June 30, 2019
In this article we prove a necessary and a sufficient condition for a finite subset of the special linear group to be dominated. These conditions are purely geometric in nature, as they only involve the trace and the eigenvectors of the matrices, and can be computed explicitly. Our sufficient condition, in particular, provides a simple algorithm for constructing a dominated set with prescribed eigenvectors. The techniques involved in our proofs take advantage of the interaction between dominated sets and two-dimensional hyperbolic geometry.
The notion of a dominated splitting for a dynamical system was introduced by Mañe [@Ma1978; @Ma1984] in his seminal work on the \(C^1\)-stability Conjecture, as a weaker version of the notion of hyperbolicity introduced by Smale. It has since appeared in various areas of the theory of dynamical systems, such as the continuity of Lyapunov exponents [@BoVi2005], differentiable dynamics [@BoDiPu2003], Anosov Representations [@BoPoSa2019], control theory [@CoKl2000] and the continuity properties of the lower spectral radius [@BoMo2015]. A detailed overview of dominated splittings, often simply called domination, can be found in the survey [@Sa2016].
The main focus of this article is dominated subsets of \(\mathrm{SL}(2,\mathbb{R})\). The two-dimensional case is particularly interesting, since domination is linked with dimension theory [@ChJu2021; @SoTa2021] and the spectral theory of Schroedinger operators [@DaFiGo2022; @Wo2022]; applications unique to the setting of \(\mathrm{SL}(2,\mathbb{R})\).
Let us start by providing the definition of domination in this setting:
Definition 1. A finite set \(\mathcal{A}=\{A_1,A_2,\dots,A_N\}\subseteq\mathrm{SL}(2,\mathbb{R})\) will be called dominated if there exist constants \(C>0\) and \(\lambda>1\), so that \[\lVert A_{i_n}A_{i_{n-1}}\cdots A_{i_1}\rVert \geqslant C \lambda^n, \quad \text{for any sequence}\;(i_j)\subseteq\{1,2,\dots,N\}\;\text{and all}\;n\in\mathbb{N}.\]
This particular characterisation of domination—that is best suited for to purposes—was obtained by Yoccoz in [@Yo2004] in his study of the hyperbolicity properties of linear cocycles. Also, in this two-dimensional setting domination is often referred to as uniform hyperbolicity (see, for example, [@AvBoYo2010; @Yo2004] and references therein).
Despite its simplicity, determining whether the set \(\mathcal{A}\) is dominated using Definition 1 is a difficult endeavour. To alleviate some of these difficulties, Avila, Bochi and Yoccoz [@AvBoYo2010] showed that domination is equivalent to the fact that \(\mathcal{A}\) acts as a uniformly contracting iterated function system on a part of the projective real line \(\mathbb{R}\mathbb{P}^1\) (see Theorem 4 for the precise statement). A higher-dimensional analogue of this condition was obtained by Bochi and Gourmelon [@BoGo2009], and was further explored by Barnsley and Vince [@BaVi2012]. Recently, these ideas have been adapted to the setting of operators on Banach spaces by Blumenthal and Morris [@BlMo2019], and Quas, Thieullen and Zarrabi [@QTZ2019].
Inspired by this geometric characterisation, we aim at providing a necessary and a sufficient condition for domination that explicitly features the geometry of the action of a pair of matrices on \(\mathbb{R}\mathbb{P}^1\).
First, we recall that a non-identity matrix \(A\in\mathrm{SL}(2,\mathbb{R})\) is called hyperbolic if \(\lvert \textrm{tr}(A)\rvert >2\), parabolic if \(\lvert \textrm{tr}(A)\rvert =2\), and elliptic otherwise. A hyperbolic matrix \(A\) has two eigenvalues \(\lvert\lambda_u(A)\rvert>\lvert\lambda_s(A)\rvert\), with distinct corresponding eigenvectors \(u(A),s(A)\in \mathbb{R}\mathbb{P}^1\). On the other hand, if \(A\) is parabolic then it has a unique eigenvector with algebraic multiplicity two, in which case we write \(u(A)=s(A)\).
Throughout the article, we identify \(\mathbb{R}\mathbb{P}^1\) with the extended real line \(\mathbb{R}\cup\{\infty\}\). With this identification in mind, for a pair of hyperbolic matrices \(A,B\in\mathrm{SL}(2,\mathbb{R})\) we define their cross-ratio as \[C(A,B)=\frac{u(A)-u(B)}{u(A)-s(B)}\frac{s(A)-s(B)}{s(A)-u(B)},\] with the standard conventions about the point at infinity. Simple arguments show that if \(A,B\) are hyperbolic matrices with pairwise distinct fixed points then \(C(A,B)\) is well-defined (i.e. both denominators are non-zero) and does not equal zero or one.
The cross-ratio, which has been previously considered by Jacques and Short [@JS], is a simple way to quantify the geometric configuration of the pair \(A,B\). For example, if \(C(A,B)\) is well-defined and \(C(A,B)<1\), then \(A\) and \(B\) exhibit behaviour similar to positive matrices, in that they map a non-trivial open and connected subset of \(\mathbb{R}\mathbb{P}^1\) compactly inside itself, meaning that \(\{A,B\}\) is dominated due to the aforementioned result of Avila, Bochi and Yoccoz (see Section 2.3 for more details).
Our first result provides an algorithm for constructing dominated sets, that depends on evaluating the cross-ratios of pairs.
Theorem 1. Let \(u_1,u_2,\dots,u_N,s_1,s_2,\dots s_N\in\mathbb{R}\mathbb{P}^1\) be \(2N\) pairwise distinct points in the projective real line. Suppose that \(A_1,A_2,\dots A_N\in\mathrm{SL}(2,\mathbb{R})\) are hyperbolic matrices with \(u(A_i)=u_i\) and \(s(A_i)=s_i\), for \(i=1,2,\dots,N\), and consider the constants \[M_{i,j} = \max\left\{1+\left\lvert C(A_i,A_j)\right\rvert,1+\frac{1}{\left\lvert C(A_i,A_j)\right\rvert}\right\},\] for all pairs \(i,j\in\{1,2,\dots,N\}\) with \(i\neq j\). If \[\label{eq:32MAIN} \tfrac{1}{2}\lvert \textrm{tr}(A_k)\rvert > \max_{i\neq j}\{M_{i,j}\}\cdot \max_{i\neq j}\left\{\frac{\lvert C(A_i,A_j)\rvert +1}{\lvert C(A_i,A_j)-1\rvert}\right\}, \quad \text{for all}\;k=1,2,\dots,N,\tag{1}\] then the set \(\{A_1,A_2,\dots, A_N\}\) is dominated.
To appreciate Theorem 1, observe that each hyperbolic matrix \(A\) is determined by three independent parameters. Two of these are its eigenvectors, and the third is its contraction rate; i.e. a measure of how much \(A\) shifts the projective space towards its fixed point \(u(A)\). This contraction rate is typically quantified by \(\lvert \textrm{tr}(A)\rvert\) (see also Section 2.2 for a more geometric description).
So, Theorem 1 tells us that if we wanted to construct an example of a dominated set \(\{A_1,A_2,\dots,A_n\}\), where each matrix \(A_i\) has some prescribed eigenvectors \(u_i,s_i\), then we only need increase the contraction rates of all \(A_1,A_2,\dots,A_n\) enough to satisfy 1 . In particular, the algebraic simplicity of the constants in 1 makes Theorem 1 an efficient tool for constructing explicit dominated examples, which are quite rare in the literature.
In order to prove Theorem 1 we establish certain constraints on the traces of a dominated pair \(\{A,B\}\). Our analysis leads us to the following dichotomy, which shows that under a certain threshold the dynamical behaviour of \(\{A,B\}\) is either very rigid, or quite chaotic.
Theorem 2. Assume that \(A,B\in\mathrm{SL}(2,\mathbb{R})\) are hyperbolic matrices such that \(C(A,B)\) is well-defined and satisfies \(C(A,B)>1\). If \[\max\left\{\tfrac{1}{2}\lvert \mathrm{tr}(A)\rvert, \tfrac{1}{2}\lvert \mathrm{tr}(B)\rvert\right\}\leqslant\;\frac{5\;C(A,B)-1}{3\;C(A,B)+1},\] then the semigroup generated by \(\{A,B,-A,-B\}\) is either a discrete subgroup of \(\mathrm{SL}(2,\mathbb{R})\), or dense in \(\mathrm{SL}(2,\mathbb{R})\).
Definition 1 immediately shows that if a finite set \(\mathcal{A}\) is dominated, then the semigroup (under matrix multiplication) generated by \(\mathcal{A}\) contains only hyperbolic matrices and is a discrete subset of \(\mathrm{SL}(2,\mathbb{R})\). Therefore, a direct corollary of Theorem 2 is the following necessary condition for domination.
Corollary 1. If the set \(\{A,B\}\subseteq\mathrm{SL}(2,\mathbb{R})\) is dominated, then \(C(A,B)\) is well-defined, and either \(C(A,B)<1\) or \[\max\left\{\tfrac{1}{2}\lvert \mathrm{tr}(A)\rvert, \tfrac{1}{2}\lvert \mathrm{tr}(B)\rvert\right\}> \;\frac{5\;C(A,B)-1}{3\;C(A,B)+1}.\]
Corollary 1 is in the same spirit as Jörgensen’s inequality [@Jo1976] from the theory of Kleinian groups, as it quantifies the the intuitive fact that if a pair is dominated then the contraction rates of the matrices cannot both be very small, unless the geometry of their action is somewhat trivial (i.e. \(C(A,B)<1\)).
Jörgensen’s inequality has been extended to a large class of matrix sets, that contains all dominated sets, by Jacques and Short [@JS] and was later used in [@Ch2022] for the study of the boundary of domination. It is also worth mentioning that a condition fulfilling a similar purpose was obtained by Avila, Bochi and Yoccoz in [@AvBoYo2010]. The main advantage of Corollary 1 is its very explicit dependence on the geometric configuration of \(\{A,B\}\), as is evident by the use of the cross-ratio.
Finally, our techniques allow us to obtain a necessary and sufficient condition for domination in the special case of a pair of matrices with equal traces.
Theorem 3. Assume that \(A,B\in\mathrm{SL}(2,\mathbb{R})\) are hyperbolic matrices such that \(C(A,B)\) is well-defined, \(C(A,B)>1\) and \(\lvert \textrm{tr}(A) \rvert =\lvert \textrm{tr}(B) \rvert=t\). Then \(\{A,B\}\) is dominated if and only if \(\tfrac{1}{2}t>\sqrt{C(A,B)}\).
In order to prove our results, we employ techniques from two-dimensional hyperbolic geometry in the complex plane. In particular, we view matrices of \(\mathrm{SL}(2,\mathbb{R})\) as Möbius transformations acting on the hyperbolic plane, and study how this action affects their domination properties. This perspective allows us to better control the geometric configuration of a finite set of matrices in order to obtain the explicit bounds present in our theorems.
The article is structured as follows: In the second section we introduce the concepts necessary for our work (Möbius transformations and hyperbolic geometry) and thoroughly analyse cross-ratios. Theorem 2 is proved in Section 3. Section 4 is dedicated to to establishing certain constraints on pairs of matrices that guarantee domination, and contains the proof of Theorem 3. The final section of this article is dedicated to the proof of our example-constructing method, Theorem 1.
