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

Renormalization of Translated Cone Exchange Transformations

Noah Cockram1, Peter Ashwin1 and Ana Rodrigues1,2,3 1 Department of Mathematics and Statistics, University of Exeter, Exeter EX4 4QE, UK 2 Departamento de Matemática, Escola de Ciências e Tecnologia, Universidade de Évora, Rua Romão Ramalho, 59, 7000–671 Évora, Portugal 3 Centro de Investigação em Matemática e Aplicações, Rua Romão Ramalho, 59, 7000–671 Évora, Portugal
Abstract.

In this paper, we investigate a class of non-invertible piecewise isometries on the upper half-plane known as Translated Cone Exchanges. These maps include a simple interval exchange on a boundary we call the baseline. We provide a geometric construction for the first return map to a neighbourhood of the vertex of the middle cone for a large class of parameters, then we show a recurrence in the first return map tied to Diophantine properties of the parameters, and subsequently prove the infinite renormalizability of the first return map for these parameters.

1. Introduction

Piecewise isometries (PWIs) are a class of maps that can be generally described as a “cutting-and-shuffling” action of a metric space, specifically a partitioning of the phase space into at most countably many convex pieces called atoms, which are each moved according to an isometry. The phase space of these maps can be partitioned into two (or three) subsets based on the dynamics – a polygon or disc packing of periodic islands known as the regular set, and its complement, the set of points whose orbit either lands on, or accumulates on, the discontinuity set. Some authors choose to further distinguish those points in the pre-images of the discontinuity and those points which accumulate on it. The most well-known and well-understood examples of such maps are the interval exchange transformations (IETs), which arise as return maps to cross-sections of some measured foliations [1] and also as generalisations of circle rotations [2, 3, 4] and their encoding spaces generalise Sturmian shifts [5]. Furthermore, interval exchanges which aren’t irrational rotations are known to be almost always weakly mixing [6] but never strongly mixing [7]. Piecewise isometries in general, however, are not as well-known and as a subset of this class, interval exchanges are in many ways exceptional, due in part to being one-dimensional, as well as the invariance of Lebesgue measure.

In the more general setting, although the inherent lack of hyperbolicity restricts the variety of possible behaviours, for example it is known that all piecewise isometries have zero topological entropy [8], piecewise isometries are still capable of quite complex behaviour; many examples show the presence of unbounded periodicity and an underlying renormalizability which structures the dynamics near the discontinuities [9, 10, 11, 12, 13, 14, 15]; numerical evidence suggests the existence of invariant curves in the exceptional set which seem fractal-like and form barriers to ergodicity [13, 14, 16, 17]; there are conjectured conditions for piecewise isometries to have sensitive dependence on initial conditions [18].

Renormalization in theoretical physics and nonlinear dynamical systems has a longstanding history, see for example [19, 20, 21, 22, 23, 24, 25], driven by the problem of understanding phenomena that occur simultaneously at many spatial and temporal scales, particularly near phase transitions, periodic points, or in the case of piecewise isometries, the set of discontinuities.

In this paper, we investigate the renormalizability of a class of piecewise isometries called Translated Cone Exchanges on the closure of the upper half-plane ¯\overline{\mathbb{H}}. This family of maps was introduced in [16] and has since been investigated in [26, 28]. In particular, we use a geometric construction to describe the action of a first return map to a subset containing the origin, and show that this map displays renormalizable behaviour locally to the origin in accordance with Diophantine approximation of one of its parameters. These results go beyond [26, 28] in that they are much less constrained in the continued fraction expansion associated with the baseline translation.

This paper is organized as follows. In Section 2, we introduce the family of maps we will investigate, namely, Translated Cone Exchange transformations. In Section 3 we will develop some tools that will be useful in the next section. Section 4 presents the preliminary results that lead to the main result of this paper, Theorem 4.7, which gives an explicit form of renormalization for the first return maps of maps in our class to a neighbourhood of 0. Finally, in Section 5 we present an example for fixed values of the parameters.

2. Translated Cone Exchange transformations

Let \mathbb{H}\subset\mathbb{C} denote the upper half plane, and let ¯\overline{\mathbb{H}} be its closure in \mathbb{C}, that is

¯={z:Im(z)0}.\overline{\mathbb{H}}=\{z\in\mathbb{C}:\mathrm{Im}(z)\geq 0\}.

A Translated Cone Exchange transformation (TCE) [16] is a PWI (𝒞,Fκ)(\mathcal{C},F_{\kappa}) defined on the closed upper half plane ¯\overline{\mathbb{H}}. For any integer d>0d>0, let 𝔹d+2\mathbb{B}^{d+2} be the set

𝔹d+2={α=(α0,,αd+1)(0,π)d+2:j=0d+1αj=π},\mathbb{B}^{d+2}=\left\{\alpha=(\alpha_{0},...,\alpha_{d+1})\in(0,\pi)^{d+2}:\sum_{j=0}^{d+1}\alpha_{j}=\pi\right\},

Next, for some α=(α0,,αd+1)𝔹d+2\alpha=(\alpha_{0},...,\alpha_{d+1})\in\mathbb{B}^{d+2}, partition the interval [0,π][0,\pi] by subintervals

Wj={[0,α0), if j=0,[α0,α0+α1], if j=1,(k=0j1αk,k=0jαk], if j{2,,d+1}.W_{j}=\begin{cases}[0,\alpha_{0}),&\text{ if }j=0,\\[4.2679pt] [\alpha_{0},\alpha_{0}+\alpha_{1}],&\text{ if }j=1,\\[4.2679pt] \left(\sum\limits_{k=0}^{j-1}\alpha_{k},\sum_{k=0}^{j}\alpha_{k}\right],&\text{ if }j\in\{2,...,d+1\}.\end{cases}

We then define the partition 𝒞\mathcal{C} as

𝒞={Cj:j{0,,d+1}},\mathcal{C}=\left\{C_{j}:j\in\{0,...,d+1\}\right\},

where

C1={0}{z:ArgzW1}\displaystyle C_{1}=\{0\}\cup\{z\in\mathbb{H}:\operatorname{Arg}z\in W_{1}\}
and
Cj={z¯:ArgzWj} for j1.\displaystyle C_{j}=\{z\in\overline{\mathbb{H}}:\operatorname{Arg}z\in W_{j}\}\text{ for $j\neq 1$}.
Refer to caption
Figure 1. An example of a partition of the closed upper half plane ¯\overline{\mathbb{H}} into 6 cones.

The mapping FκF_{\kappa} is defined as a composition

(2.1) Fκ(z)=GE(z),F_{\kappa}(z)=G\circ E(z),

where EE is a permutation of the cones C1,,CdC_{1},...,C_{d}, GG is a piecewise horizontal translation, and κ\kappa is a tuple of the parameters. Formally, let τ\tau be a permutation of {1,,d}\{1,...,d\}, that is a bijection τ:{1,,d}{1,,d}\tau:\{1,...,d\}\rightarrow\{1,...,d\}, and let

θj(α,τ)=τ(k)<τ(j)αkk<jαk.\theta_{j}(\alpha,\tau)=\sum_{\tau(k)<\tau(j)}\alpha_{k}-\sum_{k<j}\alpha_{k}.

When α\alpha and τ\tau are unambiguous, we may refer to θj(α,τ)\theta_{j}(\alpha,\tau) simply as θj\theta_{j}. The map E:¯¯E:\overline{\mathbb{H}}\rightarrow\overline{\mathbb{H}} is then defined as

E(z)={z if zC0Cd+1zeiθj if zCj,j{1,,d}.E(z)=\begin{cases}z&\text{ if }z\in C_{0}\cup C_{d+1}\\ ze^{i\theta_{j}}&\text{ if }z\in C_{j},j\in\{1,...,d\}\end{cases}.
Refer to caption
Figure 2. The same partition from figure 1 after applying the cone exchange map EE.

Note that EE is invertible Lebesgue-almost everywhere in ¯\overline{\mathbb{H}}. We define the middle cone CcC_{c} of FκF_{\kappa} as

Cc=j=1dCj=(C0Cd+1).C_{c}=\bigcup_{j=1}^{d}C_{j}=\mathbb{H}\setminus(C_{0}\cup C_{d+1}).

The map G:¯¯G:\overline{\mathbb{H}}\rightarrow\overline{\mathbb{H}} is defined as

(2.2) G(z)={z+λ if zCd+1,zη if zCc,zρ if zC0,G(z)=\begin{cases}z+\lambda&\text{ if }z\in C_{d+1},\\ z-\eta&\text{ if }z\in C_{c},\\ z-\rho&\text{ if }z\in C_{0},\end{cases}

where ρ,λ(0,)\rho,\lambda\in(0,\infty) are rationally independent, i.e. λ/ρ\lambda/\rho\in\mathbb{R}\setminus\mathbb{Q}, and λ<η<ρ-\lambda<\eta<\rho. Finally, we collect the parameters into the tuple κ=(α,τ,λ,η,ρ)\kappa=(\alpha,\tau,\lambda,\eta,\rho). See figures 4, 5 and 9 for plots of the orbit structure for FκF_{\kappa} for example choices of parameters κ\kappa.

Refer to caption
Figure 3. The example partition from figure 1 after the TCE F=GEF=G\circ E is performed. Note the overlapping of the cones in a region containing 0.

For a>0a>0, let FκF_{\kappa^{\prime}} denote the TCE with parameters κ=(α,τ,λ/a,η/a,ρ/a)\kappa^{\prime}=(\alpha,\tau,\lambda/a,\eta/a,\rho/a). Define sa:¯¯s_{a}:\overline{\mathbb{H}}\rightarrow\overline{\mathbb{H}} as uniform scaling by aa about the origin

(2.3) sa(z)=az.s_{a}(z)=az.
Proposition 2.1.

We have the following conjugacy:

(2.4) Fκ(z)=sa1Fκsa(z).F_{\kappa^{\prime}}(z)=s_{a}^{-1}\circ F_{\kappa}\circ s_{a}(z).
Proof.

Firstly, observe that for all j{0,,d+1}j\in\{0,...,d+1\},

aCj=Cj,aC_{j}=C_{j},

from which we can deduce that

(2.5) azCj if and only if zCj.az\in C_{j}\text{ if and only if }z\in C_{j}.

Let z¯z\in\overline{\mathbb{H}}. From (2.3), we get

sa1Fκsa(z)=1aFκ(az),s_{a}^{-1}\circ F_{\kappa}\circ s_{a}(z)=\frac{1}{a}F_{\kappa}(az),

and by expanding FκF_{\kappa} as in (2.1), we have

sa1Fκsa(z)={1a(az+λ) if azCd+11a(eiθj(az)η) if azCj,j{1,,d}1a(azρ) if azC0.s_{a}^{-1}\circ F_{\kappa}\circ s_{a}(z)=\begin{cases}\frac{1}{a}(az+\lambda)&\text{ if }az\in C_{d+1}\\ \frac{1}{a}(e^{i\theta_{j}}(az)-\eta)&\text{ if }az\in C_{j},j\in\{1,...,d\}\\ \frac{1}{a}(az-\rho)&\text{ if }az\in C_{0}\end{cases}.

Recalling (2.5), distributing the multiplication by 1/a1/a, we finally see that

sa1Fκsa(z)\displaystyle s_{a}^{-1}\circ F_{\kappa}\circ s_{a}(z) ={z+λ/a if zCd+1eiθjzη/a if zCj,j{1,,d}zρ/a if zC0\displaystyle=\begin{cases}z+\lambda/a&\text{ if }z\in C_{d+1}\\ e^{i\theta_{j}}z-\eta/a&\text{ if }z\in C_{j},j\in\{1,...,d\}\\ z-\rho/a&\text{ if }z\in C_{0}\end{cases}
=Fκ(z).\displaystyle=F_{\kappa^{\prime}}(z).

Clearly from this proposition, we can normalize ρ=1\rho=1 without any loss of generality. Indeed, normalizing in this way proves very helpful for establishing recurrence results, as we shall see.

Refer to caption
Figure 4. A plot of the first 3000 elements of the forward orbits of 1000 points (omitting the first 1500 to remove transients) chosen uniformly in the box [ρ,λ]×[0,1][-\rho,\lambda]\times[0,1] under a TCE with parameters α=(0.5,π/7,π/4,17π/280.5)\alpha=(0.5,\pi/7,\pi/4,17\pi/28-0.5), τ:12,21\tau:1\mapsto 2,2\mapsto 1, λ=2/2\lambda=\sqrt{2}/2, η=1λ\eta=1-\lambda and ρ=1\rho=1. Each orbit is given a (non-unique) colour to illustrate the trajectories.

Simulations of the orbits of points under some TCEs appears to reveal complex behaviour even at small scales close to the real line, such as in figures 4 and 5. One way to investigate this behaviour is by applying the tools of renormalization. In particular, let h:Cc{0}h:C_{c}\rightarrow\mathbb{N}\setminus\{0\} denote the first return time of zCcz\in C_{c} to CcC_{c} under FκF_{\kappa}, that is

(2.6) h(z)=inf{n>0:Fκn(z)Cc}.h(z)=\inf\left\{n>0:F_{\kappa}^{n}(z)\in C_{c}\right\}.

