y_{i}) and (aj,1)Φ=(xj,yj)(a_{j},1)\Phi=(x_{j},y_{j}) with xj1x_{j}\neq 1. Then, letting DD be a number greater than CC,

d(μ((ajDai,1)Φ,(ajDaiajD,1)Φ,(ai,1)Φ),μ((ajDai,1),(ajDaiajD,1),(ai,1))Φ)\displaystyle d\left(\mu\left((a_{j}^{-D}a_{i},1)\Phi,(a_{j}^{-D}a_{i}a_{j}^{D},1)\Phi,(a_{i},1)\Phi\right),\mu\left((a_{j}^{-D}a_{i},1),(a_{j}^{-D}a_{i}a_{j}^{D},1),(a_{i},1)\right)\Phi\right)
=\displaystyle=\; d(μ((xjD,yjDyi),(1,yjDyiyjD),(1,yi)),(μ1(ajDai,ajDaiajD,ai),μ2(1,1,1))Φ)\displaystyle d\left(\mu\left((x_{j}^{-D},y_{j}^{-D}y_{i}),(1,y_{j}^{-D}y_{i}y_{j}^{D}),(1,y_{i})\right),\left(\mu_{1}(a_{j}^{-D}a_{i},a_{j}^{-D}a_{i}a_{j}^{D},a_{i}),\mu_{2}(1,1,1)\right)\Phi\right)
=\displaystyle=\; d((μ1(xjD,1,1),μ2(yjDyi,yjDyiyjD,yi)),(ajDai,1)Φ)\displaystyle d\left(\left(\mu_{1}(x_{j}^{-D},1,1),\mu_{2}(y_{j}^{-D}y_{i},y_{j}^{-D}y_{i}y_{j}^{D},y_{i})\right),(a_{j}^{-D}a_{i},1)\Phi\right)
\displaystyle\geq\; d(μ1(xjD,1,1),xjD)\displaystyle d(\mu_{1}(x_{j}^{-D},1,1),x_{j}^{-D})
=\displaystyle=\; d(1,xjD)\displaystyle d(1,x_{j}^{-D})
>\displaystyle>\; C\displaystyle C

which contradicts the fact that Φ\Phi is coarse-median preserving with constant CC. So, 1XX={1}1\in X\implies X=\{1\}. The same holds for Y,ZY,\,Z and WW, and that can be seen analogously.

We now prove that we cannot have YY and WW to be nontrivial simultaneously, which shows that Φ\Phi cannot be of type I, II or V. Suppose that Y{1}Y\neq\{1\} and W{1}W\neq\{1\}. In view of the above, 1YW1\not\in Y\cup W. We know that there is some word vFmv\in F_{m} and nonzero exponents qiq_{i} and sjs_{j} such that yi=vqiy_{i}=v^{q_{i}} and wj=vsjw_{j}=v^{s_{j}}. For DD\in\mathbb{N}, we have that

μ((aiDsj,bjDqi)Φ,(aisgn(sj)+2Dsj,bjDqi)Φ,(ai2Dsj,bj2Dqi)Φ)\displaystyle\mu\left((a_{i}^{Ds_{j}},b_{j}^{-Dq_{i}})\Phi,(a_{i}^{\text{sgn}(s_{j})+2Ds_{j}},b_{j}^{-Dq_{i}})\Phi,(a_{i}^{2Ds_{j}},b_{j}^{-2Dq_{i}})\Phi\right)
=\displaystyle=\; μ((xiDsjzjDqi,1),(xi2Dsj+sgn(sj)zjDqi,vqi(sgn(sj)+Dsj)),(xi2Dsjzj2Dqi,1))\displaystyle\mu\left((x_{i}^{Ds_{j}}z_{j}^{-Dq_{i}},1),(x_{i}^{2Ds_{j}+\text{sgn}(s_{j})}z_{j}^{-Dq_{i}},v^{q_{i}(\text{sgn}(s_{j})+Ds_{j})}),(x_{i}^{2Ds_{j}}z_{j}^{-2Dq_{i}},1)\right)
=\displaystyle=\; (μ1(xiDsjzjDqi,xi2Dsjsgn(sj)zjDqi,xi2Dsjzj2Dqi),μ2(1,vqi(sgn(sj)+Dsj),1))\displaystyle\left(\mu_{1}(x_{i}^{Ds_{j}}z_{j}^{-Dq_{i}},x_{i}^{2Ds_{j}\text{sgn}(s_{j})}z_{j}^{-Dq_{i}},x_{i}^{2Ds_{j}}z_{j}^{-2Dq_{i}}),\mu_{2}(1,v^{q_{i}(\text{sgn}(s_{j})+Ds_{j})},1)\right)