As mentioned in the introduction, we identify the real projective line \(\mathbb{R}\mathbb{P}^1\) with the extended real line \(\overline{\mathbb{R}}\vcentcolon=\mathbb{R}\cup\{\infty\}\). We will frequently refer to intervals of the extended real line, which are the usual intervals of \(\mathbb{R}\) along with sets of the form \((\infty,a)\cup(b,\infty)\), for some \(a\leqslant b\), and usual modifications to include some (or all) endpoints.
With this convention \(\mathrm{SL}(2,\mathbb{R})\) acts on \(\mathbb{R}\mathbb{P}^1\) as Möbius transformations with real entries. That is, we have the following correspondence \[\mathrm{PSL}(2,\mathbb{R})\ni \begin{pmatrix} a & b \\ c & d \end{pmatrix} \mapsto \frac{ax+b}{cx+d}.\] The use of the group \(\mathrm{PSL}(2,\mathbb{R})=\mathrm{SL}(2,\mathbb{R})/\{\pm I\}\) instead of \(\mathrm{SL}(2,\mathbb{R})\) is justified by the fact that the matrices \(A\) and \(-A\) induce the same Möbius transformation. So for the rest of this article we will consider finite sets of Möbius transformations \(\{f_1,f_2,\dots, f_n\}\subseteq\mathrm{PSL}(2,\mathbb{R})\) instead of matrices. With this correspondence, the product of two matrices \(A_1\cdot A_2\) corresponds to the composition \(f_1\circ f_2\), where \(f_1\) is the Möbius transformation induced by \(A_1\) and similarly for \(f_2\).
A key point of our analysis in the characterisation of domination in terms of the projective action due to Avila, Bochi and Yoccoz [@AvBoYo2010], which with our notation can be written as follows.
Theorem 4. A set \(\{f_1,f_2,\dots,f_n\}\subseteq\mathrm{PSL}(2,\mathbb{R})\) is dominated if and only if there exists a finite union \(X\subsetneq\overline{\mathbb{R}}\) of open intervals, with disjoint closures, such that \(\overline{f_i(X)}\subseteq X\), for all \(i=1,2,\dots,n\).
One of the main advantages of considering \(\mathrm{PSL}(2,\mathbb{R})\) as a group of Möbius transformations is that its action on \(\overline{\mathbb{R}}\) can be extended to the upper half-plane of \(\mathbb{H}=\{z\in\mathbb{C}\colon \mathrm{Im}\;z >0\}\) of the complex plane. If we now equip \(\mathbb{H}\) with the hyperbolic distance \(d_\mathbb{H}\), i.e. the distance induced by the Riemannian metric \(\lvert dz \rvert/\relax{z}\), then \(\mathrm{PSL}(2,\mathbb{R})\) is exactly the group of conformal isometries of the complete metric space \((\mathbb{H},d_\mathbb{H})\). A thorough analysis of the hyperbolic geometry of \(\mathbb{H}\), along with most of the material presented in this subsection, can be found in [@Be1995].
One can easily compute the following closed formula for \(d_\mathbb{H}\) (see, for example, [@Be1995]):
\[\label{eq:32hyperbolic32metric32in32H} d_\mathbb{H}(z_1,z_2)=\log\frac{\lvert z_1 - \overline{z_2}\rvert + \lvert z_1- z_2\rvert}{\lvert z_1 - \overline{z_2}\rvert - \lvert z_1- z_2\rvert},\quad \text{for all}\;z_1,z_2\in\mathbb{H},\tag{2}\]
or, equivalently
\[\label{dist-formula} \tanh\tfrac{1}{2}d_\mathbb{H}(z_1,z_2) = \frac{\lvert z_1-z_2\rvert}{\lvert z_1-\overline{z_2}\rvert}.\tag{3}\]
The geodesics of the hyperbolic distance \(d_\mathbb{H}\) are the vertical half-lines and semicircles orthogonal to \(\mathbb{R}\). We say that two hyperbolic geodesics are disjoint if they do not meet in \(\mathbb{H}\) and have distinct endpoints. The distance between two distinct geodesics \(\gamma_1\) and \(\gamma_2\) is defined as \[d_\mathbb{H}(\gamma_1,\gamma_2)\vcentcolon=\inf\{d_\mathbb{H}(z_1,z_2)\colon z_1\in\gamma_1,\;z_2\in\gamma_2\}.\]
Now suppose that \(f(x)=(ax+b)/(cx+d)\) is the Möbius transformation induced by a matrix \(A\in\mathrm{SL}(2,\mathbb{R})\). We define the trace of \(f\) as \(\lvert\textrm{tr}(f) \rvert \vcentcolon=\lvert \textrm{tr}(A)\rvert = \lvert a+ d\rvert\). The use of the absolute value here is important since it alleviates any ambiguity in the sign. With this definition the classification of matrices to hyperbolic, parabolic and elliptic we mentioned in the introduction extends naturally to Möbius transformations. That is, \(f\) is called hyperbolic if the inducing matrix \(A\) is hyperbolic, and similarly for parabolic and elliptic transformations.
An equivalent classification can be obtained by considering the fixed points in \(\mathbb{H}\cup\overline{\mathbb{R}}\) of the transformation. So, a non-identity Möbius transformation is hyperbolic, parabolic or elliptic depending on whether it has two fixed points \(\overline{\mathbb{R}}\), one fixed point in \(\overline{\mathbb{R}}\), or one fixed point in \(\mathbb{H}\), respectively.
Suppose that \(f\) is a hyperbolic transformation in \(\mathrm{PSL}(2,\mathbb{R})\). The fixed points of \(f\) will be denoted by \(u(f),s(f)\) and are exactly the projections in \(\overline{\mathbb{R}}\) of the eigenvectors of the matrix inducing \(f\) (hence the similar notation). If we let \(f^n\vcentcolon=f\circ f\circ \dots \circ f\) denote the \(n\)-th iterate of \(f\), i.e. the composition of \(f\) with itself \(n\)-times, then the fixed point \(u(f)\) of \(f\) has the property that the sequence \((f^n)\) converges to the constant \(u(f)\) uniformly on compact subsets of \(\mathbb{H}\). We thus call \(u(f)\) the attracting fixed point of \(f\). The other fixed point is called the repelling fixed point.
The unique hyperbolic geodesic joining the fixed points \(u(f),s(f)\) of a hyperbolic transformation \(f\) is called the axis of \(f\) and is denoted by \(\mathrm{Ax}(f)\). Since \(f\) is an isometry of \(d_\mathbb{H}\), the curve \(\mathrm{Ax}(f)\) is preserved set-wise by \(f\).
All our results require a quantifiable way to measure the contraction rate of hyperbolic transformation. The prevalent method to do so is the translation length \(T_f\) of a hyperbolic transformation \(f\), defined as follows \[T_f\vcentcolon=\inf\{d_\mathbb{H}(z,f(z))\colon z\in \mathbb{H}\}.\] This infimum is in fact a minimum that is attained for any point \(z_0\in\mathrm{Ax}(f)\). The translation length is directly related to the trace of the transformation with the following formula, found in [@Be1995]. \[\label{eq:32trace32and32translation32length} \cosh\tfrac{1}{2}T_f= \tfrac{1}{2}\lvert \textrm{tr}(f)\rvert,\tag{4}\] meaning that a simple way to quantify the contraction rate is by the trace, as we mentioned in the introduction.
Furthermore, observe that since \(f\) fixes \(\mathrm{Ax}(f)\) set-wise, \(f^n(z_0)\in\mathrm{Ax}(f)\) for any \(z_0\in\mathrm{Ax}(f)\) and all \(n\in\mathbb{N}\). Using the facts that \(\mathrm{Ax}(f)\) is a hyperbolic geodesic and \(f\) is an isometry we get that \[\begin{align} T_{f^n}&=d_\mathbb{H}(f^n(z_0),z_0)=d_\mathbb{H}(f^n(z_0),f^{n-1}(z_0))+d_\mathbb{H}(f^{n-1}(z_0),f^{n-2}(z_0))+\cdots+d_\mathbb{H}(f(z_0),z_0)\nonumber\\ &=d_\mathbb{H}(f(z_0),z_0)+d_\mathbb{H}(f(z_0),z_0)+\cdots+d_\mathbb{H}(f(z_0),z_0)=n\;T_f. \label{eq:32translation32length32of32iterates} \end{align}\tag{5}\]
We will often find it convenient to switch to the disc model \(\mathbb{D}=\{z\in\mathbb{C}\colon \lvert z \rvert<1\}\) of hyperbolic space. In \(\mathbb{D}\) the hyperbolic distance is induced by the metric \(\frac{2\lvert dz \rvert}{1-\lvert z\rvert^2}\). Since the half-plane and the disc models are conformally equivalent, we can move freely between them and will do so without explicit mention. In particular, we use the disc model for most of our figures where a hyperbolic transformation will be portrayed by drawing its axis as a directed hyperbolic geodesic pointing towards its attracting fixed point as in Figure 1.
Here we explore how the cross-ratio of two hyperbolic transformations can be used to determine their geometric configuration. First, recall that the cross-ratio of two hyperbolic transformations \(f,g\in\mathrm{PSL}(2,\mathbb{R})\) is given by \[\label{eq:32cross-ratio32def} C(f,g)\vcentcolon=[u(f),s(f),u(g),s(g)]=\frac{u(f)-u(g)}{u(f)-s(g)}\frac{s(f)-s(g)}{s(f)-u(g)}.\tag{6}\] If any of the fixed points of \(f\) and \(g\) is the point at infinity, then we simply omit any terms of the above fractions that contain it. Note that the cross-ratio is well-defined only when \(u(f)\neq s(g)\) and \(u(g)\neq s(f)\). If, however, any of these two conditions does not hold, then it is easy to show that the pair \(\{f,g\}\) is not dominated (see, for example, [@JS]). So, since we are only interested in the domination properties of sets of transformations, we always assume that the cross-ratio of a pair is well-defined.
Simple calculations show that \(C(f,g)=1\) if and only if one of \(f,g\) is parabolic. Therefore, for any two hyperbolic transformations, we have that \(C(f,g)\neq 1\). Furthermore, we have the following standard properties for the cross-ratio: \(C(f,g)=C(g,f)\) and \(C(f^{-1},g)\cdot C(f,g)=1\).
Now, let us establish formulas that relate the cross-ratio \(C(f,g)\) of two hyperbolic transformations \(f\) and \(g\) with the geometric configuration of their axes. Equivalent versions of the formulae in Lemma 1, to follow, appear in [@Be1995], but we include a proof for the sake of completeness, and since none is provided in [@Be1995].
We first need to mention that if the axes of two hyperbolic transformations \(f\) and \(g\) cross at a point \(p\in\mathbb{H}\), we define the angle \(\theta\in[0,\pi]\) between \(\mathrm{Ax}(f)\) and \(\mathrm{Ax}(g)\) to be the angle at \(p\) of the hyperbolic triangle defined by \(\alpha(f),\alpha(g)\) and \(p\). In this case, we say that the axes of \(f\) and \(g\) cross at an angle \(\theta\).
Lemma 1. Suppose that \(f\) and \(g\) are hyperbolic transformations.
If the axes of \(f\) and \(g\) cross at an angle \(\theta\), then \(C(f,g)=-\tan^2\tfrac{1}{2}\theta\).