The first return map Rκ:CcCcR_{\kappa}:C_{c}\rightarrow C_{c} of FκF_{\kappa} to CcC_{c} is then defined as

(2.7) Rκ(z)=Fκh(z)(z).R_{\kappa}(z)=F_{\kappa}^{h(z)}(z).

Observe that for all zCcz\in C_{c}, Rκ(z)=Fκh(z)(z)=Gh(z)E(z)R_{\kappa}(z)=F_{\kappa}^{h(z)}(z)=G^{h(z)}\circ E(z), since EE is the identity outside of CcC_{c}.

We will now state the main theorem of our paper in a simplified form, which we shall restate in more detail later as Theorem 4.7, after establishing some terminology and preliminary results.

Theorem 2.2.

Let α𝔹d+2\alpha\in\mathbb{B}^{d+2}, τ:{1,,d}{1,,d}\tau:\{1,...,d\}\rightarrow\{1,...,d\} be a bijection, λ[0,1)\lambda\in[0,1)\setminus\mathbb{Q}, λ<η=pqλ<1-\lambda<\eta=p-q\lambda<1 for some p,q{0}p,q\in\mathbb{N}\setminus\{0\} and set κ=(α,τ,λ,η,1)\kappa=(\alpha,\tau,\lambda,\eta,1). Then there exist λ[0,1)\lambda^{\prime}\in[0,1)\setminus\mathbb{Q}, η\eta^{\prime}\in\mathbb{R} of the form λ<η=pqλ<1-\lambda^{\prime}<\eta^{\prime}=p^{\prime}-q^{\prime}\lambda^{\prime}<1 for some p,q{0}p^{\prime},q^{\prime}\in\mathbb{N}\setminus\{0\}, and a convex, positive area set VCcV\subset C_{c} containing 0 such that

Rκ(z)=11λ1λRκ((1λ1λ)z),R_{\kappa^{\prime}}(z)=\frac{1}{1-\lambda_{1}\lambda}R_{\kappa}((1-\lambda_{1}\lambda)z),

for all zVz\in V, where κ=(α,τ,λ,η,1)\kappa^{\prime}=(\alpha,\tau,\lambda^{\prime},\eta^{\prime},1).

Refer to caption
Figure 5. A similar plot to figure 4, this time of the first 3000 elements (excluding the first 1500 for transients) of the forward orbits of 1500 points chosen uniformly in the box [ρ,λ]×[0,1][-\rho,\lambda]\times[0,1] under a TCE with parameters α=(π/2+0.1,π/8,0.2,π/50.1,7π/400.2)\alpha=(\pi/2+0.1,\pi/8,0.2,\pi/5-0.1,7\pi/40-0.2), τ:13,22,31\tau:1\mapsto 3,2\mapsto 2,3\mapsto 1, λ=π/4\lambda=\pi/4, and η=1λ\eta=1-\lambda. Note that the phase space is very sparse compared to figures 4 and 9.

3. Tools

In this section we prove some preliminary results that will serve as tools for more detailed investigation of the renormalization of TCEs. Firstly, we note the following smaller observations.

Proposition 3.1.

Let FκF_{\kappa} be as in (2.1) and GG as in (2.2). Then for all xx\in\mathbb{R},

Fκ(x)=G(x).F_{\kappa}(x)=G(x).
Proof.

This is clear from the observation that EE is the identity on \mathbb{R}. ∎

Proposition 3.2.

The dynamics of FκF_{\kappa} on \mathbb{R} and on \mathbb{H} are separate in the sense that Fκ()F_{\kappa}(\mathbb{R})\subset\mathbb{R} and Fκ()F_{\kappa}(\mathbb{H})\subset\mathbb{H}.

Proof.

The statement Fκ()F_{\kappa}(\mathbb{R})\subset\mathbb{R} is clear from Proposition 3.1 and the fact that GG is a horizontal translation.

We now prove that Fκ()F_{\kappa}(\mathbb{H})\subset\mathbb{H}. Suppose that zz\in\mathbb{H} such that Fκ(z)F_{\kappa}(z)\in\mathbb{R}. Then E(z)E(z)\in\mathbb{R}, as GG is a horizontal translation. If zCcz\in\mathbb{H}\setminus C_{c} then E(z)=z\mathbb{R}\ni E(z)=z\in\mathbb{H}, which is a contradiction. On the other hand, if zCcz\in C_{c}, then E(z)CcE(z)\in C_{c} But Cc={0}\mathbb{R}\cap C_{c}=\{0\}, meaning that E(z)E(z)\in\mathbb{R} implies E(z)=0E(z)=0. But this is only the case if z=0z=0\in\mathbb{R}, which is a contradiction. ∎

Let ι:¯{1,0,1}\iota:\overline{\mathbb{H}}\rightarrow\{-1,0,1\}^{\mathbb{N}} denote the itinerary for FκF_{\kappa}, defined by

ι(z)=ι0ι1ι2,\iota(z)=\iota_{0}\iota_{1}\iota_{2}...,

where

ιn=ιn(z)={1, if Fκn(z)Cd+1,0, if Fκn(z)Cc,1, if Fκn(z)C0.\iota_{n}=\iota_{n}(z)=\begin{cases}-1,&\text{ if }F_{\kappa}^{n}(z)\in C_{d+1},\\ 0,&\text{ if }F_{\kappa}^{n}(z)\in C_{c},\\ 1,&\text{ if }F_{\kappa}^{n}(z)\in C_{0}.\end{cases}

This map is similar to, but distinct from the true notion of the encoding map, since here we do not distinguish between the cones C1C_{1}, C2C_{2}, …, CdC_{d} within the middle cone CcC_{c}. The next Lemma provides a crucial tool in the proof of Theorem 4.5, since the dynamics on the interval [ρ,λ)[-\rho,\lambda) is that of a rotation of the circle (except at the point 0, in which case Fκ(0)=ηF_{\kappa}(0)=-\eta), which is more easily understood than that of an arbitrary point in CcC_{c}.

Lemma 3.3.

Let α𝔹d+2\alpha\in\mathbb{B}^{d+2}, τ:{1,,d}{1,,d}\tau:\{1,...,d\}\rightarrow\{1,...,d\} be a bijection, λ,ρ\lambda,\rho\in\mathbb{R} such that λ/ρ\lambda/\rho\notin\mathbb{Q}, and λ<η<1-\lambda<\eta<1. If zCcz\in C_{c}, then for all 1jh(z)1\leq j\leq h(z),

Fκj(z)=E(z)+Fκj(0).F_{\kappa}^{j}(z)=E(z)+F_{\kappa}^{j}(0).
Proof.

Suppose not, for a contradiction. Then there is some nn with 0nh(z)0\leq n\leq h(z) such that Fκn(z)E(z)+Fκn(0)F_{\kappa}^{n}(z)\neq E(z)+F_{\kappa}^{n}(0), and without loss of generality assume that nn is the smallest such integer. Clearly n>1n>1, so for all 0jn10\leq j\leq n-1, Fκj(z)E(z)=Fκj(0)F_{\kappa}^{j}(z)-E(z)=F_{\kappa}^{j}(0), and therefore ιj(z)=ιj(0)\iota_{j}(z)=\iota_{j}(0) for all 0jn20\leq j\leq n-2, but ιn1(z)ιn1(0)\iota_{n-1}(z)\neq\iota_{n-1}(0). Since nh(z)n\leq h(z), we cannot have Fκj(z)CcF_{\kappa}^{j}(z)\in C_{c} for all 1jn11\leq j\leq n-1. Hence for addresses of the (n1)th(n-1)^{\text{th}} iterates of zz and 0 to disagree, one of two cases must occur:

  1. 1.

    Fκn1(z)Cd+1F_{\kappa}^{n-1}(z)\in C_{d+1} and Fκn1(0)CcC0F_{\kappa}^{n-1}(0)\in C_{c}\cup C_{0}; or

  2. 2.

    Fκn1(z)C0F_{\kappa}^{n-1}(z)\in C_{0} and Fκn1(0)CcCd+1F_{\kappa}^{n-1}(0)\in C_{c}\cup C_{d+1}.

Since the orbit of 0 is restricted to \mathbb{R} and since E(0)=0E(0)=0, the second parts of each case become Gn1(0)0G^{n-1}(0)\geq 0 and Gn1(0)0G^{n-1}(0)\leq 0, respectively.

Suppose E(z)=zE(z)=z^{\prime}, and let ε1=Im(z)cot(αd+1)\varepsilon_{1}=\mathrm{Im}(z^{\prime})\operatorname{cot}(\alpha_{d+1}) and ε2=Im(z)cot(α0)\varepsilon_{2}=\mathrm{Im}(z^{\prime})\operatorname{cot}(\alpha_{0}). Then zCcz^{\prime}\in C_{c} if and only if ε1Re(z)ε2-\varepsilon_{1}\leq\mathrm{Re}(z^{\prime})\leq\varepsilon_{2}. Note that since GG is a horizontal translation, Fκn1(z)=Gn1(E(z))=Gn1(z)Cd+1F_{\kappa}^{n-1}(z)=G^{n-1}(E(z))=G^{n-1}(z^{\prime})\in C_{d+1} if and only if Re(Gn1(z))<ε1\mathrm{Re}(G^{n-1}(z^{\prime}))<-\varepsilon_{1}. Similarly Gn1(z)C0G^{n-1}(z^{\prime})\in C_{0} if and only if Re(Gn1(z))>ε2\mathrm{Re}(G^{n-1}(z^{\prime}))>\varepsilon_{2}. The two above cases above can thus be reformulated as:

  1. 1.

    Re(Gn1(z))<ε1\mathrm{Re}(G^{n-1}(z^{\prime}))<-\varepsilon_{1} and Gn1(0)0G^{n-1}(0)\geq 0; or

  2. 2.

    Re(Gn1(z))>ε2\mathrm{Re}(G^{n-1}(z^{\prime}))>\varepsilon_{2} and Gn1(0)0G^{n-1}(0)\leq 0.

In case 1, we get 0Gn1(0)=Gn1(z)z=Re(Gn1(z))Re(z)0\leq G^{n-1}(0)=G^{n-1}(z^{\prime})-z^{\prime}=\mathrm{Re}(G^{n-1}(z^{\prime}))-\mathrm{Re}(z^{\prime}). Hence Re(Gn1(z))Re(z)\mathrm{Re}(G^{n-1}(z))\geq\mathrm{Re}(z^{\prime}) and thus

ε1Re(z)Re(Gn1(z))<ε1,-\varepsilon_{1}\leq\mathrm{Re}(z^{\prime})\leq\mathrm{Re}(G^{n-1}(z^{\prime}))<-\varepsilon_{1},

which is a contradiction. Case 2 leads to a similar contradiction. Therefore there is no such nn. ∎

3.1. Continued Fractions

Recall from the theory of continued fractions that the nthn^{\text{th}} convergent to a positive, irrational real number λ=[λ0;λ1,λ2,]\lambda=[\lambda_{0};\lambda_{1},\lambda_{2},...] is a fraction pn/qn=[λ0;λ1,,λn]p_{n}/q_{n}=[\lambda_{0};\lambda_{1},...,\lambda_{n}], where pn,qnp_{n},q_{n} are coprime integers and qn>0q_{n}>0. The numbers pn,qnp_{n},q_{n} can be generated by the recursive relations:

(3.1) p0\displaystyle p_{0} =λ0,\displaystyle=\lambda_{0}, q0=1,\displaystyle q_{0}=1,
p1\displaystyle p_{1} =λ1λ0+1,\displaystyle=\lambda_{1}\lambda_{0}+1, q1=λ1,\displaystyle q_{1}=\lambda_{1},
pn\displaystyle p_{n} =λnpn1+pn2,\displaystyle=\lambda_{n}p_{n-1}+p_{n-2}, qn=λnqn1+qn2.\displaystyle q_{n}=\lambda_{n}q_{n-1}+q_{n-2}.

Furthermore, the convergents to λ\lambda satisfy the property that for all positive integers s<qn+1s<q_{n+1} and all rr\in\mathbb{Z},

|qnλpn||sλr|,|q_{n}\lambda-p_{n}|\leq|s\lambda-r|,

with equality only when (r,s)=(pn,qn)(r,s)=(p_{n},q_{n}). Also observe that we can use the recurrence relation in (3.1) to set

(3.2) p1\displaystyle p_{-1} =1\displaystyle=1 q1=0.\displaystyle q_{-1}=0.

Let g:[0,1][0,1]g:[0,1]\rightarrow[0,1] denote the Gauss map, given by

g(x)=1x1x.g(x)=\frac{1}{x}-\left\lfloor\frac{1}{x}\right\rfloor.

In particular, if λ=[0;λ1,λ2,λ3,][0,1]\lambda=[0;\lambda_{1},\lambda_{2},\lambda_{3},...]\in[0,1], then

g(λ)=1λλ1=[0;λ2,λ3,].g(\lambda)=\frac{1}{\lambda}-\lambda_{1}=[0;\lambda_{2},\lambda_{3},...].