and

μ((aiDsj,bjDqi),(aisgn(sj)+2Dsj,bjDqi),(ai2Dsj,bj2Dqi))Φ\displaystyle\mu\left((a_{i}^{Ds_{j}},b_{j}^{-Dq_{i}}),(a_{i}^{\text{sgn}(s_{j})+2Ds_{j}},b_{j}^{-Dq_{i}}),(a_{i}^{2Ds_{j}},b_{j}^{-2Dq_{i}})\right)\Phi
=\displaystyle=\; (μ1(aiDsj,aisgn(sj)+2Dsj,ai2Dsj),μ2(bjDqi,bjDqi,bj2Dqi))Φ\displaystyle\left(\mu_{1}(a_{i}^{Ds_{j}},a_{i}^{\text{sgn}(s_{j})+2Ds_{j}},a_{i}^{2Ds_{j}}),\mu_{2}(b_{j}^{-Dq_{i}},b_{j}^{-Dq_{i}},b_{j}^{-2Dq_{i}})\right)\Phi
=\displaystyle=\; (ai2Dsj,bjDqi)Φ.\displaystyle(a_{i}^{2Ds_{j}},b_{j}^{-Dq_{i}})\Phi.

Then, for D>CD>C, the distance between

μ((aiDsj,bjDqi)Φ,(aisgn(sj)+2Dsj,bjDqi)Φ,(ai2Dsj,bj2Dqi)Φ)\mu\left((a_{i}^{Ds_{j}},b_{j}^{-Dq_{i}})\Phi,(a_{i}^{\text{sgn}(s_{j})+2Ds_{j}},b_{j}^{-Dq_{i}})\Phi,(a_{i}^{2Ds_{j}},b_{j}^{-2Dq_{i}})\Phi\right)

and

(μ((aiDsj,bjDqi),(aisgn(sj)+2Dsj,bjDqi),(ai2Dsj,bj2Dqi))Φ\left(\mu((a_{i}^{Ds_{j}},b_{j}^{-Dq_{i}}),(a_{i}^{\text{sgn}(s_{j})+2Ds_{j}},b_{j}^{-Dq_{i}}),(a_{i}^{2Ds_{j}},b_{j}^{-2Dq_{i}})\right)\Phi

is greater than or equal to

d(μ2(1,vqi(sgn(sj)+Dsj),1),vDsjqi)=d(1,vDsjqi)C,d(\mu_{2}(1,v^{q_{i}(\text{sgn}(s_{j})+Ds_{j})},1),v^{Ds_{j}q_{i}})=d(1,v^{Ds_{j}q_{i}})\geq C,

which contradicts the fact that Φ\Phi is coarse-median preserving with constant CC. The same argument yields that we cannot have both XX and ZZ nontrivial, so type III is also dealt with.

Since coarse-median and uniform continuity coincide for free groups, it follows from [8, Lemma 3.2] that an endomorphism of type VI is coarse-median preserving if and only if it is uniformly continuous.