If the axes of \(f\) and \(g\) are disjoint and a hyperbolic distance \(d\) apart, then \[C(f,g)= \begin{cases} \tanh^2\tfrac{1}{2}d, &\text{if}\quad 0<C(f,g)<1,\\[5pt] \coth^2\tfrac{1}{2}d, &\text{if}\quad C(f,g)>1. \end{cases}\]
Proof. For part (i), suppose that the axes of \(f\) and \(g\) cross at an angle \(\theta\). Conjugate \(f\) and \(g\) by a transformation \(\phi\in\mathrm{PSL}(2,\mathbb{C})\) so that they act on the unit disc, their axes meet at the origin and the diameter through \(i\) and \(-i\) dissects the angle \(\theta\) (see Figure 2 on the left). If we denote \(F=\phi\circ f \circ \phi^{-1}, G=\phi\circ g \circ \phi^{-1}\), then \(C(F,G)=C(f,g)\) by the invariance of the cross-ratio. Due to the symmetry of this configuration we have that \(u(F)=\overline{s(G)},\;u(F)=-s(F)\) and \(u(G)=-s(G)\). Hence, \[C(F,G)=\frac{u(F)-u(G)}{u(F)-s(G)}\frac{s(F)-s(G)}{s(F)-u(G)}= \frac{-2\relax s(F)}{-2i\relax s(F)} \frac{2\relax s(F)}{2i\relax s(F)}= -\left(\frac{\relax s(F)}{\relax s(F)}\right)^2.\] The law of cosines applied to the Euclidean triangles with vertices \(0, u(F),s(G)\) and \(0,u(F),u(G)\), yield \(\left(2\relax u(F)\right)^2 = 2(1-\cos\theta )\) and \(\left(2\relax u(F)\right)^2=2(1-\cos(\pi-\theta))\), respectively. Therefore, \[C(F,G) = -\frac{1-\cos\theta}{1-\cos(\pi-\theta)}=-\frac{1-\cos\theta}{1+\cos\theta}=-\tan^2\tfrac{1}{2}\theta.\] Moving on to part (ii), assume that the axes of \(f\) and \(g\) are disjoint and a hyperbolic distance \(d\) apart. Since \(C(f^{-1},g)=1/C(f,g)\) and \(C(f,g)\neq 1\), it suffices to show that \(C(f,g)=\tanh^2\tfrac{1}{2}d\) whenever \(0<C(f,g)<1\). Conjugate \(f\) and \(g\) in \(\mathrm{PSL}(2,\mathbb{R})\) so that \(f\) fixes \(-\sigma, \sigma\) and \(g\) fixes \(-\lambda, \lambda\), for some \(0<\sigma<\lambda\), as shown in Figure 2 on the right. Then, \[C(f,g)=\frac{(\lambda-\sigma)^2}{(\lambda+\sigma)^2}.\] Since the imaginary half-axis is perpendicular to both \(\textrm{Ax}(f)\) and \(\textrm{Ax}(g)\), we have that \(d=\rho_{\mathbb{H}}(i\sigma,i\lambda)\). Using formula 3 yields \[\tanh^2\tfrac{1}{2}d=\tanh^2\tfrac{1}{2}\rho_{\mathbb{H}}(i\sigma,i\lambda)=\frac{\lvert i\sigma - i \lambda \rvert^2}{\lvert i\sigma -\overline{i\lambda} \rvert^2} = \frac{(\lambda-\sigma)^2}{(\lambda+\sigma)^2}=C(f,g),\] completing our proof. ◻
Note that the formulas in Lemma 1 yield a classification for the geometric configuration of the axes of two hyperbolic transformations, as shown in Figure 3. This figure, along with the geometric characterisation of domination in Theorem 4, immediately show that if \(C(f,g)<1\), then \(\{f,g\}\) is dominated, as we mentioned in the introduction.
Using the standard formulas for the trigonometric and hyperbolic trigonometric functions we can easily obtain the next formulas that will prove useful to our calculations.
Corollary 2. Suppose that \(f\) and \(g\) are hyperbolic transformations.
If the axes of \(f\) and \(g\) cross at an angle \(\theta\), then \(\cos^2\tfrac{1}{2}\theta=\frac{1}{1-C(f,g)}\).
If the axes of \(f\) and \(g\) are disjoint and a hyperbolic distance \(d\) apart, then \(\cosh d = \frac{C(f,g)+1}{\lvert C(f,g)-1\rvert}\).
In this section we prove Theorem 2 which immediately yields the necessary condition for matrix domination given in Corollary 1.
The main ingredient for the proof is the following lemma, which shows that if the traces of two transformations are small, relative to their geometric configuration, then the semigroup they generate contains elliptic transformations.
Lemma 2. Let \(f,g\) be two hyperbolic transformations \(C(f,g)>1\). If \[\max\left\{\tfrac{1}{2}\lvert \mathrm{tr}(f)\rvert, \tfrac{1}{2}\lvert \mathrm{tr}(g)\rvert\right\}\leqslant\;\frac{5C(f,g)-1}{3C(f,g)+1},\] then there exist \(n,m\in\mathbb{N}\) so that \(f^n\circ g^m\) is elliptic.
Proof. Since \(C(f,g)>1\), the axes \(\mathrm{Ax}(f)\) and \(\mathrm{Ax}(g)\), of \(f\) and \(g\) respectively, are disjoint. Let \(d>0\) be the hyperbolic distance between \(\mathrm{Ax}(f)\) and \(\mathrm{Ax}(g)\). The
proof revolves around evaluating the trace of the composition of \(f\) and \(g\), which is given by the following equation found in [@Be1995]: \[\label{eq:32antiparallel32trace32eq}
\frac{1}{2}\lvert\text{tr}(f\circ g)\rvert=\lvert\cosh d\sinh\tfrac{1}{2}T_f\sinh\tfrac{1}{2}T_g-\cosh\tfrac{1}{2}T_f\cosh\tfrac{1}{2}T_g\rvert,\tag{7}\] where \(T_f,T_g\) are the translation lengths of \(f\) and \(g\), respectively. Let \(\mathbb{R}_+^2\) be the first quadrant of \(\mathbb{R}^2\), endowed with the Euclidean
topology, and consider the function \(h\colon \mathbb{R}_+^2\to\mathbb{R}\), with \[\label{eq:32antiparallel32def32of32h}
h(x,y)=\cosh d\sinh x\sinh y-\cosh x\cosh y.\tag{8}\] So, equation 7 can be rewritten as \(\frac{1}{2}\lvert\text{tr}(f\circ g)\rvert = \rvert
h\left(\tfrac{1}{2}T_f,\tfrac{1}{2}T_g\right)\rvert\). Also, from 5 we have that \(T_{f^n}=n\;T_f\), meaning that \[\label{n44m-traces}
\frac{1}{2}\lvert\text{tr}(f^n\circ g^m)\rvert=\left\lvert h\left(\tfrac{n}{2}T_f,\tfrac{m}{2}T_g\right) \right\rvert, \quad \text{for all}\;n,m\in\mathbb{N}.\tag{9}\] Define the open set \(D=\left\{(x,y)\in\mathbb{R}_+^2\colon -1<h(x,y)<1\right\}\). Since a transformation \(h\) is elliptic if and only if \(\lvert \mathrm{tr}(h)\rvert <2\),
we have that the composition \(f^n\circ g^m\) is elliptic, for some \(n,m\in\mathbb{N}\), if and only if \(\left\lvert h\left(\tfrac{n}{2}T_f,\tfrac{m}{2}T_g\right)
\right\rvert<1\), or equivalently \(\left(\tfrac{n}{2}T_f,\tfrac{m}{2}T_g\right)\in D\).
So, our goal is to show that if \[\max\left\{\tfrac{1}{2}\lvert \mathrm{tr}(f)\rvert, \tfrac{1}{2}\lvert \mathrm{tr}(g)\rvert\right\}\leqslant\;\frac{5C(f,g)-1}{3C(f,g)+1},\] then there exist integers \(n,m\geqslant 1\), so that \(\left(\tfrac{n}{2}T_f,\tfrac{m}{2}T_g\right)\in D\).
Let \(a,b>0\) be such that \[\begin{align}
\label{sinh}
&\sinh b =\frac{\sqrt{2}}{\sqrt{\cosh d-1}} &\text{and} & &\sinh a= \frac{\sqrt{2}}{(2\cosh d+1)\sqrt{\cosh d-1}}.
\end{align}\tag{10}\] These are well-defined since \(d>0\). Also, 10 are equivalent to \[\begin{align}
\label{cosh}
&\cosh b =\sqrt{\frac{\cosh d+1}{\cosh d-1}} &\text{and} & &\cosh a= \frac{(2\cosh d-1)}{(2\cosh d+1)}\sqrt{\frac{\cosh d+1}{\cosh d-1}}.
\end{align}\tag{11}\] Observe that \(b>a\), since \(\sinh b > \sinh a\). Let \(S\) be the square with vertices \((a,b), (b,b), (b,a), (a,a)\) and \(C\) the open region in \(\mathbb{R}^2_+\) bounded by \(S\). That is, \(\overline{C}=S\cup C\). Our main task is to show that \(\overline{C}\setminus\{(a,b), (b,b), (b,a)\}\subseteq D\), as shown in Figure 4.
We start by showing that the edges of \(S\), without the vertices \((a,b), (b,b), (b,a)\), are contained in \(D\). We shall make heavy use of the fact
that the line \(y=x\) is an axis of symmetry for \(h\); i.e. \(h(x,y)=h(y,x)\), for all \(x,y>0\).
Using the formula for \(h\) in 8 and equations 10 , 11 we can evaluate \(h\) on the vertices of
\(S\) as follows: \[\begin{align}
h(a,a)&=\cosh d \sinh^2a-\cosh^2a=(\cosh d -1)\sinh^2a-1=\frac{2}{(2\cosh d+1)^2}-1, \tag{12}\\
h(b,b)&= \cosh d \sinh^2b - \cosh^2b= \cosh d \frac{2}{\cosh d-1} - \frac{\cosh d +1}{\cosh d-1}=1,\tag{13}
\end{align}\] and \[\begin{align}
\label{hab}
h(b,a)=h(a,b) &= \cosh d \sinh a \sinh b - \cosh a \cosh b=\frac{2\cosh d - (2\cosh d-1)(\cosh d+1)}{(2\cosh d+1)(\cosh d-1)}=\nonumber \\
&=\frac{-2\cosh^2 d+\cosh d+1}{(2\cosh d+1)(\cosh d-1)}=-1.
\end{align}\tag{14}\] Observe that 12 implies that \[\label{eq:32antiparallel32haa322}
-1<h(a,a)<-\frac{7}{9}.\tag{15}\] Now consider the partial derivative \[\label{increasing}
\frac{\partial h}{\partial x}(x,b) = \cosh d \cosh x \sinh b - \sinh x \cosh b, \quad \text{for} \quad x>0.\tag{16}\] Equations 10 and 11 yield \[\frac{1}{\cosh
d}\;\frac{\cosh b}{\sinh b}=\frac{\sqrt{\cosh d +1}}{\sqrt{2}\;\cosh d}<1<\frac{\cosh x}{\sinh x},\quad \text{for all}\: x>0.\] Applying this inequality to 16 , we obtain that \(h\) is strictly increasing on the horizontal line \(y=b\). This implies that \[-1=h(a,b)<h(x,b)<h(b,b)=1,\quad \text{for any}\: x\in(a,b),\] meaning that
\(h(x,b)\in D\), for all \(x\in(a,b)\). This proves that the horizontal edge of \(S\) joining \((a,b)\) and \((b,b)\) (without these vertices) is contained in \(D\). By the symmetry of \(h\) we also obtain that the vertical edge of \(S\)
joining \((b,a)\) and \((b,b)\) (excluding the vertices again) is contained in \(D\).