Let λ=[0;λ1,λ2,][0,1)\lambda=[0;\lambda_{1},\lambda_{2},...]\in[0,1)\setminus\mathbb{Q}. To start, let

𝒩λ={(m,n)2:0nλm+1},\mathcal{N}_{\lambda}=\{(m,n)\in\mathbb{N}^{2}:0\leq n\leq\lambda_{m+1}\},

and define the function wλ:𝒩λw_{\lambda}:\mathcal{N}_{\lambda}\rightarrow\mathbb{N} by

wλ(m,n)={n if m=0,λ1++λm+n if m>0.w_{\lambda}(m,n)=\begin{cases}n&\text{ if }m=0,\\ \lambda_{1}+...+\lambda_{m}+n&\text{ if }m>0.\end{cases}

Note that wλw_{\lambda} is surjective and in fact

(3.3) wλ(m+1,0)=wλ(m,λm+1).w_{\lambda}(m+1,0)=w_{\lambda}(m,\lambda_{m+1}).

Furthermore, if we define the subset 𝒩λ<𝒩λ\mathcal{N}_{\lambda}^{<}\subset\mathcal{N}_{\lambda} to be

𝒩λ<={(m,n)2:0n<λm+1},\mathcal{N}_{\lambda}^{<}=\{(m,n)\in\mathbb{N}^{2}:0\leq n<\lambda_{m+1}\},

then wλ|𝒩λ<w_{\lambda}|_{\mathcal{N}_{\lambda}^{<}} is a bijection.

From now on, we denote the jthj^{\text{th}} coefficient of the continued fraction expansion of gm(λ)g^{m}(\lambda) by gm(λ)jg^{m}(\lambda)_{j}. The next proposition gives us a nice relationship between the set 𝒩λ\mathcal{N}_{\lambda} and the Gauss map gg.

Proposition 3.4.

Let j,m,nj,m,n\in\mathbb{N}. Then (m+j,n)𝒩λ(m+j,n)\in\mathcal{N}_{\lambda} if and only if (j,n)𝒩gm(λ)(j,n)\in\mathcal{N}_{g^{m}(\lambda)}. Moreover,

wλ(m+j,n)=λ1++λm+wgm(λ)(j,n).w_{\lambda}(m+j,n)=\lambda_{1}+...+\lambda_{m}+w_{g^{m}(\lambda)}(j,n).
Proof.

We have that (m+j,n)𝒩λ(m+j,n)\in\mathcal{N}_{\lambda} is equivalent to 0nλm+j+10\leq n\leq\lambda_{m+j+1}. We also have that λm+j+1=gm(λ)j+1\lambda_{m+j+1}=g^{m}(\lambda)_{j+1}, so 0ngm(λ)j+10\leq n\leq g^{m}(\lambda)_{j+1}. This is equivalent to (j,n)𝒩gm(λ)(j,n)\in\mathcal{N}_{g^{m}(\lambda)}.

If m=j=0m=j=0, then the second part of our lemma is clearly true.

Assume j=0j=0 and m>0m>0. Then

wλ(m+j,n)=wλ(m,n)=λ1++λm+n=λ1++λm+wgm(λ)(0,n),w_{\lambda}(m+j,n)=w_{\lambda}(m,n)=\lambda_{1}+...+\lambda_{m}+n=\lambda_{1}+...+\lambda_{m}+w_{g^{m}(\lambda)}(0,n),

where the final equality is true since (m,n)𝒩λ(m,n)\in\mathcal{N}_{\lambda} is equivalent to (0,n)𝒩gm(λ)(0,n)\in\mathcal{N}_{g^{m}(\lambda)}.

Finally, suppose m,j>0m,j>0. Then

wλ(m+j,n)\displaystyle w_{\lambda}(m+j,n) =λ1++λm+λm+1++λm+j+n\displaystyle=\lambda_{1}+...+\lambda_{m}+\lambda_{m+1}+...+\lambda_{m+j}+n
=λ1++λm+gm(λ)1++gm(λ)j+n.\displaystyle=\lambda_{1}+...+\lambda_{m}+g^{m}(\lambda)_{1}+...+g^{m}(\lambda)_{j}+n.

Note that since (m+j,n)𝒩λ(m+j,n)\in\mathcal{N}_{\lambda} is equivalent to (j,n)𝒩gm(λ)(j,n)\in\mathcal{N}_{g^{m}(\lambda)}, we have

wλ(m+j,n)=λ1++λm+wgm(λ)(j,n).w_{\lambda}(m+j,n)=\lambda_{1}+...+\lambda_{m}+w_{g^{m}(\lambda)}(j,n).

The bijection wλw_{\lambda} mainly serves as a way to show that there is a “natural” well-ordering for the set 𝒩λ<\mathcal{N}_{\lambda}^{<}, which allows us to meaningfully index sequences by 𝒩λ<\mathcal{N}_{\lambda}^{<} and, as we shall see later, define the notion of a maximal element of a finite subset of 𝒩λ\mathcal{N}_{\lambda}.

We define the one-sided convergents (or semiconvergents) to λ\lambda as the fractions

Pm,n(λ)Qm,n(λ)={[0;λ1,,λm] if n=0,[0;λ1,,λm,n] if n>0,\frac{P_{m,n}(\lambda)}{Q_{m,n}(\lambda)}=\begin{cases}[0;\lambda_{1},...,\lambda_{m}]&\text{ if }n=0,\\ [0;\lambda_{1},...,\lambda_{m},n]&\text{ if }n>0,\end{cases}

Indeed, this formula is compatible with the indexing wλw_{\lambda} in that

Pm,0(λ)Qm,0(λ)=Pm1,λm(λ)Qm1,λm(λ).\frac{P_{m,0}(\lambda)}{Q_{m,0}(\lambda)}=\frac{P_{m-1,\lambda_{m}}(\lambda)}{Q_{m-1,\lambda_{m}}(\lambda)}.

A standard result in the theory of continued fractions is that for (m,n)𝒩λ<(m,n)\in\mathcal{N}_{\lambda}^{<}, we have

(3.4) Pm,n(λ)Qm,n(λ)={npm(λ)+pm1(λ)nqm(λ)+qm1(λ), if n0pm(λ)qm(λ) if n=0.\frac{P_{m,n}(\lambda)}{Q_{m,n}(\lambda)}=\begin{cases}\frac{np_{m}(\lambda)+p_{m-1}(\lambda)}{nq_{m}(\lambda)+q_{m-1}(\lambda)},&\text{ if }n\neq 0\\[5.69054pt] \frac{p_{m}(\lambda)}{q_{m}(\lambda)}&\text{ if }n=0.\end{cases}

One way to interpret these fractions is as being the best rational approximates of λ\lambda from one “direction”. In particular, borrowing notation from the beginning of section 2.2 in [28], we have

(Pm,nQm,n)(m,n)𝒩λ<,m even\displaystyle\left(\frac{P_{m,n}}{Q_{m,n}}\right)_{(m,n)\in\mathcal{N}_{\lambda}^{<},m\text{ even}} =(pkqk)k, and\displaystyle=\left(\frac{p_{k}^{\prime}}{q_{k}^{\prime}}\right)_{k\in\mathbb{N}},\text{ and}
(Pm,nQm,n)(m,n)𝒩λ<,m odd\displaystyle\left(\frac{P_{m,n}}{Q_{m,n}}\right)_{(m,n)\in\mathcal{N}_{\lambda}^{<},m\text{ odd}} =(pk′′qk′′)k,\displaystyle=\left(\frac{p_{k}^{\prime\prime}}{q_{k}^{\prime\prime}}\right)_{k\in\mathbb{N}},

where (pk/qk)k(p_{k}^{\prime}/q_{k}^{\prime})_{k} are the best rational approximates from above in the sense that pk/qk>λp_{k}^{\prime}/q_{k}^{\prime}>\lambda and for all rational numbers r/spk/qkr/s\neq p_{k}^{\prime}/q_{k}^{\prime} such that r/s>λr/s>\lambda and 1s<qk+11\leq s<q_{k+1}^{\prime}, we have

(3.5) 0<|qkλpk|<|sλr|,0<|q_{k}^{\prime}\lambda-p_{k}^{\prime}|<|s\lambda-r|,

and in a similar fashion (pk′′/qk′′)k(p_{k}^{\prime\prime}/q_{k}^{\prime\prime})_{k} are the best rational approximates from below in the same sense except that pk′′/qk′′<λp_{k}^{\prime\prime}/q_{k}^{\prime\prime}<\lambda and

(3.6) 0<|qk′′λpk′′|<|sλr|,0<|q_{k}^{\prime\prime}\lambda-p_{k}^{\prime\prime}|<|s\lambda-r|,

holds for r/spk′′/qk′′r/s\neq p_{k}^{\prime\prime}/q_{k}^{\prime\prime} such that r/s<λr/s<\lambda and 1s<qk+1′′1\leq s<q_{k+1}^{\prime\prime}.

We define the sequence (Δm,n(λ))(m,n)𝒩λ(\Delta_{m,n}(\lambda))_{(m,n)\in\mathcal{N}_{\lambda}} by

(3.7) Δm,n(λ)=Qm,n(λ)λPm,n(λ).\Delta_{m,n}(\lambda)=Q_{m,n}(\lambda)\lambda-P_{m,n}(\lambda).

We see immediately from the above discussion that Pm,n(λ)/Qm,n(λ)<λP_{m,n}(\lambda)/Q_{m,n}(\lambda)<\lambda if and only if mm is odd and n>0n>0 or mm is even and n=0n=0, that is

(3.8) Δm,n(λ)>0 if and only if {m is even and n=0, orm is odd and n>0.\Delta_{m,n}(\lambda)>0\text{ if and only if }\begin{cases}\text{$m$ is even and $n=0$, or}\\ \text{$m$ is odd and $n>0$.}\end{cases}

By expanding the definitions of Pm,nP_{m,n} and Qm,nQ_{m,n}, we see

(3.9) Δm,n(λ)\displaystyle\Delta_{m,n}(\lambda) =n(qmλpm)+qm1λpm1\displaystyle=n(q_{m}\lambda-p_{m})+q_{m-1}\lambda-p_{m-1}
=nΔm,0(λ)+Δm1,0(λ),\displaystyle=n\Delta_{m,0}(\lambda)+\Delta_{m-1,0}(\lambda),

for (m,n)𝒩λ(m,n)\in\mathcal{N}_{\lambda} with m1m\geq 1. Moreover, by expanding the recurrence relation for pmp_{m} and qmq_{m} and rearranging terms, we have the additional property

Δm,0(λ)\displaystyle\Delta_{m,0}(\lambda) =qmλpm\displaystyle=q_{m}\lambda-p_{m}
=λm(qm1λpm1)+qm2λpm2\displaystyle=\lambda_{m}(q_{m-1}\lambda-p_{m-1})+q_{m-2}\lambda-p_{m-2}
=λmΔm1,0(λ)+Δm2,0(λ),\displaystyle=\lambda_{m}\Delta_{m-1,0}(\lambda)+\Delta_{m-2,0}(\lambda),

for m,m2m\in\mathbb{N},m\geq 2.

Note that in agreement with the function wλw_{\lambda}, we have Δm1,λm(λ)=Δm,0(λ)\Delta_{m-1,\lambda_{m}}(\lambda)=\Delta_{m,0}(\lambda). A result by Bates et al. [27] presents an interesting connection between iterates of the Gauss map and consecutive errors in the approximation of λ\lambda by its convergents.

Lemma 3.5 (Theorem 10 of [27]).

Let λ=[0;λ1,λ2,][0,1)\lambda=[0;\lambda_{1},\lambda_{2},...]\in[0,1)\setminus\mathbb{Q}. For all mm\in\mathbb{N},

(3.10) gm(λ)=qmλpmpm1qm1λ.g^{m}(\lambda)=\frac{q_{m}\lambda-p_{m}}{p_{m-1}-q_{m-1}\lambda}.

Equation (3.10) can be equivalently formulated as

(3.11) gm(λ)=Δm,0(λ)Δm1,0(λ).g^{m}(\lambda)=-\frac{\Delta_{m,0}(\lambda)}{\Delta_{m-1,0}(\lambda)}.
Lemma 3.6.

Let λ=[0;λ1,λ2,][0,1)\lambda=[0;\lambda_{1},\lambda_{2},...]\in[0,1)\setminus\mathbb{Q}. For all (m,n)𝒩λ(m,n)\in\mathcal{N}_{\lambda} with m1m\geq 1,

(3.12) Δ0,n(gm(λ))=Δm,n(λ)Δm1,0(λ).\Delta_{0,n}(g^{m}(\lambda))=-\frac{\Delta_{m,n}(\lambda)}{\Delta_{m-1,0}(\lambda)}.
Proof.

For n=0n=0, Δ0,0(gm(λ))=gm(λ)\Delta_{0,0}(g^{m}(\lambda))=g^{m}(\lambda), so (3.12) holds.

Next, observe that

Δj,0(gm(λ))=gm(λ)jΔj1,0(gm(λ))+Δj2,0(gm(λ)).\Delta_{j,0}(g^{m}(\lambda))=g^{m}(\lambda)_{j}\Delta_{j-1,0}(g^{m}(\lambda))+\Delta_{j-2,0}(g^{m}(\lambda)).

Since gm(λ)j=λm+jg^{m}(\lambda)_{j}=\lambda_{m+j}, this becomes

Δj,0(gm(λ))=λm+jΔj1,0(gm(λ))+Δj2,0(gm(λ)).\Delta_{j,0}(g^{m}(\lambda))=\lambda_{m+j}\Delta_{j-1,0}(g^{m}(\lambda))+\Delta_{j-2,0}(g^{m}(\lambda)).