If Φ\Phi is a type IV endomorphism defined as (x,y)Φ=(yϕ,yψ)(x,y)\Phi=(y\phi,y\psi), then, for all (x1,x2),(y1,y2),(z1,z2)Fn×Fm(x_{1},x_{2}),(y_{1},y_{2}),(z_{1},z_{2})\in F_{n}\times F_{m} and C>0C>0, we have that

d(μ((x1,x2),(y1,y2),(z1,z2))Φ,μ((x1,x2)Φ,(y1,y2)Φ,(z1,z2)Φ))C\displaystyle d(\mu((x_{1},x_{2}),(y_{1},y_{2}),(z_{1},z_{2}))\Phi,\mu((x_{1},x_{2})\Phi,(y_{1},y_{2})\Phi,(z_{1},z_{2})\Phi))\leq C
\displaystyle\iff\; d((μ1(x1,y1,z1),μ2(x2,y2,z2))Φ,μ((x2ϕ,x2ψ),(y2ϕ,y2ψ),(z2ϕ,z2ψ)))C\displaystyle d((\mu_{1}(x_{1},y_{1},z_{1}),\mu_{2}(x_{2},y_{2},z_{2}))\Phi,\mu((x_{2}\phi,x_{2}\psi),(y_{2}\phi,y_{2}\psi),(z_{2}\phi,z_{2}\psi)))\leq C
\displaystyle\iff\; d((μ2(x2,y2,z2)ϕ,μ2(x2,y2,z2)ψ),(μ1(x2ϕ,y2ϕ,z2ϕ),μ2(x2ψ,y2ψ,z2ψ)))C\displaystyle d((\mu_{2}(x_{2},y_{2},z_{2})\phi,\mu_{2}(x_{2},y_{2},z_{2})\psi),(\mu_{1}(x_{2}\phi,y_{2}\phi,z_{2}\phi),\mu_{2}(x_{2}\psi,y_{2}\psi,z_{2}\psi)))\leq C
\displaystyle\iff\; {d((μ2(x2,y2,z2)ϕ,μ1(x2ϕ,y2ϕ,z2ϕ))Cd(μ2(x2,y2,z2)ψ,μ2(x2ψ,y2ψ,z2ψ))C.\displaystyle\begin{cases}d((\mu_{2}(x_{2},y_{2},z_{2})\phi,\mu_{1}(x_{2}\phi,y_{2}\phi,z_{2}\phi))\leq C\\ d(\mu_{2}(x_{2},y_{2},z_{2})\psi,\mu_{2}(x_{2}\psi,y_{2}\psi,z_{2}\psi))\leq C\end{cases}.

Hence Φ\Phi is coarse-median preserving if and only if both ϕ\phi and ψ\psi are. Since, in view of Proposition 5.1 this is equivalent to having injective or trivial ϕ\phi and ψ\psi. These are precisely the uniformly continuous type IV endomorphisms of Fn×FmF_{n}\times F_{m} (see [10, Proposition 6.2]).

Finally, if Φ\Phi is a type VII endomorphism defined as (x,y)Φ=(yψ,xϕ)(x,y)\Phi=(y\psi,x\phi), then, for all (x1,x2),(y1,y2),(z1,z2)Fn×Fm(x_{1},x_{2}),(y_{1},y_{2}),(z_{1},z_{2})\in F_{n}\times F_{m} and C>0C>0, we have that

d(μ((x1,x2),(y1,y2),(z1,z2))Φ,μ((x1,x2)Φ,(y1,y2)Φ,(z1,z2)Φ))C\displaystyle d(\mu((x_{1},x_{2}),(y_{1},y_{2}),(z_{1},z_{2}))\Phi,\mu((x_{1},x_{2})\Phi,(y_{1},y_{2})\Phi,(z_{1},z_{2})\Phi))\leq C
\displaystyle\iff\; d((μ1(x1,y1,z1),μ2(x2,y2,z2))Φ,μ((x2ψ,x1ϕ),(y2ψ,y1ϕ),(z2ψ,z1ϕ)))C\displaystyle d((\mu_{1}(x_{1},y_{1},z_{1}),\mu_{2}(x_{2},y_{2},z_{2}))\Phi,\mu((x_{2}\psi,x_{1}\phi),(y_{2}\psi,y_{1}\phi),(z_{2}\psi,z_{1}\phi)))\leq C
\displaystyle\iff\; d((μ2(x2,y2,z2)ψ,μ1(x1,y1,z1)ϕ),(μ1(x2ψ,y2ψ,z2ψ),μ2(x1ϕ,y1ϕ,z1ϕ)))C\displaystyle d((\mu_{2}(x_{2},y_{2},z_{2})\psi,\mu_{1}(x_{1},y_{1},z_{1})\phi),(\mu_{1}(x_{2}\psi,y_{2}\psi,z_{2}\psi),\mu_{2}(x_{1}\phi,y_{1}\phi,z_{1}\phi)))\leq C
\displaystyle\iff\; {d((μ2(x2,y2,z2)ψ,μ1(x2ψ,y2ψ,z2ψ))Cd(μ1(x1,y1,z1)ϕ,μ2(x1ϕ,y1ϕ,z1ϕ))C.\displaystyle\begin{cases}d((\mu_{2}(x_{2},y_{2},z_{2})\psi,\mu_{1}(x_{2}\psi,y_{2}\psi,z_{2}\psi))\leq C\\ d(\mu_{1}(x_{1},y_{1},z_{1})\phi,\mu_{2}(x_{1}\phi,y_{1}\phi,z_{1}\phi))\leq C\end{cases}.