We now turn to the vertical edge of \(S\) joining \((a,a)\) and \((a,b)\). Consider the partial derivative \[\frac{\partial
h}{\partial y}(a,y) = \cosh d \sinh a \cosh y - \cosh a \sinh y, \quad \text{for}\quad y>0.\] Since \(\frac{\partial h}{\partial y}(a,a)= (\cosh d -1) \sinh a \cosh a>0\), by continuity there exists \(y_+\in(a,b)\), close enough to \(a\), so that \(\frac{\partial h}{\partial y}(a,y_+)>0\).
Suppose that \(\frac{\partial h}{\partial y}(a,y)\neq0\) for all \(y\in(a,b)\). Then, again by continuity and the fact that \(\frac{\partial h}{\partial
y}(a,y_+)>0\), we have that \(\frac{\partial h}{\partial y}(a,y)>0\), for all \(y\in(a,b)\), which implies that \(h(a,y)\) is increasing in
\((a,b)\). But \(h(a,a)>-1=h(a,b)\), from 14 and 15 , which is a contradiction.
We conclude that \(\frac{\partial h}{\partial y}(a,y_0)=0\), for some \(y_0\in(a,b)\). That is \[\begin{align}
&\frac{\partial h}{\partial y}(a,y_0)=\cosh d \sinh a \sinh y_0 - \cosh a \sinh y_0=0 & \iff& &\coth y_0 = \frac{\coth a}{\cosh d }.
\end{align}\] If \(0<y<y_0\), then \[\begin{align}
&\coth y >\coth y_0 = \frac{\coth a}{\cosh d } & &\iff& &\cosh d \cosh y \sinh a > \cosh a \sinh y & &\iff& &\frac{\partial h}{\partial y}(a,y)>0.
\end{align}\] Similarly we can show that if \(y>y_0\), then \(\frac{\partial h}{\partial y}(a,y)<0\). Therefore the maximum of \(h(a,y)\), for
all \(y>0\), is attained only at \(y_0\). This means that \[\min\{h(a,a), h(a,b)\} < h(a,y) \leqslant h(a,y_0),\quad \text{for all}\;y\in(a,b).\] But,
\(h(a,a)>-1=h(a,b)\), again by 14 and 15 , and so \(h(a,y)>-1\), for all \((a,b)\).
In order to conclude that \(h(a,y)\in D\), for all \(y\in(a,b)\), it suffices to show that \(h(a,y_0)<0\), which we now prove. \[\begin{align}
h(a,y_0)<0 &\iff \cosh d \sinh a \sinh y_0 <\cosh a \cosh y_0 \\
&\iff \cosh d < \coth a \coth y_0= \coth a\;\frac{\coth a}{\cosh d }\\
&\iff \cosh d <\coth a.
\end{align}\] But, equations 10 and 11 yield that \[\coth a = \frac{(2\cosh d-1)\sqrt{\cosh d+1}}{\sqrt{2}}=\frac{(\cosh d + \cosh d -1)\sqrt{\cosh d+1}}{\sqrt{2}}
>\frac{\cosh d\sqrt{2}}{\sqrt{2}}=\cosh d,\] since \(d>0\). This concludes the proof of the fact that \(h(a,y)\in D\), for all \(y\in(a,b)\),
which implies that the vertical edge of \(S\) joining \((a,a)\) and \((a,b)\) (apart from the vertex \((a,b)\)) lies in
\(D\). Using the symmetry of \(h\), again, we obtain that the horizontal edge of \(S\) joining \((a,a)\) and \((b,a)\) (apart from the vertex \((b,a)\)) lies in \(D\).
So, as we claimed, all edges of \(S\) lie in \(D\), apart from the vertices \((a,b),(b,b),(b,a)\), which lie on the boundary of \(D\), since \(h(a,b)=h(b,a)=-1\) and \(h(b,b)=1\). With these facts in mind, we move on to the proof of the inclusion \(C\subseteq
D\).
First, a simple computation shows that the function \(h(x,x)\) is strictly increasing for all \(x>0\). Therefore, recalling 8 and 13 , we get that \[\label{eq:32anitparallel32y61x}
-1=\lim_{x\to0^+}h(x,x)<h(x,x)<h(b,b)=1,\quad \text{for all}\quad 0<x<b.\tag{17}\] This means that \(C\cap\{(x,x)\colon x>0\}\subseteq D\). Now, fix \(x_0>0\)
and consider the derivative \[\label{eq:32antiparallel32final32der}
\frac{\partial h}{\partial y}(x_0,y)=\cosh d \sinh x_0 \cosh y- \cosh x_0 \sinh y.\tag{18}\] Since the hyperbolic tangent is an increasing function, we obtain that for \(y<x_0\)
\[\label{eq:32antiparallel32final32der322}
\tanh y < \tanh x_0 < \cosh d \tanh x_0.\tag{19}\] Combining 18 and 19 yields that \(h(x_0,y)\) is
strictly increasing for all \(y\in(0,x_0)\), and any \(x_0>0\). So, if \(x_0\in(a,b)\) then \[-1<h(x_0,a)<h(x_0,y)<h(x_0,x_0)<1,\] where the inequality \(-1<h(x_0,a)\) follows from the fact that \((x_0,a)\) lies on an edge of \(S\) and inequality \(h(x_0,x_0)<1\) is 17 applied to \(x_0\). We conclude that the vertical line segment joining
\((x_0,a)\) and \((x_0,x_0)\), which is contained in \(C\), lies in \(D\) for all \(x_0\in(a,b)\). Using the symmetry of \(h\) one last time, we obtain that \(C\subseteq D\), as claimed.
To finish the proof of the lemma, note that since the edges of \(S\) have length \(b-a\), if a point \((x,y)\in\mathbb{R}_+^2\) is such that \(\max\{x,y\}<b-a\), then there exist integers \(n,m\geqslant 1\) such that \((nx,my)\in C\subseteq D\). Therefore, recalling equation 9 , we have that if \[\max\left\{\tfrac{1}{2}T_f, \tfrac{1}{2}T_g \right\}< b-a,\] then there exist integers \(n,m\geqslant 1\) such that \(f^n\circ g^m\) is elliptic. Next, assume that \((x,y)\in\mathbb{R}_+^2\) is such that \(\max\{x,y\}=b-a\). If \(x<y\) (or,
similarly, \(y<x\)), then there exist \(n,m\geqslant 1\) so that \((nx,my)\in \overline{C}\setminus\{(a,b),(b,b),(b,a)\}\subseteq D\) and we are done once
again. On the other hand, if \(x=y=b-a\), then since \(b-a<b\) we can use 17 to deduce that \(-1<h(b-a,b-a)<1\), meaning that \((x,y)\in D\).
We have thus shown that if \[\max\left\{\tfrac{1}{2}T_f, \tfrac{1}{2}T_g \right\}\leqslant b-a,\] then there exist integers \(n,m\geqslant 1\) so that \(f^n\circ
g^m\) is elliptic. As such, the proof of this lemma will now be complete upon showing that \[\max\left\{\tfrac{1}{2}T_f, \tfrac{1}{2}T_g \right\}\leqslant b-a \quad \iff \quad \max\left\{
\tfrac{1}{2}\lvert\mathrm{tr}(f)\rvert, \tfrac{1}{2}\lvert \mathrm{tr}(g)\rvert\right\}\leqslant\frac{5\;C(f,g)-1}{3\;C(f,g)+1}.\] Recall that from 4 we have that \(\cosh \left(\tfrac{1}{2}T_f\right) = \tfrac{1}{2} \lvert \mathrm{tr}(f)\rvert\) and so we only need to show that \[\label{eq:32antiparallel32end32goal}
\cosh (b-a) = \frac{5C(f,g)-1}{3C(f,g)+1}.\tag{20}\] But, \(\cosh(b-a) = \cosh b\cosh a - \sinh b\sinh a\), and all quantities on the right-hand side are given by equations 10 and 11 . So, \[\begin{align}
\cosh(b-a) &= \frac{(2\cosh d -1)}{(2\cosh d+1)}\frac{(\cosh d+1)}{(\cosh d-1)} - \frac{2}{(2\cosh d+1)(\cosh d-1)}\nonumber \\ &=\frac{2\cosh^2 d +\cosh d-3}{(2\cosh d+1)(\cosh d-1)}\nonumber \\ &=\frac{2\cosh d+3}{2\cosh d+1}. \label{coshba}
\end{align}\tag{21}\] Now, by Corollary 2 (ii) and the fact that \(C(f,g)>1\), we have that \(\cosh
d = \frac{C(f,g)+1}{C(f,g)-1}\) which when substituted to 21 yields 20 . ◻
Some of the arguments in the proof of Lemma 2 can be used to prove the following result, which will be used in the proof of Theorem 3 in the next section.
Lemma 3. Assume that \(f,g\in\mathrm{PSL}(2,\mathbb{R})\) are hyperbolic matrices such, \(C(f,g)>1\) and \(\lvert \textrm{tr}(f) \rvert =\lvert \textrm{tr}(g) \rvert=t\). If \(\tfrac{1}{2}t\leqslant\sqrt{C(f,g)}\), then the transformation \(f\circ g\) is either elliptic or parabolic.
Proof. Since \(C(f,g)>1\) the axes of \(f,g\) are disjoint and a hyperbolic distance \(d>0\) apart. Also, from 4 and our assumption, we have that \[\tfrac{1}{2}t=\tfrac{1}{2}\lvert \textrm{tr}(f)\rvert =\cosh \tfrac{1}{2} T_f = \tfrac{1}{2}\lvert \textrm{tr}(g) \rvert =\cosh \tfrac{1}{2}
T_g.\] This means that \(T\vcentcolon=T_f=T_g\).
As in the proof of Lemma 2, we have that \[\label{eq:32equal32trace32eq1}
\frac{1}{2}\lvert \textrm{tr}(f\circ g)\rvert= \lvert h\left(\tfrac{1}{2}T_f,\tfrac{1}{2}T_g\right)\rvert= \lvert h\left(\tfrac{1}{2}T,\tfrac{1}{2}T\right)\rvert,\tag{22}\] where \(h(x,y)=\cosh d\sinh x\sinh y-\cosh
x\cosh y\). In 17 we showed that \(-1<h(x,x)<h(b,b)=1\), for all \(x\in(0,b)\), where \(b>0\)
satisfies \[\label{eq:32equal32trace32eq2}
\cosh b =\sqrt{\frac{\cosh d+1}{\cosh d-1}}.\tag{23}\] So, 22 tells us that if \(\tfrac{1}{2}T\leqslant b\) we have that \(f\circ g\)
is either elliptic (when \(T<b\)) or parabolic (when \(T=b\)). Equivalently, by 23 , the inequality \(\tfrac{1}{2}T\leqslant b\) can be written as \[\label{eq:32equal32trace32eq3}
\tfrac{1}{2}t=\cosh\tfrac{1}{2}T\leqslant\cosh b = \sqrt{\frac{\cosh d+1}{\cosh d-1}}.\tag{24}\] But, from Corollary 2 (ii), and the fact that \(C(f,g)>1\), we have that \[\cosh d = \frac{C(f,g)+1}{C(f,g)-1},\] which when substituted to 24 yields the desired result. ◻
Before proceeding to the proof of Theorem 2, we require a some additional results. The first is [@JS], where an elliptic transformation \(h\in\mathrm{PSL}(2,\mathbb{R})\) is called of finite order if \(h^m=\mathrm{Id}\) for some \(m\in\mathbb{N}\) and of infinite order otherwise.