We thus see that

Δ0,n(gm(λ))=n(q0gm(λ)p0)+q1gm(λ)p1,\Delta_{0,n}(g^{m}(\lambda))=n(q_{0}g^{m}(\lambda)-p_{0})+q_{-1}g^{m}(\lambda)-p_{-1},

and by recalling p1p_{-1}, q1q_{-1}, p0p_{0}, and q0q_{0} from (3.1), we get

Δ0,n(gm(λ))=ngm(λ)1.\Delta_{0,n}(g^{m}(\lambda))=ng^{m}(\lambda)-1.

Using (3.11), we can substitute gm(λ)g^{m}(\lambda) and rearrange terms to get

Δ0,n(gm(λ))\displaystyle\Delta_{0,n}(g^{m}(\lambda)) =n(Δm,0(λ)Δm1,0(λ)+1)\displaystyle=-n\left(\frac{\Delta_{m,0}(\lambda)}{\Delta_{m-1,0}(\lambda)}+1\right)
=nΔm,0(λ)+Δm1,0(λ)Δm1,0(λ).\displaystyle=-\frac{n\Delta_{m,0}(\lambda)+\Delta_{m-1,0}(\lambda)}{\Delta_{m-1,0}(\lambda)}.

Finally, from this and (3.9) we get (3.12). ∎

Corollary 3.7.

For all (m+j,n)𝒩λ(m+j,n)\in\mathcal{N}_{\lambda}, where m,jm,j\in\mathbb{N}, m1m\geq 1, we have

(3.13) Δ0,n(gm+j(λ))=Δm,n(gj(λ))Δm1,0(gj(λ))\Delta_{0,n}(g^{m+j}(\lambda))=-\frac{\Delta_{m,n}(g^{j}(\lambda))}{\Delta_{m-1,0}(g^{j}(\lambda))}
Proof.

This follows from Lemma 3.6 with gj(λ)g^{j}(\lambda) instead of λ\lambda. ∎

Our next Lemma is an important tool for determining scaling properties of these errors.

Lemma 3.8.

Let λ=[0;λ1,λ2,][0,1)\lambda=[0;\lambda_{1},\lambda_{2},...]\in[0,1)\setminus\mathbb{Q}. For all (m+j,n)𝒩λ(m+j,n)\in\mathcal{N}_{\lambda} such that m,jm,j\in\mathbb{N} and m1m\geq 1,

(3.14) Δj,n(gm(λ))=Δm+j,n(λ)Δm1,0(λ).\Delta_{j,n}(g^{m}(\lambda))=-\frac{\Delta_{m+j,n}(\lambda)}{\Delta_{m-1,0}(\lambda)}.
Proof.

Let us prove first that (3.14) holds for n=0n=0. By multiplying and dividing by Δm+k1,0\Delta_{m+k-1,0} for all 0kj0\leq k\leq j, we get

Δm+j,0(λ)Δm1,0(λ)=k=0jΔm+k,0(λ)Δm+k1,0(λ)\frac{\Delta_{m+j,0}(\lambda)}{\Delta_{m-1,0}(\lambda)}=\prod_{k=0}^{j}\frac{\Delta_{m+k,0}(\lambda)}{\Delta_{m+k-1,0}(\lambda)}

Then, using (3.11), we get

Δm+j,0(λ)Δm1,0(λ)=k=0jΔ0,0(gm+k(λ)).\frac{\Delta_{m+j,0}(\lambda)}{\Delta_{m-1,0}(\lambda)}=\prod_{k=0}^{j}-\Delta_{0,0}(g^{m+k}(\lambda)).

Rearranging this last expression and using (3.13), we get

Δm+j,0(λ)Δm1,0(λ)=(1)j+1Δ0,0(gm(λ))k=1jΔk,0(gm(λ))Δk1,0(gm(λ))\frac{\Delta_{m+j,0}(\lambda)}{\Delta_{m-1,0}(\lambda)}=(-1)^{j+1}\Delta_{0,0}(g^{m}(\lambda))\prod_{k=1}^{j}-\frac{\Delta_{k,0}(g^{m}(\lambda))}{\Delta_{k-1,0}(g^{m}(\lambda))}

We then simplify the product by cancelling terms in the numerator and denominator to get

Δm+j,0(λ)Δm1,0(λ)\displaystyle\frac{\Delta_{m+j,0}(\lambda)}{\Delta_{m-1,0}(\lambda)} =(1)j+1Δ0,0(gm(λ))((1)jΔj,0(gm(λ))Δ0,0(gm(λ)))\displaystyle=(-1)^{j+1}\Delta_{0,0}(g^{m}(\lambda))\left((-1)^{j}\frac{\Delta_{j,0}(g^{m}(\lambda))}{\Delta_{0,0}(g^{m}(\lambda))}\right)
=Δj,0(gm(λ)).\displaystyle=-\Delta_{j,0}(g^{m}(\lambda)).

Finally, for general (m+j,n)𝒩λ(m+j,n)\in\mathcal{N}_{\lambda}, m,jm,j\in\mathbb{N}, m1m\geq 1, we have

Δm+j,n(λ)Δm1,0(λ)=Δm+j,n(λ)Δm+j1,0(λ)Δm+j1,0(λ)Δm1,0(λ)\frac{\Delta_{m+j,n}(\lambda)}{\Delta_{m-1,0}(\lambda)}=\frac{\Delta_{m+j,n}(\lambda)}{\Delta_{m+j-1,0}(\lambda)}\frac{\Delta_{m+j-1,0}(\lambda)}{\Delta_{m-1,0}(\lambda)}

Using (3.12) and (3.14) for n=0n=0, we get

Δm+j,n(λ)Δm1,0(λ)=Δ0,n(gm+j(λ))Δj1,0(gm(λ)),\frac{\Delta_{m+j,n}(\lambda)}{\Delta_{m-1,0}(\lambda)}=\Delta_{0,n}(g^{m+j}(\lambda))\Delta_{j-1,0}(g^{m}(\lambda)),

and then using (3.13) gives us

Δm+j,n(λ)Δm1,0(λ)\displaystyle\frac{\Delta_{m+j,n}(\lambda)}{\Delta_{m-1,0}(\lambda)} =Δj,n(gm(λ))Δj1,0(gm(λ))Δj1,0(gm(λ))\displaystyle=-\frac{\Delta_{j,n}(g^{m}(\lambda))}{\Delta_{j-1,0}(g^{m}(\lambda))}\Delta_{j-1,0}(g^{m}(\lambda))
=Δj,n(gm(λ)),\displaystyle=-\Delta_{j,n}(g^{m}(\lambda)),

as required. ∎

With these properties in mind, we will now define the sets which will partition a neighbourhood of the middle cone CcC_{c}. Recall that 𝒩λ<\mathcal{N}_{\lambda}^{<} denotes the subset of 𝒩λ\mathcal{N}_{\lambda} defined by

𝒩λ<={(m,n)2:0n<λm+1}.\mathcal{N}_{\lambda}^{<}=\{(m,n)\in\mathbb{N}^{2}:0\leq n<\lambda_{m+1}\}.

For (m,n)𝒩λ<(m,n)\in\mathcal{N}_{\lambda}^{<}, let Sm,n(λ)S_{m,n}(\lambda) be the set defined by

(3.15) Sm,n(λ)={(C0Δm,0(λ))Cc(CcΔm,n+1(λ))(Cd+1(nΔm,0(λ)+Δm1,0(λ))),if m is even,(C0(nΔm,0(λ)+Δm1,0(λ)))Cc(CcΔm,n+1(λ))(Cd+1Δm,0(λ)),if m is odd.S_{m,n}(\lambda)=\begin{cases}\begin{aligned} &(C_{0}-\Delta_{m,0}(\lambda))\cap C_{c}\cap(C_{c}-\Delta_{m,n+1}(\lambda))\\ &\cap(C_{d+1}-(n\Delta_{m,0}(\lambda)+\Delta_{m-1,0}(\lambda)))\end{aligned},&\text{if }m\text{ is even},\\[17.07164pt] \begin{aligned} &(C_{0}-(n\Delta_{m,0}(\lambda)+\Delta_{m-1,0}(\lambda)))\cap C_{c}\\ &\cap(C_{c}-\Delta_{m,n+1}(\lambda))\cap(C_{d+1}-\Delta_{m,0}(\lambda))\end{aligned},&\text{if }m\text{ is odd}.\end{cases}

For brevity, we will drop the argument λ\lambda from Sm,n(λ)S_{m,n}(\lambda) if it is unambiguous. Additionally for the purposes of the case that m=0m=0, and recalling (3.2), we have

Δ1,0(λ)=q1λp1=1.\Delta_{-1,0}(\lambda)=q_{-1}\lambda-p_{-1}=-1.

Recall from (3.8) that for (m,n)𝒩λ<(m,n)\in\mathcal{N}_{\lambda}^{<}, Δm,n>0\Delta_{m,n}>0 if and only if mm is odd and n>0n>0 or mm is even and n=0n=0. Thus we can clearly see that Sm,nS_{m,n}\neq\emptyset for all (m,n)𝒩λ<(m,n)\in\mathcal{N}_{\lambda}^{<}. Additionally, since every point in Sm,nS_{m,n} has positive imaginary part, the boundary of Sm,nS_{m,n} consists of segments of the non-horizontal boundary lines of C0C_{0}, CcC_{c}, and Cd+1C_{d+1}, and all of these lines either have angle α0\alpha_{0} or παd+1\pi-\alpha_{d+1}. Thus, Sm,nS_{m,n} is a quadrilateral, and its opposing sides must be parallel, so it is a parallelogram.

Indeed, since opposite edges of Sm,nS_{m,n} are parallel, the side lengths of Sm,nS_{m,n} are uniquely determined by the horizontal distances between opposing edges. In the case that mm is even, these are precisely the distances between the vertices of the pairs of cones C0Δm,0C_{0}-\Delta_{m,0} and CcC_{c}, and CcΔm,n+1C_{c}-\Delta_{m,n+1} and Cd+1Δm,nC_{d+1}-\Delta_{m,n}. In the case that mm is odd, the horizontal distances are determined by the distance between the vertices of pairs of cones C0Δm,nC_{0}-\Delta_{m,n} and CcΔm,n+1C_{c}-\Delta_{m,n+1}, and CcC_{c} and Cd+1Δm,0C_{d+1}-\Delta_{m,0}. Since Δm,n+1Δm,n=Δm,0\Delta_{m,n+1}-\Delta_{m,n}=\Delta_{m,0}, we know that these distances are equal. Therefore, the side lengths of opposing edges of Sm,nS_{m,n} are equal and can be calculated as

(3.16) Δm,nsinα0sin(α0+αd+1) and Δm,nsinαd+1sin(α0+αd+1)\frac{\Delta_{m,n}\sin\alpha_{0}}{\sin(\alpha_{0}+\alpha_{d+1})}\text{ and }\frac{\Delta_{m,n}\sin\alpha_{d+1}}{\sin(\alpha_{0}+\alpha_{d+1})}

These sidelengths are equal only when α0=αd+1\alpha_{0}=\alpha_{d+1}, in which case Sm,nS_{m,n} is a rhombus for all (m,n)𝒩λ<(m,n)\in\mathcal{N}_{\lambda}^{<}. See figure 6 for an example of the geometry of the construction of sets Sm,nS_{m,n}.

Refer to caption
Figure 6. An illustration of the construction of the parallelogram S0,1(λ)S_{0,1}(\lambda) for the parameters in figure 4. Here, the angles shown indicate the cones used to construct S1,0(λ)S_{1,0}(\lambda). In this case, the vertices of these cones can be verified via (3.15) to be Δ0,0(λ)=λ-\Delta_{0,0}(\lambda)=-\lambda, Δ1,1(λ)=12λ-\Delta_{1,1}(\lambda)=1-2\lambda, 0 and (Δ0,0(λ)+Δ1,0)=1λ-(\Delta_{0,0}(\lambda)+\Delta_{-1,0})=1-\lambda.

An interesting property of these sets can be found by an application of Lemma 3.8.

Theorem 3.9.

Let λ[0,1)\lambda\in[0,1)\setminus\mathbb{Q}. For all (m+j,n)𝒩λ<(m+j,n)\in\mathcal{N}_{\lambda}^{<} such that m,jm,j\in\mathbb{N} and m2m\geq 2 is even,

1|Δm1,0(λ)|Sm+j,n(λ)=Sj,n(gm(λ)).\frac{1}{|\Delta_{m-1,0}(\lambda)|}S_{m+j,n}(\lambda)=S_{j,n}(g^{m}(\lambda)).
Proof.

Firstly, recall from (3.8) that since mm is even, Δm1,0(λ)<0\Delta_{m-1,0}(\lambda)<0, and thus

1|Δm1,0(λ)|=1Δm1,0(λ)>0.\frac{1}{|\Delta_{m-1,0}(\lambda)|}=\frac{1}{-\Delta_{m-1,0}(\lambda)}>0.