It follows, as in the type IV case, that Φ\Phi is coarse-median preserving if and only if both ϕ\phi and ψ\psi are. Since, in view of Proposition 5.1 this is equivalent to having injective or trivial ϕ\phi and ψ\psi. These are precisely the uniformly continuous type VII endomorphisms of Fn×FmF_{n}\times F_{m} (see [10, Proposition 6.2]). ∎

We will now denote by d1d_{1} the prefix metric on FnF_{n}, by d2d_{2} the prefix metric on FmF_{m} and by dd the product metric obtained by taking d1d_{1} and d2d_{2} in the factors. We can now prove the main result of the section.

Theorem 5.3.

Let ΦAut(Fn×Fm)\Phi\in\text{Aut}(F_{n}\times F_{m}), dd be the product metric given by taking the prefix metric in each factor and let Φ^\widehat{\Phi} be its continuous extension to the completion with respect to dd. Then every point in Fn×Fm^\widehat{F_{n}\times F_{m}} is either wandering or periodic.

Proof. Let ΦAut(Fn×Fm)\Phi\in\text{Aut}(F_{n}\times F_{m}) be a type VI automorphism of Fn×FmF_{n}\times F_{m}. Then Φ\Phi is of the form (x,y)(xϕ,yψ)(x,y)\mapsto(x\phi,y\psi), for some ϕAut(Fn)\phi\in\text{Aut}(F_{n}) and ψAut(Fm)\psi\in\text{Aut}(F_{m}). By [10, Proposition 6.2], Φ\Phi is uniformly continuous with respect to the metric dd and, by uniqueness of the extension Φ^\widehat{\Phi} is given by (x,y)(xϕ^,yψ^)(x,y)\mapsto(x\widehat{\phi},y\widehat{\psi}). Let (x,y)Fn^×Fm^(x,y)\in\widehat{F_{n}}\times\widehat{F_{m}} be a nonwandering point. Then,

ε>0Nn>N:B((x,y),ε)Φ^nB((x,y),ε).\forall\varepsilon>0\;\forall N\in\mathbb{N}\;\exists n>N:\;\;B((x,y),\varepsilon)\widehat{\Phi}^{n}\cap B((x,y),\varepsilon)\neq\emptyset.

Rewriting, we obtain that

ε>0Nn>N(z,w)Fn^×Fm^:{d((x,y),(z,w))<εd((x,y),(z,w)Φ^n)<ε,\forall\varepsilon>0\;\forall N\in\mathbb{N}\;\exists n>N\;\exists(z,w)\in\widehat{F_{n}}\times\widehat{F_{m}}:\;\;\begin{cases}d((x,y),(z,w))<\varepsilon\\ d((x,y),(z,w)\widehat{\Phi}^{n})<\varepsilon\end{cases},

which is the same as saying that

ε>0Nn>N(z,w)Fn^×Fm^:{d1(x,z)<εd1(x,zϕ^n)<εd2(y,w)<εd2(y,wψ^n)<ε.\forall\varepsilon>0\;\forall N\in\mathbb{N}\;\exists n>N\;\exists(z,w)\in\widehat{F_{n}}\times\widehat{F_{m}}:\;\;\begin{cases}d_{1}(x,z)<\varepsilon\\ d_{1}(x,z\widehat{\phi}^{n})<\varepsilon\\ d_{2}(y,w)<\varepsilon\\ d_{2}(y,w\widehat{\psi}^{n})<\varepsilon\end{cases}.