Lemma 4 ([@JS]). Suppose that the semigroup generated by \(\{f,g\}\subseteq\mathrm{PSL}(2,\mathbb{R})\) contains a transformation \(w=w_1w_2\cdots w_n\) that is elliptic of finite order, where \(w_i\in\{f,g\}\) and not all \(w_i\) are the same. Then the semigroup and the group generated by \(\{f,g\}\) coincide.
Next, we require [@BBC1996].
Theorem 5 ([@BBC1996]). If \(f,g\in\mathrm{PSL}(2,\mathbb{R})\) are non-commuting transformations, and \(f\) is elliptic of infinite order, then the semigroup generated by \(\{f,g\}\) is dense in \(\mathrm{PSL}(2,\mathbb{R})\).
Finally, we have a standard result on discrete subgroups of \(\mathrm{PSL}(2,\mathbb{R})\) (see, for example, [@Be1995]).
Theorem 6. A non-elementary, non-discrete subgroup of \(\mathrm{PSL}(2,\mathbb{R})\) contains an elliptic element of infinite order.
We should mention that a subsgroup \(G\) of \(\mathrm{PSL}(2,\mathbb{R})\) is called non-elementary if the orbit \(G(z)=\{g(z)\colon g\in G\}\) is infinite for any \(z\in\mathbb{H}\). Elementary groups are thoroughly understood, since they can be split into three main categories as described in [@Be1995]. For our purposes it suffices to note that \(G\) is non-elementary whenever it contains two hyperbolic transformations with distinct fixed points (i.e. with well-defined and non-zero cross-ratio).
We are now ready to prove Theorem 2, which we restate below in the language of \(\mathrm{PSL}(2,\mathbb{R})\), for the convenience of the reader.
Theorem 1. Assume that \(f,g\in\mathrm{PSL}(2,\mathbb{R})\) are hyperbolic transformations such that \(C(f,g)\) is well-defined and \(C(f,g)>1\). If \[\max\left\{\tfrac{1}{2}\lvert \mathrm{tr}(f)\rvert, \tfrac{1}{2}\lvert \mathrm{tr}(g)\rvert\right\}\leqslant\;\frac{5\;C(f,g)-1}{3\;C(f,g)+1},\] then semigroup generated by \(\{f,g\}\) is either a discrete subgroup of \(\mathrm{PSL}(2,\mathbb{R})\) or is dense in \(\mathrm{PSL}(2,\mathbb{R})\).
Proof. By assumption, Lemma 2 is applicable and yields that \(w=f^n\circ g^m\) is elliptic for some integers \(n,m\geqslant 1\). Assume first that \(w\) has infinite order. Since \(C(f,g)>1\), the transformations \(f\) and \(g\) do not commute, and so neither do \(w\) and \(f\). Thus Theorem 5 is applicable and yields that the semigroup generated by \(\{f,w\}\) is dense in \(\mathrm{PSL}(2,\mathbb{R})\).
If, on the other hand, \(w\) has finite order, Lemma 4 implies that the semigroup generated by \(\{f,g\}\) is in fact a group. If this group is discrete, we are done. Otherwise, again because \(C(f,g)>1\), we have that the group generated by \(\{f,g\}\)
is non-elementary. So, using Theorem 6 we get that the semigroup generated by \(\{f,g\}\) contains an elliptic of
infinite order and we can follow the arguments we presented in the previous paragraph to establish that it is dense in \(\mathrm{PSL}(2,\mathbb{R})\). ◻
In preparation for the proof of Theorem 1, that will take place in the next section, we prove certain conditions on pairs of hyperbolic transformations that guarantee domination. This section also contains the proof of Theorem 3.
Let us first state some additional properties of hyperbolic transformations. Any hyperbolic transformation \(f\) can be written as a composition \(f=\sigma_1\circ \sigma_2\), where each \(\sigma_i\) is the reflection in a hyperbolic geodesic \(\gamma_i\) that is perpendicular to the axis \(\mathrm{Ax}(f)\), for \(i=1,2\). The geodesics \(\gamma_1\) and \(\gamma_2\) are not unique, but the distance between any such geodesics is fixed and equal to half the translation length; that is \[d_\mathbb{H}(\gamma_1,\gamma_2)=\tfrac{1}{2}T_f.\] We will say that an interval \(I\subseteq\overline{\mathbb{R}}\) is symmetric with respect to \(f\) if the unique hyperbolic geodesic with the same endpoints as \(I\) is perpendicular to the axis \(\mathrm{Ax}(f)\).
The following observation will prove important in what follows:
Remark 7. Let \(f\) be a hyperbolic transformation and \(C>0\) a constant so that \(T_f> C\). Then, we can find intervals \(U_f,S_f\subseteq\overline{\mathbb{R}}\), depending on \(C\), with the following properties:
\(U_f\) and \(S_f\) are symmetric with respect to \(f\),
\(f\left(S_f^c\right)\subseteq U_f\), where \(S_f^c=\overline{\mathbb{R}}\setminus S_f\),
if \(\gamma_U\) is the hyperbolic geodesic of \(\mathbb{H}\) with the same endpoints as \(U_f\), and \(\gamma_S\) is the hyperbolic geodesic of \(\mathbb{H}\) with the same endpoints as \(S_f\), then \(d_\mathbb{H}(\gamma_U,\gamma_S)=C\).
Lemma 5. Let \(f,g\) be two hyperbolic transformations with \(C(f,g)>0\). If \[\min\left\{\tfrac{1}{2}\lvert \mathrm{tr}(f)\rvert, \tfrac{1}{2}\lvert \mathrm{tr}(g)\rvert\right\} > \max\left\{\sqrt{C(f,g)},\frac{1}{\sqrt{C(f,g)}}\right\},\] then there exist open intervals \(U_f, S_f,U_g,S_g\) in \(\overline{\mathbb{R}}\), with pairwise disjoint closures, that satisfy the following properties: the intervals \(U_f,S_f\) are symmetric with respect to \(f\) and \(f(S_f^c)\) is contained in \(U_f\); the intervals \(U_g,S_g\) are symmetric with respect to \(g\) and \(g(S_g^c)\) is contained in \(U_g\). In particular, \(\{f,g\}\) is dominated.
Proof. Let \(f,g\) be hyperbolic transformations as in the statement of the theorem. Observe that if we prove the existence of the intervals \(U_f,S_f,U_g,S_g\), then Theorem 4 immediately tells us that \(\{f,g\}\) is dominated. Since \(C(f,g)>0\) the axes of \(f\)
and \(g\) are disjoint and a hyperbolic distance \(d>0\) apart. We first consider the case where \(C(f,g)>1\). By assumption we have that
\[\label{eq:32newlem32proof32eq1}
\min\left\{\tfrac{1}{2}\lvert \mathrm{tr}(f) \rvert, \tfrac{1}{2}\lvert \mathrm{tr}(g) \rvert \right\}> \sqrt{C(f,g)}=\max\left\{\sqrt{C(f,g)},\frac{1}{\sqrt{C(f,g)}}\right\}.\tag{25}\] Equation 4 tells us that \(\tfrac{1}{2}\lvert \textrm{tr}(f)\rvert =\cosh \tfrac{1}{2}T_f\), and similarly for \(g\). Thus, 25 is equivalent to \[\min\left\{\cosh\tfrac{1}{2}T_f, \cosh\tfrac{1}{2}T_g \right\}> \sqrt{C(f,g)},\] which in turn can be rewritten as
\[\label{eq:32newlem32proof32eq2}
\min\left\{\sinh\tfrac{1}{2}T_f, \sinh\tfrac{1}{2}T_g \right\}> \sqrt{C(f,g)-1}.\tag{26}\] Recall that Lemma 1 tell us that \(C(f,g)=\coth^2\tfrac{1}{2}d\), meaning that \(\frac{1}{\sinh^2\tfrac{1}{2}d}=C(f,g)-1\). We are thus led to \[\label{eq:32newlem32proof32eq3}
\min\left\{\sinh\tfrac{1}{2}T_f, \sinh\tfrac{1}{2}T_g \right\}> \frac{1}{\sinh \tfrac{1}{2}d}.\tag{27}\] Now, let \(\ell\) be the unique hyperbolic geodesic that is perpendicular to \(\mathrm{Ax}(f)\) and \(\mathrm{Ax}(g)\), and let \(\sigma\) the reflection in \(\ell\). Also, define the reflections \(\sigma_f=f\circ \sigma\) and \(\sigma_g=\sigma\circ g\) and let \(\ell_f, \ell_g\) be their geodesics of reflection respectively. Observe that \(\ell_f\) is also perpendicular to \(\mathrm{Ax}(f)\) and \(\ell_g\) is perpendicular to \(\mathrm{Ax}(g)\) (see Figure 5).
To fully justify the configuration shown in Figure 5, we first claim that the geodesic \(\ell_f\) does not intersect \(\mathrm{Ax}(g)\) in \(\overline{\mathbb{H}}\) and \(\ell_g\) does not intersect \(\mathrm{Ax}(f)\). Assuming that the geodesic \(\ell_f\) meets \(\mathrm{Ax}(g)\) in \(\overline{\mathbb{H}}\) at an angle \(\phi\in [0,\pi/2)\) we can see that the geodesics \(\ell_f,\mathrm{Ax}(f),\ell\) and \(\mathrm{Ax}(g)\) form a quadrilateral in \(\mathbb{H}\) that has three right angles and the fourth is \(\phi\). Moreover, \(\phi\) cannot be \(\pi/2\) because then \(\ell_f\) would coincide with \(\ell\). Applying [@Be1995] to this quadrilateral implies that \[\sinh\tfrac{1}{2}T_g =\frac{\cos\phi}{\sinh d }\leqslant\frac{1}{\sinh\tfrac{1}{2}
d},\] which directly contradicts 27 . We reach the same contradiction if we assume that \(\ell_g\) intersects \(\mathrm{Ax}(f)\) in \(\overline{\mathbb{H}}\).