Hence,

1|Δm1,0(λ)|(Ckx)=Ck+xΔm1,0(λ),\frac{1}{|\Delta_{m-1,0}(\lambda)|}(C_{k}-x)=C_{k}+\frac{x}{\Delta_{m-1,0}(\lambda)},

for all k{1,,d}k\in\{1,...,d\} and all xx\in\mathbb{R}. Thus, we have

1|Δm1,0(λ)|Sm+j,n(λ)\displaystyle\frac{1}{|\Delta_{m-1,0}(\lambda)|}S_{m+j,n}(\lambda)
={(C0+Δm+j,0(λ)Δm1,0(λ))(Cc+Δm+j,n+1(λ)Δm1,0(λ))Cc(Cd+1+nΔm+j,0(λ)+Δm+j1,0(λ)Δm1,0(λ)),if m+j is even,(C0+nΔm+j,0(λ)+Δm+j1,0(λ)Δm1,0(λ))Cc(Cc+Δm+j,n+1(λ)Δm1,0(λ))(Cd+1+Δm+j,0(λ)Δm1,0(λ)),if m+j is odd.\displaystyle=\begin{cases}\begin{aligned} &\left(C_{0}+\frac{\Delta_{m+j,0}(\lambda)}{\Delta_{m-1,0}(\lambda)}\right)\cap\left(C_{c}+\frac{\Delta_{m+j,n+1}(\lambda)}{\Delta_{m-1,0}(\lambda)}\right)\\[5.69054pt] &\cap C_{c}\cap\left(C_{d+1}+\frac{n\Delta_{m+j,0}(\lambda)+\Delta_{m+j-1,0}(\lambda)}{\Delta_{m-1,0}(\lambda)}\right),\end{aligned}&\text{if }m+j\text{ is even,}\\[31.29802pt] \begin{aligned} &\left(C_{0}+\frac{n\Delta_{m+j,0}(\lambda)+\Delta_{m+j-1,0}(\lambda)}{\Delta_{m-1,0}(\lambda)}\right)\cap C_{c}\\[5.69054pt] &\cap\left(C_{c}+\frac{\Delta_{m+j,n+1}(\lambda)}{\Delta_{m-1,0}(\lambda)}\right)\cap\left(C_{d+1}+\frac{\Delta_{m+j,0}(\lambda)}{\Delta_{m-1,0}(\lambda)}\right),\end{aligned}&\text{if }m+j\text{ is odd.}\end{cases}

Using (3.14) many times give us

1|Δm1,0(λ)|Sm+j,n(λ)\displaystyle\frac{1}{|\Delta_{m-1,0}(\lambda)|}S_{m+j,n}(\lambda)
={(C0Δj,0(gm(λ)))(CcΔj,n+1(gm(λ)))Cc(Cd+1(nΔj,0(gm(λ))+Δj1,0(gm(λ)))),if j is even,(C0(nΔj,0(gm(λ))+Δj1,0(gm(λ))))Cc(CcΔj,n+1(gm(λ)))(Cd+1Δj,0(gm(λ))),if j is odd.\displaystyle=\begin{cases}\begin{aligned} &(C_{0}-\Delta_{j,0}(g^{m}(\lambda)))\cap(C_{c}-\Delta_{j,n+1}(g^{m}(\lambda)))\\ &\cap C_{c}\cap(C_{d+1}-(n\Delta_{j,0}(g^{m}(\lambda))+\Delta_{j-1,0}(g^{m}(\lambda)))),\end{aligned}&\text{if }j\text{ is even,}\\[14.22636pt] \begin{aligned} &(C_{0}-(n\Delta_{j,0}(g^{m}(\lambda))+\Delta_{j-1,0}(g^{m}(\lambda))))\cap C_{c}\\ &\cap(C_{c}-\Delta_{j,n+1}(g^{m}(\lambda)))\cap(C_{d+1}-\Delta_{j,0}(g^{m}(\lambda))),\end{aligned}&\text{if }j\text{ is odd.}\end{cases}

Finally, comparing this with (3.15) gets

1|Δm1,0(λ)|Sm+j,n(λ)=Sj,n(gm(λ)).\frac{1}{|\Delta_{m-1,0}(\lambda)|}S_{m+j,n}(\lambda)=S_{j,n}(g^{m}(\lambda)).

Refer to caption
Figure 7. An illustration of the EE-image of the partition for the first return map RκR_{\kappa} of the TCE FκF_{\kappa} in figure 4. Observe the alternating stacks of parallelograms decreasing in size and cascading towards the origin.

This theorem seems to suggest the possibility of infinite renormalizability of the first return maps to CcC_{c} for a whole class of TCEs. At the very least, if indeed the first return map of a TCE to CcC_{c} is an isometry on Sm,n(λ)S_{m,n}(\lambda) for (m,n)𝒩λ<(m,n)\in\mathcal{N}_{\lambda}^{<} with mm0m\geq m_{0} and some m0m_{0}\in\mathbb{N}, then at the very least the partition matches with its potential renormalization.

Refer to caption
Figure 8. A similar illustration to figure 7 of the EE-image of the partition for RκR_{\kappa}, but with the parameters from figure 5.

4. Renormalization around zero

In this section we investigate renormalizability of TCEs around the origin for a broad range of values of λ\lambda and η\eta.

Let λ[0,1)\lambda\in[0,1)\setminus\mathbb{Q} and η\eta\in\mathbb{R} such that λ<η=pqλ<1-\lambda<\eta=p-q\lambda<1 for some p,qp,q\in\mathbb{N}. Note that 𝒩λ<\mathcal{N}_{\lambda}^{<} has a well-ordering <<^{\prime} induced by the indexing function wλw_{\lambda} so that

(m,n)<(r,s) if and only if wλ(m,n)<wλ(r,s).(m,n)<^{\prime}(r,s)\text{ if and only if }w_{\lambda}(m,n)<w_{\lambda}(r,s).

Thus, the notion of a “maximal element” of a finite subset of 𝒩λ<\mathcal{N}_{\lambda}^{<} is well-defined.

Let (m0,n0)(m_{0},n_{0}) be the largest element of 𝒩λ<\mathcal{N}_{\lambda}^{<} such that

(4.1) Pm0,n0<p or Qm0,n0<q.P_{m_{0},n_{0}}<p\text{ or }Q_{m_{0},n_{0}}<q.

The pair (m0,n0)(m_{0},n_{0}) is well-defined since p,q1p,q\geq 1 but P0,0=0P_{0,0}=0. Thus, wλ(m0,n0)wλ(0,0)w_{\lambda}(m_{0},n_{0})\geq w_{\lambda}(0,0).

Note that for all (m,n)𝒩λ<(m,n)\in\mathcal{N}_{\lambda}^{<},

Δm,n=Qm,nλPm,n=η+(Qm,nq)λ(Pm,np).\Delta_{m,n}=Q_{m,n}\lambda-P_{m,n}=-\eta+(Q_{m,n}-q)\lambda-(P_{m,n}-p).

Define the sequence (hm,n)(m,n)𝒩λ<(h_{m,n})_{(m,n)\in\mathcal{N}_{\lambda}^{<}} of positive integers by

(4.2) hm,n=(Qm,nq)+(Pm,np)+1.h_{m,n}=(Q_{m,n}-q)+(P_{m,n}-p)+1.

We can establish some recurrence relations for the sequence hm,nh_{m,n} using those of Pm,nP_{m,n} and Qm,nQ_{m,n}.

Proposition 4.1.

Let (m,n)𝒩λ<(m,n)\in\mathcal{N}_{\lambda}^{<}. Then

hm,n+1=(n+1)hm,0+hm1,0+(n+1)(p+q1).h_{m,n+1}=(n+1)h_{m,0}+h_{m-1,0}+(n+1)(p+q-1).

Moreover, if wλ(m,n)>wλ(m0,n0)w_{\lambda}(m,n)>w_{\lambda}(m_{0},n_{0}), then hm,n>0h_{m,n}>0.

Proof.

Recall from the definition of Pm,nP_{m,n} and Qm,nQ_{m,n} in (3.4) that

Qm,n+1=(n+1)Qm,0+Qm1,0,Q_{m,n+1}=(n+1)Q_{m,0}+Q_{m-1,0},

and

Pm,n+1=(n+1)Pm,0+Pm1,0.P_{m,n+1}=(n+1)P_{m,0}+P_{m-1,0}.

By applying this to the formula (4.2), we get

hm,n+1\displaystyle h_{m,n+1} =(Qm,n+1q)+(Pm,n+1p)+1\displaystyle=(Q_{m,n+1}-q)+(P_{m,n+1}-p)+1
=((n+1)Qm,0+Qm1,0q)+((n+1)Pm,0+Pm1,0p)+1\displaystyle=((n+1)Q_{m,0}+Q_{m-1,0}-q)+((n+1)P_{m,0}+P_{m-1,0}-p)+1
=(n+1)(Qm,0+Pm,0)+((Qm1,0q)+(Pm1,0p)+1).\displaystyle=(n+1)(Q_{m,0}+P_{m,0})+\left((Q_{m-1,0}-q)+(P_{m-1,0}-p)+1\right).

Recalling the formula (4.2) for hm,0h_{m,0}, we get

hm,n+1\displaystyle h_{m,n+1} =(n+1)((Qm,0q)+(Pm,0p)+1+(p+q1))+hm1,0\displaystyle=(n+1)((Q_{m,0}-q)+(P_{m,0}-p)+1+(p+q-1))+h_{m-1,0}
=(n+1)hm,0+hm1,0+(n+1)(p+q1).\displaystyle=(n+1)h_{m,0}+h_{m-1,0}+(n+1)(p+q-1).

Recall that (m0,n0)(m_{0},n_{0}) is the largest pair in 𝒩λ<\mathcal{N}_{\lambda}^{<} such that either Qm0,n0<qQ_{m_{0},n_{0}}<q or Pm0,n0<pP_{m_{0},n_{0}}<p. Thus, if wλ(m,n)>wλ(m0,n0)w_{\lambda}(m,n)>w_{\lambda}(m_{0},n_{0}), then necessarily Qm,nqQ_{m,n}\geq q and Pm,npP_{m,n}\geq p, which further implies hm,n>0h_{m,n}>0 by (4.2). ∎

We won’t attempt to find recursive relations for all iterates of FκF_{\kappa} at 0, but we will at least calculate the orbit of 0 at iterates given by the sequence (hm,n)(m,n)𝒩λ<(h_{m,n})_{(m,n)\in\mathcal{N}_{\lambda}^{<}}.

Lemma 4.2.

Let (m,n)𝒩λ<(m,n)\in\mathcal{N}_{\lambda}^{<} such that wλ(m,n)>wλ(m0,n0)w_{\lambda}(m,n)>w_{\lambda}(m_{0},n_{0}). Then

Fκhm,n(0)=Δm,n.F_{\kappa}^{h_{m,n}}(0)=\Delta_{m,n}.
Proof.

Suppose, for a contradiction, that Fκhm,n(0)Δm,nF_{\kappa}^{h_{m,n}}(0)\neq\Delta_{m,n}. Let (at)t(a_{t})_{t\in\mathbb{N}} and (bt)n(b_{t})_{n\in\mathbb{N}} denote the sequences defined by

(4.3) Fκt(0)=η+btλat.F_{\kappa}^{t}(0)=-\eta+b_{t}\lambda-a_{t}.

Note that since η=pqλ\eta=p-q\lambda, we have

Fκt(0)=qλp+btλat=(bt+q)λ(at+p),F_{\kappa}^{t}(0)=q\lambda-p+b_{t}\lambda-a_{t}=(b_{t}+q)\lambda-(a_{t}+p),

which is never equal to 0 since λ\lambda is irrational and bt+qb_{t}+q and at+pa_{t}+p are both positive. Observe that (at)t(a_{t})_{t} and (bt)t(b_{t})_{t} are both non-decreasing and obey the following rule:

(at+1,bt+1)={(at,bt+1), if Fκt(0)<0,(at+1,bt), if Fκt(0)>0.(a_{t+1},b_{t+1})=\begin{cases}(a_{t},b_{t}+1),&\text{ if }F_{\kappa}^{t}(0)<0,\\ (a_{t}+1,b_{t}),&\text{ if }F_{\kappa}^{t}(0)>0.\end{cases}

Given that a0=a1=b0=b1=0a_{0}=a_{1}=b_{0}=b_{1}=0, we can deduce that the sequences (at)t(a_{t})_{t} and (bt)t(b_{t})_{t} achieve every non-negative integer value, that is, for any NN\in\mathbb{N}, there is some tNt\in N such that at=Na_{t}=N, and similarly there is some tt^{\prime}\in\mathbb{N} such that bt=Nb_{t^{\prime}}=N. Moreover, a simple inductive argument shows that for all integers t1t\geq 1,

at+bt+1=t.a_{t}+b_{t}+1=t.

Next, observe that Fκhm,n(0)Δm,nF_{\kappa}^{h_{m,n}}(0)\neq\Delta_{m,n} is equivalent to the statement that ahm,nPm,na_{h_{m,n}}\neq P_{m,n} and bhm,nQm,nb_{h_{m,n}}\neq Q_{m,n}. However, notice that

(Qm,nq)+(Pm,np)+1=hm,n=ahm,n+bhm,n+1.(Q_{m,n}-q)+(P_{m,n}-p)+1=h_{m,n}=a_{h_{m,n}}+b_{h_{m,n}}+1.

This implies one of two cases.