In particular, the above condition yields that xx is ϕ^\widehat{\phi}-nonwandering and yy is ψ^\widehat{\psi}-nonwandering. By [8, Corollary 5.11], xx must be ϕ^\widehat{\phi}-periodic and yy must be ψ^\widehat{\psi}-periodic, and so (x,y)(x,y) is Φ^\widehat{\Phi}-periodic.

Now, suppose that n=mn=m and let ΦAut(Fn×Fm)\Phi\in\text{Aut}(F_{n}\times F_{m}) be a type VII automorphism of Fn×FnF_{n}\times F_{n}. Then Φ\Phi is of the form (x,y)(yψ,xϕ)(x,y)\mapsto(y\psi,x\phi), for some ϕ,ψAut(Fn)\phi,\psi\in\text{Aut}(F_{n}). Again, by [10, Proposition 6.2], Φ\Phi is uniformly continuous with respect to the metric dd and, by uniqueness of the extension Φ^\widehat{\Phi} is given by (x,y)(yψ^,xϕ^)(x,y)\mapsto(y\widehat{\psi},x\widehat{\phi}). We start by proving that Φ^\widehat{\Phi} satisfies a Hölder condition. Notice that, since ϕ\phi and ψ\psi are uniformly continuous with respect to d1d_{1}, then ϕ^\widehat{\phi} and ψ^\widehat{\psi} satisfy a Hölder condition. Let (K1,r1)(K_{1},r_{1}) and (K2,r2)(K_{2},r_{2}) be the constants for ϕ^\widehat{\phi} and ψ^\widehat{\psi}, respectively.

We have that, for all (x,y),(z,w)Fn^×Fn^(x,y),(z,w)\in\widehat{F_{n}}\times\widehat{F_{n}},

d((x,y)Φ^,(z,w)Φ^)\displaystyle d((x,y)\widehat{\Phi},(z,w)\widehat{\Phi}) =d((yψ^,xϕ^),(wψ^,zϕ^))\displaystyle=d((y\widehat{\psi},x\widehat{\phi}),(w\widehat{\psi},z\widehat{\phi}))
=max{d1(yψ^,wψ^),d1(xϕ^,zϕ^)}\displaystyle=\max\{d_{1}(y\widehat{\psi},w\widehat{\psi}),d_{1}(x\widehat{\phi},z\widehat{\phi})\}
max{K2d1(y,w)r2,K1d1(x,z)r1}\displaystyle\leq\max\{K_{2}d_{1}(y,w)^{r_{2}},K_{1}d_{1}(x,z)^{r_{1}}\}
max{K1,K2}max{d1(y,w),d1(x,z)}min{r1,r2}\displaystyle\leq\max\{K_{1},K_{2}\}\max\{d_{1}(y,w),d_{1}(x,z)\}^{\min\{r_{1},r_{2}\}}
=max{K1,K2}d((x,y),(z,w))min{r1,r2}.\displaystyle=\max\{K_{1},K_{2}\}d((x,y),(z,w))^{\min\{r_{1},r_{2}\}}.

We remark that the last inequality holds since d1(x,z),d1(y,w)1d_{1}(x,z),d_{1}(y,w)\leq 1, by definition of the prefix metric.