Next, we show that the geodesics \(\ell_f,\ell_g\) do not intersect in \(\overline{\mathbb{H}}\). Working towards a contradiction we suppose that \(\ell_f\)
and \(\ell_g\) meet in \(\overline{\mathbb{H}}\) at an angle \(\theta \in[0,\pi/2]\). Then, because the pairs \(\ell_f,\mathrm{Ax}(g)\) and \(\ell_g,\mathrm{Ax}(f)\) are disjoint, we get that the geodesics \(\ell,\ell_f,\ell_g,\mathrm{Ax}(f), \mathrm{Ax}(g)\) define a
pentagon in \(\overline{\mathbb{H}}\) with four right angles and the fifth being \(\theta\). Using [@Be1995] we get
\[\label{eq:32newlem32proof32eq4}
\sinh\tfrac{1}{2}T_f\sinh\tfrac{1}{2}T_g\cosh d - \cosh\tfrac{1}{2}T_f\cosh\tfrac{1}{2}T_g = \cos\theta\leqslant 1.\tag{28}\] Consider the function \(h\colon \mathbb{R}^2\to \mathbb{R}\), with \(h(x,y)=\sinh x\sinh y\cosh d-\cosh x\cosh y\), and let \(b\in\mathbb{R}\) be such that \[\sinh b=\frac{1}{\sinh \tfrac{1}{2}d}=\sqrt{\frac{2}{\cosh d-1}}.\] The
function \(h\) and the number \(b\) also appeared in the proof of Lemma 2, where we showed that \(h(x,b)\) is increasing for all \(x>0\) (see the arguments succeeding 16 ), and \(h(b,b)=1\) (see 13 ). Also, \(h(x,x)=\cosh d \sinh^2x -\cosh^2x\) is an increasing function of \(x\). Combining these facts we obtain that \(h(x,b),h(x,x)>1\), for all \(x>b\). Let us fix \(x_0>b\). Now, again in the proof of Lemma 2 we showed that \(h(x_0,y)\) is increasing for all \(y\in(0,x_0)\) (see the arguments succeeding 18 ). We thus have
that \[h(x_0,y)>h(x_0,b)>1,\quad\text{for all}\;y\in (b,x_0).\] Since \(h(x,y)=h(y,x)\), we can conclude that all points \((x,y)\in\mathbb{R}^2\) with
\(\min\{x,y\}>b\) satisfy \(h(x,y)>1\). Now note that by 27 we have that \(\tfrac{1}{2}T_f,\tfrac{1}{2}T_g>b\), meaning that \(h\left(\tfrac{1}{2}T_f,\tfrac{1}{2}T_f\right)>1\), contradicting 28 .
To sum up, we just showed that \(\ell_f\) is disjoint with \(\mathrm{Ax}(g)\) and \(\ell_g\), and similarly \(\ell_g\) is
disjoint with \(\mathrm{Ax}(f)\) and \(\ell_f\). Hence, if we denote by \(U_f\) the open interval in \(\overline{\mathbb{R}}\) that has the same endpoints as \(\ell_f\) and contains \(u(f)\), and by \(I_g\) the open interval that
has the same endpoints as \(\ell_g\) and contains \(s(g)\), we obtain that \(\overline{U_f}\cap \overline{I_g}=\emptyset\). Also, let \(I_f=\sigma(U_f)\) and \(U_g=\sigma(I_g)\). Since \(f=\sigma_f\circ\sigma\) and \(g = \sigma\circ \sigma_g\), we have that \(f(I_f^c)= \overline{U_f}\) and \(g(I_g^c)= \overline{U_g}\). It is now easy to see that we can choose slightly larger intervals \(I_f\subseteq S_f\) and \(I_g\subseteq S_g\) so that \(U_f,S_f,U_g,S_g\) all have disjoint closures, \(f(S_f^c)\subseteq U_f\) and \(g(S_g^c)\subseteq
U_g\), as required. This concludes our proof for the case \(C(f,g)>1\).
Recall that because \(f,g\) are hyperbolic, we have that \(C(f,g)\neq1\). So the next case to consider is when \(0<C(f,g)<1\). But, we have that \(C(f^{-1},g)=1/C(f,g)>1\) and so we can apply the first case to the pair \(\{f^{-1},g\}\), in order to obtain the desired intervals. ◻
Using Lemma 5 and Lemma 3 we can now prove Theorem 3.
Proof of Theorem 3. We assume that \(f,g\) are hyperbolic transformations with \(C(f,g)>1\) and \(\lvert \textrm{tr}(f)\rvert=\lvert \textrm{tr}(g)\rvert=t\). Our goal is to show that \(\{f,g\}\) is dominated if and only if \(\tfrac{1}{2}t>\sqrt{C(f,g)}\).
By Lemma 3 we have that if \(\tfrac{1}{2} t\leqslant\sqrt{C(f,g)}\) then \(f\circ
g\) is either elliptic or parabolic. Since the semigroup generated by a dominated set only contains hyperbolic transformations, \(\{f,g\}\) is not dominated in this case. On the other hand, if \(\tfrac{1}{2} t> \sqrt{C(f,g)}\), Lemma 5 immediately yields that \(\{f,g\}\) is dominated, since \[\max\left\{\sqrt{C(f,g)},\frac{1}{\sqrt{C(f,g)}}\right\} = \sqrt{C(f,g)}.\qedhere\] ◻
Next, we prove the analogue of Lemma 5 for two hyperbolic transformations with crossing axes.
Lemma 6. Suppose that \(f\) and \(g\) are hyperbolic transformations such that \(C(f,g)<0\). If \[\min\left\{\tfrac{1}{2}\lvert\textrm{tr}(f)\rvert, \tfrac{1}{2}\lvert\textrm{tr}(g)\rvert\right\}>\max\left\{\sqrt{1-C(f,g)},\sqrt{1-\frac{1}{C(f,g)}}\right\},\] then there exist open intervals \(U_f, S_f,U_g,S_g\) in \(\overline{\mathbb{R}}\), with pairwise disjoint closures, that satisfy the following properties: the intervals \(U_f,S_f\) are symmetric with respect to \(f\), and \(f(S_f^c)\) is contained in \(U_f\); the intervals \(U_g,S_g\) are symmetric with respect to \(g\), and \(g(S_g^c)\) is contained in \(U_g\).
Proof. Since \(C(f,g)<0\), the axes of \(f\) and \(g\) meet at an angle \(\theta\in(0, \pi)\). For
convenience, we conjugate \(f\) and \(g\) with a transformation in \(\mathrm{PSL}(2,\mathbb{C})\), so that they act on the unit disc \(\mathbb{D}\), their axes cross at the origin and the Euclidean diameter joining \(i\) and \(-i\) bisects \(\theta\). By the
symmetry of this configuration, we also have that the Euclidean diameter joining \(1\) and \(-1\) bisects \(\pi-\theta\). Throughout the proof, we use the
notation \([z,w]\) to denote the closed arc of the unit circle running counter-clockwise from \(z\in\partial\mathbb{D}\) to \(w\in\partial\mathbb{D}\). Similarly, \((z,w)\) will denote the respective open arc. By applying a Euclidean rotation about the origin, we can also assume that both \(f\) and \(g\) map the arc \([-1,1]\) inside itself, and without loss of generality assume that the arc \([u(f),u(g)]\) does not
contain \(s(f)\) and \(s(g)\). See Figure 6 for the geometry of the configuration of \(f\) and \(g\).
We first consider the case where \(\theta\in(0, \tfrac{\pi}{2}]\). Due to Corollary 2 this is equivalent to \(C(f,g)\in[-1,0)\), meaning that in this case \[\label{eq:32intersecting32eq0}
\max\left\{\sqrt{1-C(f,g)},\sqrt{1-\frac{1}{C(f,g)}}\right\}=\sqrt{1-\frac{1}{C(f,g)}}.\tag{29}\] We shall work independently for \(f\) and \(g\). Starting with \(f\), let \(\ell_f\) be the hyperbolic geodesic of \(\mathbb{D}\) landing at \(-i\) that is perpendicular to \(\mathrm{Ax}(f)\), the axis of \(f\). Also, write \(z_f\in\partial\mathbb{D}\) for the other endpoint of \(\ell_f\). Since \(\theta\leqslant\tfrac{\pi}{2}\) and the diameter joining \(-i\) and \(i\) bisects \(\theta\), we have that \(z_f\) lies in \([-1,-i]\). Let \(a_f\in\mathbb{D}\) be the point of intersection of \(\ell_f\) and \(\mathrm{Ax}(f)\). By the symmetry of our configuration, we have that the hyperbolic geodesic \(\widetilde{\ell_f}\) with endpoints \(i\) and \(-z_f\) is also perpendicular to \(\mathrm{Ax}(f)\) and \(-z_f\) lies in \([1,i]\). Also, the point of intersection of \(\widetilde{\ell_f}\) and \(\mathrm{Ax}(f)\) is \(-a_f\). By Remark 7 we have that
if \(T_f>d_\mathbb{D}(a_f,-a_f)\), then there exist open arcs \(U_f\subseteq(z_f,-i)\) and \(S_f\subseteq(-z_f,i)\), that are symmetric with respect to
\(f\), so that \(f(S_f^c)\subseteq U_f\).
Our goal now is to show that the condition \(T_f>d_\mathbb{D}(a_f,-a_f)\) is equivalent to \(\tfrac{1}{2}\lvert\textrm{tr}(f)\rvert>\sqrt{1-\frac{1}{C(f,g)}}\). Observe that we have
\(d_\mathbb{D}(a_f,-a_f)=d_\mathbb{D}(a_f,0)+d_\mathbb{D}(0,-a_f)\) and \(d_\mathbb{D}(a_f,0)=d_\mathbb{D}(0,-a_f)\). So the condition \(T_f>d_\mathbb{D}(a_f,-a_f)\) is equivalent to \(\tfrac{1}{2}T_f>d_\mathbb{D}(a_f,0)\). Consider the hyperbolic triangle with vertices \(a_f,0\) and \(-i\). This triangle has angles \(0,\tfrac{\pi}{2}\) and \(\tfrac{\theta}{2}\), meaning that we can use the Angle of Parallelism [@Be1995] in order to obtain that \[\label{eq:32intersecting32eq1}
\cosh d_\mathbb{D}(a_f,0)\;\sin\tfrac{\theta}{2}=1.\tag{30}\] Equation 30 can be rewritten as \[\label{eq:32intersecting32eq2}
\cosh d_\mathbb{D}(a_f,0) = \frac{1}{\sqrt{1-\cos^2\tfrac{\theta}{2}}}=\sqrt{1-\frac{1}{C(f,g)}},\tag{31}\] where for the last equality we used Corollary 2 (i).
Hence, using 31 and the equation \(\cosh\tfrac{1}{2}T_f=\tfrac{1}{2}\lvert\textrm{tr}(f)\rvert\), i.e. 4 , we have that \[\begin{align}
\tfrac{1}{2}T_f>d_\mathbb{D}(a_f,0)& & \iff & &\tfrac{1}{2}\lvert\textrm{tr}(f)\rvert=\cosh\tfrac{1}{2}T_f> \cosh d_\mathbb{D}(a_f,0) = \sqrt{1-\frac{1}{C(f,g)}}.
\end{align}\] Following the same arguments for \(g\) in place of \(f\), we can see that if \(\tfrac{1}{2}\lvert\textrm{tr}(g)\rvert>\sqrt{1-\frac{1}{C(f,g)}}\), then there exists a point \(z_g\in[-i,1]\) and arcs \(U_g\subseteq(-i,z_g)\) and \(S_g\subseteq(i,-z_g)\), symmetric with respect to \(g\), and such that \(g(S_g^c)\subseteq U_g\). By construction, we have that the arcs \(U_f,S_f,U_g,S_g\) have pairwise disjoint closures, which concludes the proof in the case where \(\theta\in(0,\tfrac{\pi}{2}]\).