  1. (1)

    ahm,n<Pm,npa_{h_{m,n}}<P_{m,n}-p and bhm,n>Qm,nqb_{h_{m,n}}>Q_{m,n}-q; or

  2. (2)

    ahm,n>Pm,npa_{h_{m,n}}>P_{m,n}-p and bhm,n<Qm,nqb_{h_{m,n}}<Q_{m,n}-q.

Suppose case (1) holds. Since (bt)t(b_{t})_{t} is non-decreasing and takes every non-negative integer value, we know that there is some non-negative integer t<hm,nt^{*}<h_{m,n} such that bt=Qm,nqb_{t^{*}}=Q_{m,n}-q. Thus,

Fκt(0)=η+(Qm,nq)λat.F_{\kappa}^{t^{*}}(0)=-\eta+(Q_{m,n}-q)\lambda-a_{t^{*}}.

Note that at<at<Pm,npa_{t^{*}}<a_{t}<P_{m,n}-p. Now, suppose Δm,n>0\Delta_{m,n}>0. Then for any 0jPm,npat0\leq j\leq P_{m,n}-p-a_{t^{*}}, we have

Fκt(0)j\displaystyle F_{\kappa}^{t^{*}}(0)-j =η+(Qm,nq)λ(at+j)\displaystyle=-\eta+(Q_{m,n}-q)\lambda-(a_{t^{*}}+j)
>(Qm,nq)λ(Pm,np)\displaystyle>(Q_{m,n}-q)\lambda-(P_{m,n}-p)
=Δm,n\displaystyle=\Delta_{m,n}
>0.\displaystyle>0.

Thus, Fκt+j(0)=Fκt(0)jF_{\kappa}^{t^{*}+j}(0)=F_{\kappa}^{t^{*}}(0)-j for 0jPm,npat0\leq j\leq P_{m,n}-p-a_{t^{*}}. In particular,

Fκt+Pm,npat(0)=η+(Qm,nq)λ(Pm,np).F_{\kappa}^{t^{*}+P_{m,n}-p-a_{t^{*}}}(0)=-\eta+(Q_{m,n}-q)\lambda-(P_{m,n}-p).

But then

t+Pm,npat\displaystyle t^{*}+P_{m,n}-p-a_{t^{*}} =(Qm,nq)+(Pm,np)+1\displaystyle=(Q_{m,n}-q)+(P_{m,n}-p)+1
=hm,n.\displaystyle=h_{m,n}.

And this implies that

Fκhm,n(0)\displaystyle F_{\kappa}^{h_{m,n}}(0) =Fκt+Pm,npat(0)\displaystyle=F_{\kappa}^{t^{*}+P_{m,n}-p-a_{t^{*}}}(0)
=η+(Qm,nq)λ(Pm,np)\displaystyle=-\eta+(Q_{m,n}-q)\lambda-(P_{m,n}-p)
=Δm,n.\displaystyle=\Delta_{m,n}.

But this contradicts our assumption that Fκhm,n(0)Δm,nF_{\kappa}^{h_{m,n}}(0)\neq\Delta_{m,n}.

Now suppose that Δm,n<0\Delta_{m,n}<0. Let 0j<Pm,npat0\leq j<P_{m,n}-p-a_{t^{*}}. Then either Fκt(0)j>0F_{\kappa}^{t^{*}}(0)-j>0 or Δm,n<Fκt(0)j<0\Delta_{m,n}<F_{\kappa}^{t^{*}}(0)-j<0. But Δm,n<Fκt(0)j<0\Delta_{m,n}<F_{\kappa}^{t^{*}}(0)-j<0 implies

|Qm,nλ(at+j+p)|<|Δm,n|=|Qm,nλPm,n|,|Q_{m,n}\lambda-(a_{t^{*}}+j+p)|<|\Delta_{m,n}|=|Q_{m,n}\lambda-P_{m,n}|,

and at+j+p<Pm,na_{t^{*}}+j+p<P_{m,n}. This contradicts the ’best approximate’ property of the semiconvergent Pm,n/Qm,nP_{m,n}/Q_{m,n}. Thus Fκt(0)j>0F_{\kappa}^{t^{*}}(0)-j>0 for all 0j<Pm,npat0\leq j<P_{m,n}-p-a_{t^{*}}. Thus, by a similar argument to before, we reach the contradiction that Fκhm,n(0)=Δm,nF_{\kappa}^{h_{m,n}}(0)=\Delta_{m,n}.

In case (2), ahm,n>Pm,npa_{h_{m,n}}>P_{m,n}-p and bhm,n<Qm,nqb_{h_{m,n}}<Q_{m,n}-q. We can reach a similar contradiction as above, by using a similar argument where the roles of (at)t(a_{t})_{t} and (bt)t(b_{t})_{t} are interchanged.

This exhausts all cases, so our assumption that Fκhm,n(0)Δm,nF_{\kappa}^{h_{m,n}}(0)\neq\Delta_{m,n} must be false. ∎

Lemma 4.3.

Let mm\in\mathbb{N} such that wλ(m,0)>wλ(m0,n0)w_{\lambda}(m,0)>w_{\lambda}(m_{0},n_{0}). Then, for all 1t<hm+1,01\leq t<h_{m+1,0},

|Fκt(0)||Δm,0|.|F_{\kappa}^{t}(0)|\geq|\Delta_{m,0}|.
Proof.

recall that Fκt(0)=η+btλatF_{\kappa}^{t}(0)=-\eta+b_{t}\lambda-a_{t}, for some at,bta_{t},b_{t}\in\mathbb{N}. Since (at)t(a_{t})_{t} and (bt)t(b_{t})_{t} are non-decreasing and t<hm+1,0t<h_{m+1,0}, we have that btqm+1qb_{t}\leq q_{m+1}-q and atpm+1pa_{t}\leq p_{m+1}-p, not both equal. Hence either bt+q<qm+1b_{t}+q<q_{m+1} or at+p<pm+1a_{t}+p<p_{m+1} and bt+qqm+1b_{t}+q\leq q_{m+1}. In either case, by the best approximate property of convergents

|Fκt(0)|=|(bt+q)λ(at+p)||qmλpm|=|Δm,0|.|F_{\kappa}^{t}(0)|=|(b_{t}+q)\lambda-(a_{t}+p)|\geq|q_{m}\lambda-p_{m}|=|\Delta_{m,0}|.

Lemma 4.4.

Let (m,n)𝒩λ<(m,n)\in\mathcal{N}_{\lambda}^{<}. Then for all 1t<hm,n+11\leq t<h_{m,n+1},

Fκt(0)\displaystyle F_{\kappa}^{t}(0) Δm,0 or Fκt(0)nΔm,0+Δm1,0 if m is even,\displaystyle\geq\Delta_{m,0}\text{ or }F_{\kappa}^{t}(0)\leq n\Delta_{m,0}+\Delta_{m-1,0}\text{ if $m$ is even,}
Fκt(0)\displaystyle F_{\kappa}^{t}(0) Δm,0 or Fκt(0)nΔm,0+Δm1,0 if m is odd.\displaystyle\leq\Delta_{m,0}\text{ or }F_{\kappa}^{t}(0)\geq n\Delta_{m,0}+\Delta_{m-1,0}\text{ if $m$ is odd.}
Proof.

From (3.5), recall that if mm is even and n>0n>0, then Pm,n/Qm,n>λP_{m,n}/Q_{m,n}>\lambda is a best approximate from above, which implies

(4.4) bλaQm,nλPm,n<0,b\lambda-a\leq Q_{m,n}\lambda-P_{m,n}<0,

for all a,ba,b\in\mathbb{Z} with 0<b<Qm,n+10<b<Q_{m,n+1} such that Pm,n/Qm,na/b>λP_{m,n}/Q_{m,n}\neq a/b>\lambda.

Let (at)t(a_{t})_{t\in\mathbb{N}} and (bt)t(b_{t})_{t\in\mathbb{N}} be the sequences described by (4.3). Suppose that 1thm,n1\leq t\leq h_{m,n} with Fκt(0)<0F_{\kappa}^{t}(0)<0. Then

btQm,n+1q and atPm,n+1p,b_{t}\leq Q_{m,n+1}-q\text{ and }a_{t}\leq P_{m,n+1}-p,

since (at)t(a_{t})_{t} and (bt)t(b_{t})_{t} are non-decreasing. Thus,

(4.5) bt+qQm,n+1 and at+pPm,n+1,b_{t}+q\leq Q_{m,n+1}\text{ and }a_{t}+p\leq P_{m,n+1},

not both equal. Therefore, from (4.3) we know that

Fκt(0)=(bt+q)λ(at+p),F_{\kappa}^{t}(0)=(b_{t}+q)\lambda-(a_{t}+p),

and by (4.4) with (4.5), we have

Fκt(0){Qm,nλPm,n, if n0,qm1λpm1, if n=0.F_{\kappa}^{t}(0)\leq\begin{cases}Q_{m,n}\lambda-P_{m,n},&\text{ if }n\neq 0,\\ q_{m-1}\lambda-p_{m-1},&\text{ if }n=0.\end{cases}

Recalling the definition of Δm,n\Delta_{m,n} as in (3.7), we get

Fκt(0){Δm,n, if n0,Δm1,0, if n=0.F_{\kappa}^{t}(0)\leq\begin{cases}\Delta_{m,n},&\text{ if }n\neq 0,\\ \Delta_{m-1,0},&\text{ if }n=0.\end{cases}

Finally, by using (3.9), we have

Fκt(0)nΔm,0+Δm1,0.F_{\kappa}^{t}(0)\leq n\Delta_{m,0}+\Delta_{m-1,0}.

If Fκt(0)>0F_{\kappa}^{t}(0)>0, then by Lemma 4.3, we have

Fκt(0)Δm,0.F_{\kappa}^{t}(0)\geq\Delta_{m,0}.

From (3.6), recall that if mm is odd and n>0n>0, then Pm,n/Qm,n<λP_{m,n}/Q_{m,n}<\lambda is a best approximate from below, that is

bλaQm,nλPm,n>0,b\lambda-a\geq Q_{m,n}\lambda-P_{m,n}>0,

for all a,ba,b\in\mathbb{Z} with 0<b<Qm,n+10<b<Q_{m,n+1} such that Pm,n/Qm,na/b<λP_{m,n}/Q_{m,n}\neq a/b<\lambda. Thus, in the case that mm is odd and Fκt(0)>0F_{\kappa}^{t}(0)>0, then

bt+qQm,n+1 and at+pPm,n+1,b_{t}+q\leq Q_{m,n+1}\text{ and }a_{t}+p\leq P_{m,n+1},

not both equal. Thus, similarly to the above case where mm is even, we have

Fκt(0)\displaystyle F_{\kappa}^{t}(0) =(bt+q)λ(at+p)\displaystyle=(b_{t}+q)\lambda-(a_{t}+p)
{Qm,nλPm,n if n0qm1λpm1 if n=0\displaystyle\geq\begin{cases}Q_{m,n}\lambda-P_{m,n}&\text{ if }n\neq 0\\ q_{m-1}\lambda-p_{m-1}&\text{ if }n=0\end{cases}
=nΔm,0+Δm1,0.\displaystyle=n\Delta_{m,0}+\Delta_{m-1,0}.

On the other hand, if Fκt(0)<0F_{\kappa}^{t}(0)<0, then by Lemma 4.3,

Fκt(0)Δm,0.F_{\kappa}^{t}(0)\leq\Delta_{m,0}.

In order to prove the next theorem, we will distinguish between the following two cases and we will prove them separately. We will first prove that

if zE1(Sm,n), then h(z)=hm,n+1,\text{if }z\in E^{-1}(S_{m,n}),\text{ then }h(z)=h_{m,n+1},

and then we will prove that

if h(z)=hm,n+1, then Fκh(z)(z)=E(z)+Δm,n+1.\text{if }h(z)=h_{m,n+1},\text{ then }F_{\kappa}^{h(z)}(z)=E(z)+\Delta_{m,n+1}.
Theorem 4.5.

Let α𝔹d+2\alpha\in\mathbb{B}^{d+2}, τ:{1,,d}{1,,d}\tau:\{1,...,d\}\rightarrow\{1,...,d\} be a bijection, λ[0,1)\lambda\in[0,1)\setminus\mathbb{Q}, λ<η=pqλ<1-\lambda<\eta=p-q\lambda<1 for some p,q{0}p,q\in\mathbb{N}\setminus\{0\} and set κ=(α,τ,λ,η,1)\kappa=(\alpha,\tau,\lambda,\eta,1). For all (m,n)𝒩λ<(m,n)\in\mathcal{N}_{\lambda}^{<} with wλ(m,n)>wλ(m0,n0)w_{\lambda}(m,n)>w_{\lambda}(m_{0},n_{0}), h(E1(Sm,n))h(E^{-1}(S_{m,n})) exists and is equal to hm,n+1h_{m,n+1}. Moreover, let zE1(Sm,n)z\in E^{-1}(S_{m,n}). Then

(4.6) Rκ(z)=E(z)+Δm,n+1(λ).R_{\kappa}(z)=E(z)+\Delta_{m,n+1}(\lambda).
Proof.