Let (x,y)Fn^×Fm^(x,y)\in\widehat{F_{n}}\times\widehat{F_{m}} be a nonwandering point. Then, by Proposition 2.3, (x,y)(x,y) must be recurrent and so there is a sequence (ni)i(n_{i})_{i\in\mathbb{N}} such that (x,y)Φ^ni(x,y)(x,y)\widehat{\Phi}^{n_{i}}\to(x,y). If (ni)(n_{i}) has infinitely many even numbers, then consider the subsequence (nki)i(n_{k_{i}})_{i\in\mathbb{N}} of even numbers such that (x,y)Φ^nki(x,y)(x,y)\widehat{\Phi}^{n_{k_{i}}}\to(x,y). This means that

(x(ϕ^ψ^)nki2,y(ψ^ϕ^)nki2)(x,y),\left(x(\widehat{\phi}\widehat{\psi})^{\frac{n_{k_{i}}}{2}},y(\widehat{\psi}\widehat{\phi})^{\frac{n_{k_{i}}}{2}}\right)\to(x,y),

and so x(ϕ^ψ^)nki2xx(\widehat{\phi}\widehat{\psi})^{\frac{n_{k_{i}}}{2}}\to x and y(ψ^ϕ^)nki2yy(\widehat{\psi}\widehat{\phi})^{\frac{n_{k_{i}}}{2}}\to y, which implies that xx is ϕ^ψ^\widehat{\phi}\widehat{\psi}-recurrent and yy is ψ^ϕ^\widehat{\psi}\widehat{\phi}-recurrent. Since recurrent elements of extensions of free group endomorphisms must be periodic by [8, Corollary 5.11], then there are p1,p2p_{1},p_{2}\in\mathbb{N} such that x(ϕ^ψ^)p1=xx(\widehat{\phi}\widehat{\psi})^{p_{1}}=x and y(ψ^ϕ^)p2=yy(\widehat{\psi}\widehat{\phi})^{p_{2}}=y. Hence,

(x,y)Φ^2p1p2=(x(ϕ^ψ^)p1p2,y(ψ^ϕ^)p2)=(x,y),(x,y)\widehat{\Phi}^{2p_{1}p_{2}}=(x(\widehat{\phi}\widehat{\psi})^{p_{1}p_{2}},y(\widehat{\psi}\widehat{\phi})^{p_{2}})=(x,y),

that is (x,y)(x,y) is Φ^\widehat{\Phi}-periodic.

If (ni)(n_{i}) has only finitely many even numbers, then it must have infinitely many odd numbers and we consider the subsequence (nki)i(n_{k_{i}})_{i\in\mathbb{N}} of odd numbers such that (x,y)Φ^nki(x,y)(x,y)\widehat{\Phi}^{n_{k_{i}}}\to(x,y). This means that

(y(ψ^ϕ^)nki12ψ^,x(ϕ^ψ^)nki12ϕ^)(x,y),\left(y(\widehat{\psi}\widehat{\phi})^{\frac{n_{k_{i}}-1}{2}}\widehat{\psi},x(\widehat{\phi}\widehat{\psi})^{\frac{n_{k_{i}}-1}{2}}\widehat{\phi}\right)\to(x,y),

and so y(ψ^ϕ^)nki12ψ^xy(\widehat{\psi}\widehat{\phi})^{\frac{n_{k_{i}}-1}{2}}\widehat{\psi}\to x and x(ϕ^ψ^)nki12ϕ^yx(\widehat{\phi}\widehat{\psi})^{\frac{n_{k_{i}}-1}{2}}\widehat{\phi}\to y. Let ε>0\varepsilon>0. It follows that

NiNd1(x(ϕ^ψ^)nki12ϕ^,y)ε2\displaystyle\exists N\in\mathbb{N}\;\forall i\geq N\;\;d_{1}(x(\widehat{\phi}\widehat{\psi})^{\frac{n_{k_{i}}-1}{2}}\widehat{\phi},y)\leq\frac{\varepsilon}{2} (5)

and since (ϕ^ψ^)nkN12ϕ^(\widehat{\phi}\widehat{\psi})^{\frac{n_{k_{N}}-1}{2}}\widehat{\phi} is uniformly continuous, there is some δ\delta such that

zFn^d1(x,z)δd1(x(ϕ^ψ^)nkN12ϕ^,z(ϕ^ψ^)nkN12ϕ^)ε2\displaystyle\forall z\in\widehat{F_{n}}\;\;d_{1}(x,z)\leq\delta\implies d_{1}(x(\widehat{\phi}\widehat{\psi})^{\frac{n_{k_{N}}-1}{2}}\widehat{\phi},z(\widehat{\phi}\widehat{\psi})^{\frac{n_{k_{N}}-1}{2}}\widehat{\phi})\leq\frac{\varepsilon}{2} (6)