For the case \(\theta\in(\tfrac{\pi}{2},\pi)\), we can apply the above arguments to the pair \(f^{-1},g\), since their axes cross at angle \(\pi-\theta\in(0,\tfrac{\pi}{2})\). Recalling that \(C(f^{-1},g)=1/C(f,g)\) yields the desired conclusion. ◻
The next theorem combines the cases presented in Lemmas 5 and 6 into a single statement.
Theorem 8. Suppose that \(f,g\) are hyperbolic transformations with \(C(f,g)\neq0\). If \[\min\left\{\tfrac{1}{2}\lvert\textrm{tr}(f)\rvert, \tfrac{1}{2}\lvert\textrm{tr}(g)\rvert\right\}>\max\left\{\sqrt{1+\left\lvert C(f,g)\right\rvert},\sqrt{1+\frac{1}{\left\lvert C(f,g)\right\rvert}}\right\},\] then there exist open intervals \(U_f, S_f,U_g,S_g\) in \(\overline{\mathbb{R}}\), with pairwise disjoint closures, that satisfy the following properties: the intervals \(U_f,S_f\) are symmetric with respect to \(f\), and \(f(S_f^c)\) is contained in \(U_f\); the intervals \(U_g,S_g\) are symmetric with respect to \(g\), and \(g(S_g^c)\) is contained in \(U_g\).
Proof. If \(C(f,g)<0\), then the the statement of the theorem is identical to the statement of Lemma 6. If \(C(f,g)>0\), then \[\begin{align} \min\left\{\tfrac{1}{2}\lvert\textrm{tr}(f)\rvert, \tfrac{1}{2}\lvert\textrm{tr}(g)\rvert\right\}&>\max\left\{\sqrt{1+\left\lvert C(f,g)\right\rvert},\sqrt{1+\frac{1}{\left\lvert C(f,g)\right\rvert}}\right\}\\ &= \max\left\{\sqrt{1+ C(f,g)},\sqrt{1+\frac{1}{ C(f,g)}}\right\}>\max\left\{\sqrt{C(f,g)},\sqrt{\frac{1}{C(f,g)}}\right\}, \end{align}\] and so the conclusion is obtained by applying Lemma 5. ◻
We end this section with a simple remark combining 7 and Theorem 8.
Remark 9. Let \(f\) and \(g\) by hyperbolic transformations with \(C(f,g)\neq 0\), satisfying \[\min\left\{\tfrac{1}{2}\lvert\textrm{tr}(f)\rvert, \tfrac{1}{2}\lvert\textrm{tr}(g)\rvert\right\}>\max\left\{\sqrt{1+\left\lvert C(f,g)\right\rvert},\sqrt{1+\frac{1}{\left\lvert C(f,g)\right\rvert}}\right\},\] as in the
assumption of Theorem 8. For any \(\varepsilon>0\) with \[\tfrac{1}{2}\lvert\textrm{tr}(f)\rvert> \max\left\{\sqrt{1+\left\lvert C(f,g)\right\rvert},\sqrt{1+\frac{1}{\left\lvert C(f,g)\right\rvert}}\right\}\cdot\varepsilon,\] the intervals \(U_f,S_f\) that appear in the conclusion of Theorem 8 can be chosen to have the following property:
If \(\ell_U,\gamma_S\) are the hyperbolic geodesics of \(\mathbb{H}\) with the same endpoints as \(U_f\) and \(S_f\),
respectively, then \(d_\mathbb{H}(\ell_U,\gamma_S)\), the hyperbolic distance between \(\ell_U\) and \(\gamma_S\), satisfies \[\max\left\{\sqrt{1+\left\lvert C(f,g)\right\rvert},\sqrt{1+\frac{1}{\left\lvert C(f,g)\right\rvert}}\right\}\cdot \varepsilon=\cosh \tfrac{1}{2}d_\mathbb{H}(\ell_U,\gamma_S).\] An analogous property will hold for the intervals
\(U_g,S_g\).
In this section we prove Theorem 1, which is restated below in the setting of \(\mathrm{PSL}(2,\mathbb{R})\), for convenience.
Theorem 2. Let \(u_1,u_2,\dots,u_n,s_1,s_2,\dots s_N\in\mathbb{R}\mathbb{P}^1\) be \(2N\) pairwise distinct points in the projective real line. Suppose that \(f_1,f_2,\dots f_N\in\mathrm{PSL}(2,\mathbb{R})\) are hyperbolic transformations with \(u(f_i)=u_i\) and \(s(f_i)=s_i\), for \(i=1,2,\dots,N\), and consider the constants \[M_{i,j} = \max\left\{1+\left\lvert C(f_i,f_j)\right\rvert,1+\frac{1}{\left\lvert C(f_i,f_j)\right\rvert}\right\},\] for all pairs \(i,j\in\{1,2,\dots,N\}\) with \(i\neq j\). If \[\label{eq:32main32assumption} \tfrac{1}{2}\lvert \textrm{tr}(f_k)\rvert >4\;\max_{i\neq j}\{M_{i,j}\}\cdot \max_{i\neq j}\left\{\frac{\lvert C(f_i,f_j)\rvert +1}{\lvert C(f_i,f_j)-1\rvert}\right\}, \quad \text{for all}\;k=1,2,\dots,N,\qquad{(1)}\] then the set \(\{f_1,f_2,\dots, f_N\}\) is dominated.
Proof. Suppose that \(f_1,f_2,\dots,f_N\) are hyperbolic transformations, with the properties described in the statement of the theorem. Note that, by assumption, the fixed points of all \(f_i\) are distinct, meaning that \(C(f_i,f_j)\) is well-defined and \(C(f_i,f_j)\neq0\), for all \(i\neq j\). Due to the
geometric characterisation of domination in Theorem 4, it suffices to show that under the trace condition in ?? we can find open intervals \(U^1,U^2,\dots,
U^k\), with disjoint closures, so that each \(f_k\) maps \(\overline{\cup_{k=1}^NU^k}\) into \(U^k\).
Let \(\varepsilon>1\) be such that for all \(i=1,2,\dots,N\) we have that \[\label{eq:32proof32epsilon}
\tfrac{1}{2}\lvert\textrm{tr}(f_i)\rvert > 4\varepsilon\;\max_{i\neq j}\{M_{i,j}\}\cdot \max_{i\neq j}\left\{\frac{\lvert C(f_i,f_j)\rvert +1}{\lvert C(f_i,f_j)-1\rvert}\right\}.\tag{32}\] Fix \(k=1,2,\dots,N\). For any \(i\neq k\), the hypothesis ?? and 32 yield \[\begin{align}
\min\{\tfrac{1}{2}\lvert \textrm{tr}(f_k)\rvert,\tfrac{1}{2}\lvert \textrm{tr}(f_i)\rvert\}&> 4\varepsilon\cdot M_{k,i}\cdot \frac{\lvert C(f_k,f_i)\rvert +1}{\lvert C(f_k,f_i)-1\rvert} \nonumber\\ &=4\varepsilon\cdot \max\left\{1+\left\lvert
C(f_k,f_i)\right\rvert,1+\frac{1}{\left\lvert C(f_k,f_i)\right\rvert}\right\}\cdot \frac{\lvert C(f_k,f_i)\rvert +1}{\lvert C(f_k,f_i)-1\rvert} \nonumber\\ &>\sqrt{\varepsilon} \cdot \max\left\{\sqrt{1+\left\lvert
C(f_k,f_i)\right\rvert},\sqrt{1+\frac{1}{\left\lvert C(f_k,f_i)\right\rvert}}\right\} \label{eq:32proof32estimate322} \\ &> \max\left\{\sqrt{1+\left\lvert C(f_k,f_i)\right\rvert},\sqrt{1+\frac{1}{\left\lvert C(f_k,f_i)\right\rvert}}\right\}\nonumber
,
\end{align}\tag{33}\] which means that Theorem 8 is applicable to the pair of hyperbolic transformations \(f_k,f_i\). So we can find intervals \(U_i^k,U_k^i,S_i^k,S_k^i\subseteq\overline{\mathbb{R}}\) with pairwise disjoint closures so that \(U_i^k,S_i^k\) are
symmetric with respect to \(\mathrm{Ax}(f_k)\), the axis of \(f_k\), and \(U_k^i,S_k^i\) are symmetric with respect to \(\mathrm{Ax}(f_i)\) . Also the intervals satisfy \(f_k\left(\overline{\mathbb{R}}\setminus S_i^k\right)\subseteq U_i^k\) and \(f_i\left(\overline{\mathbb{R}}\setminus
S_k^i\right)\subseteq U_k^i\).
Moreover, in 33 we got that \[\min\{\tfrac{1}{2}\lvert \textrm{tr}(f_k)\rvert,\tfrac{1}{2}\lvert \textrm{tr}(f_i)\rvert\}> \sqrt{\varepsilon}\cdot \max\left\{\sqrt{1+\left\lvert
C(f_k,f_i)\right\rvert},\sqrt{1+\frac{1}{\left\lvert C(f_k,f_i)\right\rvert}}\right\}.\] Hence, according to Remark 9, we can choose
the intervals \(U_i^k,U_k^i,S_i^k,S_k^i\) to have the following properties:
Let \(\ell_i^k,\ell_k^i\) be the hyperbolic geodesics of \(\mathbb{H}\) with the same endpoints as \(U_i^k\) and \(U_k^i\),
respectively. Similarly, \(\gamma_i^k,\gamma_k^i\) are the hyperbolic geodesics of \(\mathbb{H}\) with the same endpoints as \(S_i^k\) and \(S_k^i\), respectively. Then the geodesics \(\ell_i^k,\ell_k^i,\gamma_i^k,\gamma_k^i\) are pairwise distinct and satisfy \[\label{eq:32proof32distance32of32intervals}
\cosh\tfrac{1}{2}d_\mathbb{H}(\ell_i^k,\gamma_i^k)= \cosh\tfrac{1}{2}d_\mathbb{H}(\ell_k^i,\gamma_k^i)=\max\left\{\sqrt{1+\left\lvert C(f_k,f_i)\right\rvert},\sqrt{1+\frac{1}{\left\lvert C(f_k,f_i)\right\rvert}}\right\}\cdot
\sqrt{\varepsilon}.\tag{34}\] By repeating this process for all \(i=1,2,\dots,N\), with \(i\neq k\) we obtain two collections of \(N-1\)
intervals \(\mathbb{U}^k=\{U_1^k,\dots, U^k_{k-1},U^k_{k+1},\dots, U_N^k\}\) and \(\mathbb{S}^k=\{S_1^k,\dots, S^k_{k-1},S^k_{k+1},\dots, S_N^k\}\). Since \(U^k_i,S^k_i\) are symmetric with respect to the axis of \(f_k\), the collections \(\mathbb{U}^k\) and \(\mathbb{S}^k\) are
totally ordered with the set inclusion operation; i.e. for any \(U^k_i,U^k_j\in\mathbb{U}^k\) we have that either \(U^k_i\subseteq U^k_j\) or \(U^k_j\subseteq
U^k_i\), and similarly for the intervals in \(\mathbb{S}^k\). Let \(U^k_n\) be a least element of \(\mathbb{U}^k\) (an inner-most interval in the
collection) and \(S^k_m\) a least element of \(\mathbb{S}^k\). Here, we say a least element, and not the, since they are not necessarily unique. Furthermore, the indices
\(n\) and \(m\) might not coincide. Let us also consider the intervals \(U^k_m\) and \(S^k_n\). Because \(U^k_n\) and \(S^k_m\) are least elements, we have that \(U^k_n\subseteq U^k_m\) and \(S^k_m\subseteq S^k_n\). Now, observe that
\[\overline{U^k_n}\cap \overline{S^k_m}\subseteq\overline{U^k_n}\cap \overline{S^k_n}=\emptyset,\] because, by construction, \(U^k_n\) and \(S^k_n\) have
disjoint closures.