Let zCcz\in C_{c}. Assume zE1(Sm,n)z\in E^{-1}(S_{m,n}). Observe that E(z)+Fκhm,n+1(0)CcE(z)+F_{\kappa}^{h_{m,n+1}}(0)\in C_{c}, so h(z)hm,n+1h(z)\leq h_{m,n+1}. By Lemma 3.3, we know that

Fκh(z)(z)=E(z)+Fκh(z)(0).F_{\kappa}^{h(z)}(z)=E(z)+F_{\kappa}^{h(z)}(0).

Suppose, for a contradiction, that h(z)<hm,n+1h(z)<h_{m,n+1}. We will prove the contradiction for even and odd mm separately, starting with the case that mm is even. By Lemma 4.4, we know that either Fκh(z)(0)Δm,0F_{\kappa}^{h(z)}(0)\geq\Delta_{m,0} or Fκh(z)(0)nΔm,0+Δm1,0F_{\kappa}^{h(z)}(0)\leq n\Delta_{m,0}+\Delta_{m-1,0}. Since mm is even, recall that

Sm,n=(C0Δm,0)(CcΔm,n+1)Cc(Cd+1(nΔm,0+Δm1,0)).S_{m,n}=(C_{0}-\Delta_{m,0})\cap(C_{c}-\Delta_{m,n+1})\cap C_{c}\cap(C_{d+1}-(n\Delta_{m,0}+\Delta_{m-1,0})).

Observe that

Fκh(z)(z)\displaystyle F_{\kappa}^{h(z)}(z) Sm,n+Fκh(z)(0)\displaystyle\in S_{m,n}+F_{\kappa}^{h(z)}(0)
(C0+(Fκh(z)(0)Δm,0))(Cd+1+(Fκh(z)(0)(nΔm,0+Δm1,0))).\displaystyle\subset(C_{0}+(F_{\kappa}^{h(z)}(0)-\Delta_{m,0}))\cap(C_{d+1}+(F_{\kappa}^{h(z)}(0)-(n\Delta_{m,0}+\Delta_{m-1,0}))).

Therefore, if Fκh(z)(0)Δm,0F_{\kappa}^{h(z)}(0)\geq\Delta_{m,0}, then

Fκh(z)(z)C0+(Fκh(z)(0)Δm,0)C0.F_{\kappa}^{h(z)}(z)\in C_{0}+(F_{\kappa}^{h(z)}(0)-\Delta_{m,0})\subset C_{0}.

But by the definition of h(z)h(z) as in (2.6), we have

Fκh(z)(z)=Rκ(z)Cc,F_{\kappa}^{h(z)}(z)=R_{\kappa}(z)\in C_{c},

which reveals a contradiction. Similarly, if Fκh(z)(0)nΔm,0+Δm1,0F_{\kappa}^{h(z)}(0)\leq n\Delta_{m,0}+\Delta_{m-1,0}, then

Fκh(z)(z)Cd+1+(Fκh(z)(0)(nΔm,0+Δm1,0))Cd+1,F_{\kappa}^{h(z)}(z)\in C_{d+1}+(F_{\kappa}^{h(z)}(0)-(n\Delta_{m,0}+\Delta_{m-1,0}))\subset C_{d+1},

which also contradicts Fκh(z)(z)CcF_{\kappa}^{h(z)}(z)\in C_{c}.

Now suppose mm is odd. Then

Sm,n=(C0(nΔm,0+Δm1,0))Cc(CcΔm,n+1))(Cd+1Δm,0).S_{m,n}=(C_{0}-(n\Delta_{m,0}+\Delta_{m-1,0}))\cap C_{c}\cap(C_{c}-\Delta_{m,n+1}))\cap(C_{d+1}-\Delta_{m,0}).

By Lemma 4.4, we know that either

Fκh(z)(0)nΔm,0+Δm1,0 or Fκh(z)(0)Δm,n.F_{\kappa}^{h(z)}(0)\geq n\Delta_{m,0}+\Delta_{m-1,0}\text{ or }F_{\kappa}^{h(z)}(0)\leq\Delta_{m,n}.

Clearly, either of these cases give similar contradictions as before. Therefore our assumption that h(z)<hm,n+1h(z)<h_{m,n+1} must be false, so in fact h(z)=hm,n+1h(z)=h_{m,n+1}.

Now, suppose h(z)=hm,n+1h(z)=h_{m,n+1}. By Lemma 3.3,

Fκh(z)(z)=E(z)+Fκh(z)(0),F_{\kappa}^{h(z)}(z)=E(z)+F_{\kappa}^{h(z)}(0),

and by Lemma 4.2 we know that

Fκhm,n+1(0)=Δm,n+1.F_{\kappa}^{h_{m,n+1}}(0)=\Delta_{m,n+1}.

Combining these two with our assumption gives us that if h(z)=hm,n+1,h(z)=h_{m,n+1}, then Fκh(z)(z)=E(z)+Δm,n+1.F_{\kappa}^{h(z)}(z)=E(z)+\Delta_{m,n+1}.

With this Theorem, as well as Theorem 3.9, we can prove the existence of a renormalization scheme around the point 0. First, we will find the definition of the first return map RκR_{\kappa} on the rest of CcC_{c}.

Lemma 4.6.

Let Sm,nS_{m,n} be as in (3.15) and (m0,n0)𝒩λ<(m_{0},n_{0})\in\mathcal{N}_{\lambda}^{<} as in (4.1). Define the set U(κ)U(\kappa)

U(κ)={0}(m,n)𝒩λ<wλ(m,n)wλ(m0,n0)Sm,n.U(\kappa)=\{0\}\cup\bigcup_{\begin{subarray}{c}(m,n)\in\mathcal{N}_{\lambda}^{<}\\ w_{\lambda}(m,n)\geq w_{\lambda}(m_{0},n_{0})\end{subarray}}S_{m,n}.

Then

U(κ)={(C0Δm0,0)Cc(Cd+1(n0Δm0,0+Δm01,0)),if m0 is even,(C0(n0Δm0,0+Δm01,0))Cc(Cd+1Δm0,0),if m0 is odd,U(\kappa)=\begin{cases}(C_{0}-\Delta_{m_{0},0})\cap C_{c}\cap(C_{d+1}-(n_{0}\Delta_{m_{0},0}+\Delta_{m_{0}-1,0})),&\text{if }m_{0}\text{ is even,}\\[4.2679pt] (C_{0}-(n_{0}\Delta_{m_{0},0}+\Delta_{m_{0}-1,0}))\cap C_{c}\cap(C_{d+1}-\Delta_{m_{0},0}),&\text{if }m_{0}\text{ is odd,}\end{cases}

and U(κ)U(\kappa) is convex.

Proof.

For brevity we will drop the parameters κ\kappa when they are unambiguous. We will first prove the equality

U(κ)={(C0Δm0,0)Cc(Cd+1(n0Δm0,0+Δm01,0)),if m0 is even,(C0(n0Δm0,0+Δm01,0))Cc(Cd+1Δm0,0),if m0 is odd..U(\kappa)=\begin{cases}(C_{0}-\Delta_{m_{0},0})\cap C_{c}\cap(C_{d+1}-(n_{0}\Delta_{m_{0},0}+\Delta_{m_{0}-1,0})),&\text{if }m_{0}\text{ is even,}\\[4.2679pt] (C_{0}-(n_{0}\Delta_{m_{0},0}+\Delta_{m_{0}-1,0}))\cap C_{c}\cap(C_{d+1}-\Delta_{m_{0},0}),&\text{if }m_{0}\text{ is odd.}\end{cases}.

Observe that for all (m,n)𝒩λ<(m,n)\in\mathcal{N}_{\lambda}^{<},

Sm,nCc.S_{m,n}\subset C_{c}.

Additionally, if (m,n)𝒩λ<(m,n)\in\mathcal{N}_{\lambda}^{<} such that wλ(m,n)=j+k01w_{\lambda}(m,n)=j+k_{0}-1, then

Sm,n{m=m0(C0Δm,0), if m0 is even,wλ(m,n)wλ(m0,n0)(C0(nΔm,0+Δm1,0)), if m0 is odd.S_{m,n}\subset\begin{cases}\bigcup_{m=m_{0}}^{\infty}(C_{0}-\Delta_{m,0}),&\text{ if $m_{0}$ is even,}\\[5.69054pt] \bigcup_{w_{\lambda}(m,n)\geq w_{\lambda}(m_{0},n_{0})}(C_{0}-(n\Delta_{m,0}+\Delta_{m-1,0})),&\text{ if $m_{0}$ is odd.}\end{cases}

Note that C0(nΔm+1,0+Δm,0)C0Δm,0C_{0}-(n\Delta_{m+1,0}+\Delta_{m,0})\subset C_{0}-\Delta_{m,0} for all (m+1,n)𝒩λ<(m+1,n)\in\mathcal{N}_{\lambda}^{<}. Thus, we have

Sm,n{C0Δm0,0, if m0 is even,C0(n0Δm0,0+Δm01,0), if m0 is odd..S_{m,n}\subset\begin{cases}C_{0}-\Delta_{m_{0},0},&\text{ if $m_{0}$ is even,}\\[4.2679pt] C_{0}-(n_{0}\Delta_{m_{0},0}+\Delta_{m_{0}-1,0}),&\text{ if $m_{0}$ is odd.}\end{cases}.

We also know that

Sm,n{wλ(m,n)wλ(m0,n0)(Cd+1(nΔm,0+Δm1,0)), if m0 is even,m=m0+1(Cd+1Δm,0), if m0 is odd.S_{m,n}\subset\begin{cases}\bigcup_{w_{\lambda}(m,n)\geq w_{\lambda}(m_{0},n_{0})}(C_{d+1}-(n\Delta_{m,0}+\Delta_{m-1,0})),&\text{ if $m_{0}$ is even,}\\[5.69054pt] \bigcup_{m=m_{0}+1}^{\infty}(C_{d+1}-\Delta_{m,0}),&\text{ if $m_{0}$ is odd.}\end{cases}

Thus, with a similar argument as above, we can show that

Sm,n{Cd+1(n0Δm0,0+Δm01,0), if m0 is even,Cd+1Δm0,0, if m0 is odd.,S_{m,n}\subset\begin{cases}C_{d+1}-(n_{0}\Delta_{m_{0},0}+\Delta_{m_{0}-1,0}),&\text{ if $m_{0}$ is even,}\\[4.2679pt] C_{d+1}-\Delta_{m_{0},0},&\text{ if $m_{0}$ is odd.}\end{cases},

Altogether, we deduce that

U(κ){(C0Δm0,0)Cc(Cd+1(n0Δm0,0+Δm01,0)),if m0 is even,(C0(n0Δm0,0+Δm01,0))Cc(Cd+1Δm0,0),if m0 is odd..U(\kappa)\subset\begin{cases}(C_{0}-\Delta_{m_{0},0})\cap C_{c}\cap(C_{d+1}-(n_{0}\Delta_{m_{0},0}+\Delta_{m_{0}-1,0})),&\text{if }m_{0}\text{ is even,}\\[4.2679pt] (C_{0}-(n_{0}\Delta_{m_{0},0}+\Delta_{m_{0}-1,0}))\cap C_{c}\cap(C_{d+1}-\Delta_{m_{0},0}),&\text{if }m_{0}\text{ is odd.}\end{cases}.

Suppose m0m_{0} is even, and let

z(C0Δm0,0)Cc(Cd+1(n0Δm0,0+Δm01,0)),z\in(C_{0}-\Delta_{m_{0},0})\cap C_{c}\cap(C_{d+1}-(n_{0}\Delta_{m_{0},0}+\Delta_{m_{0}-1,0})),

with z0z\neq 0. Then there is some (m,n)𝒩λ<(m,n)\in\mathcal{N}_{\lambda}^{<} with mm is odd and wλ(m,n)wλ(m0,n0)w_{\lambda}(m,n)\geq w_{\lambda}(m_{0},n_{0}) such that

zC0(nΔm,0+Δm1,0),z\in C_{0}-(n\Delta_{m,0}+\Delta_{m-1,0}),

and there is some (m,n)𝒩λ<(m^{\prime},n^{\prime})\in\mathcal{N}_{\lambda}^{<} with mm^{\prime} even and wλ(m,n)wλ(m0,n0)w_{\lambda}(m^{\prime},n^{\prime})\geq w_{\lambda}(m_{0},n_{0}) such that

zCd+1(nΔm,0+Δm1,0).z\in C_{d+1}-(n\Delta_{m,0}+\Delta_{m-1,0}).

Suppose that (m,n)(m,n) and (m,n)(m^{\prime},n^{\prime}) are the largest such pairs, which is well-defined since z0z\neq 0. Since mm is odd and mm^{\prime} is even, we either have m<mm<m^{\prime} or m<mm^{\prime}<m. Suppose m<mm<m^{\prime}. Then

z(C0(nΔm,0+Δm1,0))Cc(Cd+1Δm,0).z\in(C_{0}-(n\Delta_{m,0}+\Delta_{m-1,0}))\cap C_{c}\cap(C_{d+1}-\Delta_{m,0}).

Since (m,n)(m,n) is the largest pair such that zC0(nΔm,0+Δm1,0)z\in C_{0}-(n\Delta_{m,0}+\Delta_{m-1,0}), and noting that (n+1)Δm,0+Δm1,0=Δm,n+1(n+1)\Delta_{m,0}+\Delta_{m-1,0}=\Delta_{m,n+1} since n+1>0n+1>0, we have that

z(CcΔm,n+1) or z(Cd+1Δm,n+1).z\in(C_{c}-\Delta_{m,n+1})\text{ or }z\in(C_{d+1}-\Delta_{m,n+1}).