Since y(ψ^ϕ^)nki12ψ^xy(\widehat{\psi}\widehat{\phi})^{\frac{n_{k_{i}}-1}{2}}\widehat{\psi}\to x,

N2iN2d1(y(ψ^ϕ^)nki12ψ^,x)δ,\displaystyle\exists N_{2}\in\mathbb{N}\;\forall i\geq N_{2}\;\;d_{1}(y(\widehat{\psi}\widehat{\phi})^{\frac{n_{k_{i}}-1}{2}}\widehat{\psi},x)\leq\delta, (7)

which, by (6) implies that

iN2d1(y(ψ^ϕ^)nki12ψ^(ϕ^ψ^)nkN12ϕ^,x(ϕ^ψ^)nkN12ϕ^)ε2.\displaystyle\forall i\geq N_{2}\;\;d_{1}(y(\widehat{\psi}\widehat{\phi})^{\frac{n_{k_{i}}-1}{2}}\widehat{\psi}(\widehat{\phi}\widehat{\psi})^{\frac{n_{k_{N}}-1}{2}}\widehat{\phi},x(\widehat{\phi}\widehat{\psi})^{\frac{n_{k_{N}}-1}{2}}\widehat{\phi})\leq\frac{\varepsilon}{2}. (8)

Combining (5) and (8), we get that

iN2d1(y(ψ^ϕ^)nki+nkN2,y)ε,\displaystyle\forall i\geq N_{2}\;\;d_{1}(y(\widehat{\psi}\widehat{\phi})^{\frac{n_{k_{i}}+n_{k_{N}}}{2}},y)\leq\varepsilon,

thus yy is (ψ^ϕ^)(\widehat{\psi}\widehat{\phi})-recurrent and so (ψ^ϕ^)(\widehat{\psi}\widehat{\phi})-periodic. Analogously, we deduce that xx is (ϕ^ψ^)(\widehat{\phi}\widehat{\psi})-periodic. Proceeding as above, we obtain that (x,y)(x,y) is Φ^\widehat{\Phi}-periodic. ∎

Corollary 5.4.

The continuous extension of every automorphism of Fn×FmF_{n}\times F_{m} to the completion obtained by taking the prefix metric in each factor has asymptotically periodic dynamics.

Acknowledgements

This work is funded by national funds through the FCT - Fundação para a Ciência e a Tecnologia, I.P., under the scope of the projects UIDB/00297/2020 and UIDP/00297/2020 (Center for Mathematics and Applications).