Now let \(k_0=1,2,\dots,N\) with \(k\neq k_0\). Following the process described above for \(k_0\) in place of \(k\), we
obtain collections \(\mathbb{U}^{k_0}, \mathbb{S}^{k_0}\) of intervals that are symmetric with respect to \(\mathrm{Ax}(f_{k_0})\). Let \(U^{k_0}_{n_0}\) and
\(S^{k_0}_{m_0}\) be some least elements in \(\mathbb{U}^{k_0}\) and \(\mathbb{S}^{k_0}\), respectively. We want to also show that the intervals \(U^k_n,S^k_m,U^{k_0}_{n_0},S^{k_0}_{m_0}\) have pairwise disjoint closures, where \(U^k_n,S^k_m\) are the intervals that were constructed above. First, observe that as we showed before, we have
that \(\overline{U^{k_0}_{n_0}}\cap\overline{S^{k_0}_{m_0}}=\emptyset\). Since \(U^k_n,U^{k_0}_{n_0}\) are least elements in their respective collections, we have that \[\overline{U^{k_0}_{n_0}}\cap\overline{U^k_n}\subseteq\overline{U^{k_0}_k}\cap\overline{U^k_{k_0}}=\emptyset,\] because by construction we have that \(U^k_{k_0},S^k_{k_0},U^{k_0}_k,S^{k_0}_k\)
have pairwise disjoint closures. Similar arguments yield the desired conclusion.
The process we just described allows us to construct intervals \(U^1,U^2,\dots,U^N, S^1,S^2,\dots,S^N\) (we suppress the indices and keep the exponents, for visual clarity) that have pairwise disjoint closures, and so that
each pair \(U^i,S^i\) is symmetric with respect to the axis \(\mathrm{Ax}(f_i)\) of \(f_i\), for \(i=1,2,\dots,N\).
Therefore, our proof will be complete upon showing that \(f_i\left(\overline{\mathbb{R}}\setminus S^i\right)\subseteq U^i\), for all \(i=1,2,\dots,N\).
Focusing on the exponent \(k\) we fixed earlier, and reintroducing the indices in our intervals, that is considering the inner-most intervals \(U^k_n\) and \(S^k_m\) described above, we have to show that \[\label{eq:32end32goal}
f_i\left(\overline{\mathbb{R}}\setminus S^k_m\right)\subseteq U^k_n.\tag{35}\] First, observe that if \(n=m\), then we already have the desired conclusion, by construction. So, for the rest of the proof, we
assume that \(n\neq m\). We will also use the intervals \(U^k_m\) and \(S^k_ n\). Let \(\ell^k_n,\ell^k_m\) be the geodesics
of \(\mathbb{H}\) with the same endpoints as \(U^k_n,U^k_m\), respectively. Similarly, \(\gamma^k_n,\gamma^k_m\) be the geodesics of \(\mathbb{H}\) with the same endpoints as \(S^k_n,S^k_m\), respectively. Because the intervals in our construction satisfied 34 , we have that
\[\label{eq:32distance32of32intervals32n}
\cosh\tfrac{1}{2}d_\mathbb{H}(\ell_n^k,\gamma_n^k)=\max\left\{\sqrt{1+\left\lvert C(f_k,f_n)\right\rvert},\sqrt{1+\frac{1}{\left\lvert C(f_k,f_n)\right\rvert}}\right\}\cdot \sqrt{\varepsilon},\tag{36}\] and
\[\label{eq:32distance32of32intervals32m}
\cosh\tfrac{1}{2}d_\mathbb{H}(\ell_m^k,\gamma_m^k)=\max\left\{\sqrt{1+\left\lvert C(f_k,f_m)\right\rvert},\sqrt{1+\frac{1}{\left\lvert C(f_k,f_m)\right\rvert}}\right\}\cdot \sqrt{\varepsilon}.\tag{37}\] Observe that if \(T_{f_k}\), the translation length of \(f_k\), satisfies \(T_{f_k}>d_\mathbb{H}(\ell_n^k,\gamma_m^k)\), then we have the desired inclusion 35 . Equivalently, it suffices to show that \[\label{eq:32end32goal322}
\tfrac{1}{2}\lvert \textrm{tr}(f_k)\rvert =\cosh\tfrac{1}{2}T_{f_k}>\cosh\tfrac{1}{2}d_\mathbb{H}(\ell_n^k,\gamma_m^k).\tag{38}\] For simplicity, let us write \(d_0=d_\mathbb{H}(\ell_n^k,\gamma_m^k)\),
\(d_n=d_\mathbb{H}(\ell_n^k,\gamma_n^k)\) and \(d_m=d_\mathbb{H}(\ell_m^k,\gamma_m^k)\). To simplify future estimates we observe that due to 36 and 37 we have that \[\begin{align}
\cosh \tfrac{1}{2}d_n\;\cosh \tfrac{1}{2}d_m &= \max\left\{\sqrt{1+\left\lvert C(f_k,f_n)\right\rvert},\sqrt{1+\frac{1}{\left\lvert C(f_k,f_n)\right\rvert}}\right\}\cdot \max\left\{\sqrt{1+\left\lvert
C(f_k,f_m)\right\rvert},\sqrt{1+\frac{1}{\left\lvert C(f_k,f_m)\right\rvert}}\right\} \cdot \varepsilon\nonumber\\ &\leqslant\left(\max_{i\neq j}\left\{ \max\left\{\sqrt{1+\left\lvert C(f_i,f_j)\right\rvert},\sqrt{1+\frac{1}{\left\lvert
C(f_i,f_j)\right\rvert}}\right\} \right\}\right)^2\cdot \varepsilon\nonumber\\ &=\max_{i\neq j} \{M_{i,j}\}\cdot \varepsilon. \label{eq:32proof32estimate}
\end{align}\tag{39}\] We consider two cases, in which we will use the trivial inequality \(\cosh(x+y)\leqslant 2\cosh x \;\cosh y\), for all \(x,y\leqslant 0\).
Case 1: \(d_0\leqslant d_n + d_m\) (see left-hand side of Figure 7)
Using 32 and 39 we get that \[\begin{align}
\cosh \tfrac{1}{2}d_0 &\leqslant\cosh \left( \tfrac{1}{2}d_n + \tfrac{1}{2}d_m\right)\leqslant 2\;\cosh \tfrac{1}{2}d_n \;\cosh\tfrac{1}{2} d_m\\ &\leqslant 2 \;\max_{i\neq j} \{M_{i,j}\} \cdot \varepsilon\leqslant 2\;\max_{i\neq j}\{M_{i,j}\}\cdot
\max_{i\neq j}\left\{\frac{\lvert C(f_i,f_j)\rvert +1}{\lvert C(f_i,f_j)-1\rvert}\right\} \cdot \varepsilon\\ &< 4\;\max_{i\neq j}\{M_{i,j}\}\cdot \max_{i\neq j}\left\{\frac{\lvert C(f_i,f_j)\rvert +1}{\lvert
C(f_i,f_j)-1\rvert}\right\}\cdot\varepsilon< \tfrac{1}{2}\lvert \textrm{tr}(f_k)\rvert ,
\end{align}\] as required for 38 .
Case 2: \(d_0>d_n+d_m\) (see right-hand side of Figure 7)
In this case we have that \(d_0=d_n+d_m+d_\mathbb{H}(\gamma_n^k,\ell_m^k)\), or equivalently \[\label{eq:32case32232eq1}
\cosh \tfrac{1}{2}d_0 = \cosh\left( \tfrac{1}{2}d_n+\tfrac{1}{2}d_m + \tfrac{1}{2}d_\mathbb{H}(\gamma_n^k,\ell_m^k)\right) \leqslant 4\;\cosh \tfrac{1}{2}d_n\;\cosh\tfrac{1}{2}d_m \;\cosh
\tfrac{1}{2}d_\mathbb{H}(\gamma_n^k,\ell_m^k).\tag{40}\] Let \(H_n^k\subseteq\mathbb{H}\) be the open hyperbolic half-plane that is bounded by the geodesic \(\gamma_n^k\) and
contains \(\ell_n^k\). Similarly, \(H_m^k\) denotes the open hyperbolic half-plane bounded by \(\ell_m^k\), that contains \(\gamma_m^k\). Note that the half-planes \(H_n^k, H_m^k\) have disjoint closures, and \[\label{eq:32case32232eq2}
\inf\left\{d_\mathbb{H}(z,w)\colon z\in H_n^k\;\text{and} \;w\in H_m^k\right\}=d_\mathbb{H}(\gamma_n^k,\ell_m^k).\tag{41}\] Moreover, by construction of the intervals \(U_n^k, S_n^k,U_m^k, S_m^k\), we have
that \(\mathrm{Ax}(f_n)\subseteq H_n^k\), and similarly \(\mathrm{Ax}(f_m)\subseteq H_m^k\). Therefore the axes of \(f_n\) and \(f_m\) are disjoint, and \[\label{eq:32case32232eq3}
d_\mathbb{H}\left( \mathrm{Ax}(f_n), \mathrm{Ax}(f_m)\right)<\inf\left\{d_\mathbb{H}(z,w)\colon z\in H_n^k\;\text{and} \;w\in H_m^k\right\}.\tag{42}\] So, due to 41 and 42 , inequality 40 yields \[\begin{align}
\cosh \tfrac{1}{2}d_0 &< 4\;\cosh \tfrac{1}{2}d_n\;\cosh\tfrac{1}{2}d_m \;\cosh\tfrac{1}{2}d_\mathbb{H}\left( \mathrm{Ax}(f_n), \mathrm{Ax}(f_m)\right)\nonumber \\ &\leqslant 4\;\cosh \tfrac{1}{2}d_n\;\cosh\tfrac{1}{2}d_m \;\cosh
d_\mathbb{H}\left( \mathrm{Ax}(f_n), \mathrm{Ax}(f_m)\right). \label{eq:32case32232eq4}
\end{align}\tag{43}\] But, according to Corollary 2, we have that \(\cosh d_\mathbb{H}\left( \mathrm{Ax}(f_n), \mathrm{Ax}(f_m)\right) =
\frac{C(f_n,f_m)+1}{\lvert C(f_n,f_m)-1\rvert}\). Substituting this to 43 and using 32 , 39 yield \[\begin{align}
\cosh \tfrac{1}{2}d_0 \leqslant 4\;\max_{i\neq j}\{M_{i,j}\}\cdot \varepsilon\cdot \frac{C(f_n,f_m)+1}{\lvert C(f_n,f_m)-1\rvert}<\tfrac{1}{2}\lvert\textrm{tr}(f_k)\rvert,
\end{align}\] and we are once more led to 38 . ◻