References

  • [1] V. Araújo and P. V. Silva. Hölder conditions for endomorphisms of hyperbolic groups. Comm. Algebra, 44(10):4483–4503, 2016.
  • [2] M. Bestvina and M. Handel. Train tracks and automorphisms of free groups. Ann. Math., 135:1–51, 1992.
  • [3] O. Bogopolski, A. Martino, O. Maslakova, and E. Ventura. The conjugacy problem is solvable in free-by-cyclic groups. Bull. London Math. Soc., 38:787–794, 2006.
  • [4] B. H. Bowditch. Notes on Gromov’s hyperbolicity criterion for path-metric spaces. In A.Verjovsky E.Ghys, A.Haefliger, editor, Group theory from a geometrical viewpoint, pages 64–167. World Scientific, 1991.
  • [5] B. H. Bowditch. Coarse median spaces and groups. Pacific J. Math., 261(1):53–93, 2013.
  • [6] B. H. Bowditch. Median algebras. preprint, available at http://homepages.warwick.ac.uk/ masgak/papers/median-algebras.pdf, 2022.
  • [7] P. Brinkmann. Detecting automorphic orbits in free groups. J. Algebra, 324(5):1083–1097, 2010.
  • [8] A. Carvalho. On the dynamics of extensions of free-abelian times free groups endomorphisms to the completion. Proc. Edinb. Math. Soc., 65(3):705–735, 2022.
  • [9] A. Carvalho. On uniformly continuous endomorphisms of hyperbolic groups. J. Algebra, 602:197–223, 2022.
  • [10] A. Carvalho. On endomorphisms of the direct product of two free groups. J. Group Theory, 26(4):693–724, 2023.
  • [11] A. Carvalho. On generalized conjugacy and some related problems. Comm. Algebra, 51(8):3528–3542, 2023.
  • [12] A. Carvalho. On the intersection of fixed subgroups of Fn×Fm{F}_{n}\times{F}_{m}. preprint, arXiv: 2306.12533, 2023.
  • [13] A. Carvalho. Quantifying Brinkmann’s problem: φ\varphi-order and φ\varphi-spectrum. preprint, arXiv: 2306.12563, 2023.
  • [14] A. Carvalho. Structure, algorithmics and dynamics of endomorphisms for certain classes of groups. PhD thesis, University of Porto, 2023.
  • [15] A. Carvalho and J. Delgado. Homomorphisms of free-abelian-by-free groups. in preparation, 2023.
  • [16] J. Cassaigne and P. V. Silva. Infinite periodic points of endomorphisms over special confluent rewriting systems. Ann. Inst. Fourier, 59.2:769–810, 2009.
  • [17] J. Cassaigne and P. V. Silva. Infinite words and confluent rewriting systems: endomorphism extensions. Int. J. Algebra Comput., 19.4:443–490, 2009.
  • [18] I. Chatterji, T. Fernós, and A. Iozzi. The median class and superrigidity of actions on CAT(0) cube complexes. J. Topol., 9(2):349–400, 2016.
  • [19] G. Levitt D. Gaboriau, A. Jaeger and M. Lustig. An index for counting fixed points of automorphisms of free groups. Duke Math. J., 93:425–452, 1998.
  • [20] J. Delgado and E. Ventura. Algorithmic problems for free-abelian times free groups. J. Algebra, 263(1):256–283, 2013.
  • [21] J. Dugundji. Topology. Boston, Mass.-London-Sydney: Allyn and Bacon, Inc. Reprinting of the 1966 original, Allyn and Bacon Series in Advanced Mathematics, 1978.
  • [22] E. Fioravanti. Roller boundaries for median spaces and algebras. Algebr. Geom. Topol., 20(3):1325–1370, 2020.
  • [23] E. Fioravanti. Coarse-median preserving automorphisms. to appear in Geom. Topol., 2022.
  • [24] R. Kannan and R. Lipton. Polynomial-time algorithm for the orbit problem. J. Assoc. Comput. Mach., 33(4), 1986.
  • [25] G. Levitt and M. Lustig. Automorphisms of free groups have asymptotically periodic dynamics. J. Reine Angew. Math., 619:1–36, 2008.
  • [26] A. D. Logan. The conjugacy problem for ascending HNN-extensions of free groups. preprint, arXiv:2209.04357, 2022.
  • [27] K. A. Mihailova. The occurrence problem for direct products of groups. Dokl. Acad. Nauk SSRR, 119:1103–1105, 1958.
  • [28] G. A. Niblo, N. Wright, and J. Zhang. A four point characterisation for coarse median spaces. Groups Geom. Dyn, 13(3):651–662, 2019.
  • [29] M. A. Roller. Poc sets, median algebras and group actions. an extended study of Dunwoody’s construction and Sageev’s theorem. preprint, 1998.
  • [30] M. Sapir and Iva S̆pakulová. Almost all one-relator groups with at least three generators are residually finite. J. Eur. Math. Soc., 13(2):331–343, 2011.
  • [31] P. V. Silva. Fixed points of endomorphisms over special confluent rewriting systems. Monatsh. Math., 161.4:417–447, 2010.
  • [32] P. V. Silva. Fixed points of endomorphisms of virtually free groups. Pacific J. Math., 263(1):207–240, 2013.
  • [33] E. Ventura. The multiple endo-twisted conjugacy problem is solvable in finitely generated free groups. preprint, available at https://enric-ventura.staff.upc.edu/ventura/files/68t.pdf, 2021.
  • [34] S. Willard. General Topology. Addison-Wesley, Reading, Mass., 1970.