However, Cd+1Δm,n+1Cd+1C_{d+1}-\Delta_{m,n+1}\subset C_{d+1} since mm is odd and so Δm,n+1>0\Delta_{m,n+1}>0. Importantly,

Cc(Cd+1Δm,n+1)=,C_{c}\cap(C_{d+1}-\Delta_{m,n+1})=\emptyset,

and thus zCcΔm,n+1z\in C_{c}-\Delta_{m,n+1}. Therefore,

z(C0(nΔm,0+Δm1,0))(CcΔm,n+1)Cc(Cd+1Δm,0)=Sm,n,z\in(C_{0}-(n\Delta_{m,0}+\Delta_{m-1,0}))\cap(C_{c}-\Delta_{m,n+1})\cap C_{c}\cap(C_{d+1}-\Delta_{m,0})=S_{m,n},

so clearly zU(κ)z\in U(\kappa). In the case that m<mm^{\prime}<m we can use a similar argument to prove that

z\displaystyle z (C0Δm,0)Cc(CcΔm,n+1)(Cd+1(nΔm,0+Δm1,0))\displaystyle\in(C_{0}-\Delta_{m^{\prime},0})\cap C_{c}\cap(C_{c}-\Delta_{m^{\prime},n^{\prime}+1})\cap(C_{d+1}-(n^{\prime}\Delta_{m^{\prime},0}+\Delta_{m^{\prime}-1,0}))
=Sm,n,\displaystyle=S_{m^{\prime},n^{\prime}},

so that zU(κ)z\in U(\kappa). If m0m_{0} is odd and

z(C0(n0Δm0,0+Δm01,0))Cc(Cd+1Δm0,0),z\in(C_{0}-(n_{0}\Delta_{m_{0},0}+\Delta_{m_{0}-1,0}))\cap C_{c}\cap(C_{d+1}-\Delta_{m_{0},0}),

with z0z\neq 0, then we can use a similar argument to prove that zU(κ)z\in U(\kappa). Hence, we have

U(κ)={(C0Δm0,0)Cc(Cd+1(n0Δm0,0+Δm01,0))if m0 is even(C0(n0Δm0,0+Δm01,0))Cc(Cd+1Δm0,0)if m0 is odd.U(\kappa)=\begin{cases}(C_{0}-\Delta_{m_{0},0})\cap C_{c}\cap(C_{d+1}-(n_{0}\Delta_{m_{0},0}+\Delta_{m_{0}-1,0}))&\text{if }m_{0}\text{ is even}\\[4.2679pt] (C_{0}-(n_{0}\Delta_{m_{0},0}+\Delta_{m_{0}-1,0}))\cap C_{c}\cap(C_{d+1}-\Delta_{m_{0},0})&\text{if }m_{0}\text{ is odd}\end{cases}.

To show that UU is convex, one must note that the cones C0C_{0}, CcC_{c}, and Cd+1C_{d+1} and all their translates are convex sets, and that the intersection of convex sets is also convex. ∎

Define

(4.7) Uk,l(κ)={0}(m,n)𝒩λ<wλ(m,n)wλ(k,l)Sm,n,U_{k,l}(\kappa)=\{0\}\cup\bigcup_{\begin{subarray}{c}(m,n)\in\mathcal{N}_{\lambda}^{<}\\ w_{\lambda}(m,n)\geq w_{\lambda}(k,l)\end{subarray}}S_{m,n},

for wλ(k,l)wλ(m0,n0)w_{\lambda}(k,l)\geq w_{\lambda}(m_{0},n_{0}) (omitting the arguments where unambiguous), and let

κ=(α,τ,λ,η,ρ),\kappa^{\prime}=(\alpha,\tau,\lambda,\eta^{\prime},\rho),

where λ<η=pqλ<ρ-\lambda<\eta^{\prime}=p^{\prime}-q^{\prime}\lambda<\rho is such that ppp^{\prime}\leq p and qqq^{\prime}\leq q. If (m0,n0)𝒩λ<(m_{0}^{\prime},n_{0}^{\prime})\in\mathcal{N}_{\lambda}^{<} is the maximal element of 𝒩λ<\mathcal{N}_{\lambda}^{<} such that Pm0,n0<pP_{m_{0}^{\prime},n_{0}^{\prime}}<p^{\prime} or Qm0,n0<qQ_{m^{\prime}_{0},n^{\prime}_{0}}<q^{\prime}, then wλ(m0,n0)wλ(m0,n0)w_{\lambda}(m_{0}^{\prime},n_{0}^{\prime})\leq w_{\lambda}(m_{0},n_{0}). Furthermore,

(4.8) Rκ|Um0,n0(κ)(z)=Rκ|Um0,n0(κ)(z).R_{\kappa}|_{U_{m_{0},n_{0}}(\kappa)}(z)=R_{\kappa^{\prime}}|_{U_{m_{0},n_{0}}(\kappa)}(z).

Given α𝔹\alpha\in\mathbb{B}, τ:{1,,d}{1,,d}\tau:\{1,...,d\}\rightarrow\{1,...,d\} be a bijection, λ[0,1)\lambda\in[0,1)\setminus\mathbb{Q}, ρ=1\rho=1, λ<η=pqλ<1-\lambda<\eta=p-q\lambda<1 for some p,qp,q\in\mathbb{N}, let (m0,n0)𝒩λ<(m_{0},n_{0})\in\mathcal{N}_{\lambda}^{<} be as in (4.1). Let p,q{0}p^{\prime},q^{\prime}\in\mathbb{N}\setminus\{0\} be defined by

p=Pm0,0(g2(λ)), and q=Qm0,0(g2(λ)),p^{\prime}=P_{m_{0},0}(g^{2}(\lambda)),\text{ and }q^{\prime}=Q_{m_{0},0}(g^{2}(\lambda)),

and let

(4.9) η=pqg2(λ).\eta^{\prime}=p^{\prime}-q^{\prime}g^{2}(\lambda).

Clearly by the definition of the one-sided convergents in (3.4), we have

g2(λ)<η<1.-g^{2}(\lambda)<\eta<1.

With this in mind, we have the following Theorem.

Theorem 4.7.

Let α𝔹d+2\alpha\in\mathbb{B}^{d+2}, τ:{1,,d}{1,,d}\tau:\{1,...,d\}\rightarrow\{1,...,d\} be a bijection, λ[0,1)\lambda\in[0,1)\setminus\mathbb{Q} and λ<η=pqλ<1-\lambda<\eta=p-q\lambda<1 for some p,qp,q\in\mathbb{N}. Set the tuples κ=(α,τ,λ,η,1)\kappa=(\alpha,\tau,\lambda,\eta,1) and κ=(α,τ,λ,η,1)\kappa^{\prime}=(\alpha,\tau,\lambda^{\prime},\eta^{\prime},1), where λ=g2(λ)\lambda^{\prime}=g^{2}(\lambda) and η\eta^{\prime} is as in (4.9). Then for all zUm0,0(κ)z\in U_{m_{0},0}(\kappa^{\prime}),

Rκ(z)=11λ1λRκ((1λ1λ)z).R_{\kappa^{\prime}}(z)=\frac{1}{1-\lambda_{1}\lambda}R_{\kappa}((1-\lambda_{1}\lambda)z).
Proof.

We begin by noting that the quantity

(4.10) (m0,n0)=max{(m,n)𝒩g2(λ)<:Pm,n(g2(λ))<p or Qm,n(g2(λ))<q},(m_{0}^{\prime},n_{0}^{\prime})=\max\{(m,n)\in\mathcal{N}_{g^{2}(\lambda)}^{<}:P_{m,n}(g^{2}(\lambda))<p^{\prime}\text{ or }Q_{m,n}(g^{2}(\lambda))<q^{\prime}\},

satisfies (m0,n0)=(m0,0)(m_{0}^{\prime},n_{0}^{\prime})=(m_{0},0) and, importantly for the reasons of (4.8), we have

(4.11) wλ(m0+2,n0)wλ(m0+2,0).w_{\lambda}(m_{0}^{\prime}+2,n_{0}^{\prime})\leq w_{\lambda}(m_{0}+2,0).

Here we recall the equality in Proposition 3.4 since (m0,n0)(m_{0}^{\prime},n_{0}^{\prime}) is being considered an element of 𝒩g2(λ)<\mathcal{N}_{g^{2}(\lambda)}^{<} and (m0,0)(m_{0},0) an element of 𝒩λ<\mathcal{N}_{\lambda}^{<}.

Recall that by Theorem 3.9, we have

11λ1λSj+2,n(λ)=Sj,n(g2(λ)),\frac{1}{1-\lambda_{1}\lambda}S_{j+2,n}(\lambda)=S_{j,n}(g^{2}(\lambda)),

for all (j+2,n)𝒩λ<(j+2,n)\in\mathcal{N}_{\lambda}^{<} with wλ(j+2,n)wλ(m0+2,0)w_{\lambda}(j+2,n)\geq w_{\lambda}(m_{0}+2,0), i.e. jm0j\geq m_{0}. Note that by Proposition 3.4,

wλ(m0+2,n0)wλ(m0+2,0),w_{\lambda}(m_{0}^{\prime}+2,n_{0}^{\prime})\leq w_{\lambda}(m_{0}+2,0),

is equivalent to the statement that

wg2(λ)(m0,n0)wg2(λ)(m0,0).w_{g^{2}(\lambda)}(m_{0}^{\prime},n_{0}^{\prime})\leq w_{g^{2}(\lambda)}(m_{0},0).

Hence,

11λ1λUm0+2,0(κ)=Um0,0(κ).\frac{1}{1-\lambda_{1}\lambda}U_{m_{0}+2,0}(\kappa)=U_{m_{0},0}(\kappa^{\prime}).

Let zCcz\in C_{c} so that (1λ1λ)zSm+2,n(λ)(1-\lambda_{1}\lambda)z\in S_{m+2,n}(\lambda) with wλ(m+2,n)wλ(m0+2,0)w_{\lambda}(m+2,n)\geq w_{\lambda}(m_{0}+2,0). Also note that since aCj=CjaC_{j}=C_{j} for all a>0a>0 and EE consists of rotations about 0, we have

(4.12) E(az)=aE(z),E(az)=aE(z),

for a>0a>0. Therefore zSm,n(g2(λ))z\in S_{m,n}(g^{2}(\lambda)) and by expanding RκR_{\kappa} as in (4.6) we get

11λ1λRκ((1λ1λ)z)=11λ1λ(E((1λ1λ)z)+Δm+2,n+1(λ)).\frac{1}{1-\lambda_{1}\lambda}R_{\kappa}\left((1-\lambda_{1}\lambda)z\right)=\frac{1}{1-\lambda_{1}\lambda}\left(E((1-\lambda_{1}\lambda)z)+\Delta_{m+2,n+1}(\lambda)\right).

Using 4.12, we get

11λ1λRκ((1λ1λ)z)\displaystyle\frac{1}{1-\lambda_{1}\lambda}R_{\kappa}\left((1-\lambda_{1}\lambda)z\right) =11λ1λ((1λ1λ)E(z)+Δm+2,n+1(λ))\displaystyle=\frac{1}{1-\lambda_{1}\lambda}\left((1-\lambda_{1}\lambda)E(z)+\Delta_{m+2,n+1}(\lambda)\right)
=E(z)+(Δm+2,n+1(λ)λ1λ1).\displaystyle=E(z)+\left(-\frac{\Delta_{m+2,n+1}(\lambda)}{\lambda_{1}\lambda-1}\right).

Now, recall that p1=1p_{1}=1 and q1=λ1q_{1}=\lambda_{1}, so that Δ1,0(λ)=λ1λ1\Delta_{1,0}(\lambda)=\lambda_{1}\lambda-1. Using (3.14) and comparing with the formula for RκR_{\kappa^{\prime}} as in (4.6), we see

11λ1λRκ((1λ1λ)z)\displaystyle\frac{1}{1-\lambda_{1}\lambda}R_{\kappa}\left((1-\lambda_{1}\lambda)z\right) =E(z)+(Δm+2,n+1(λ)Δ1,0(λ))\displaystyle=E(z)+\left(-\frac{\Delta_{m+2,n+1}(\lambda)}{\Delta_{1,0}(\lambda)}\right)
=E(z)+Δm,n+1(g2(λ))\displaystyle=E(z)+\Delta_{m,n+1}(g^{2}(\lambda))
=Rκ(z).\displaystyle=R_{\kappa^{\prime}}(z).

The immediate consequence of Theorem 4.7 is that when λ[0,1)\lambda\in[0,1)\setminus\mathbb{Q} and λ<η=pqλ<1-\lambda<\eta=p-q\lambda<1, one can renormalize infinitely “towards” 0 in the sense that the domains of each renormalization are shrinking neighbourhoods of 0 in CcC_{c}. Additionally, the renormalizations are determined by the orbit of λ\lambda under the square of the Gauss map g2(x)g^{2}(x). We shall now apply our results in this and the previous section to an example.

5. An Example

As an example inspired by [28], we will set α𝔹d+2\alpha\in\mathbb{B}^{d+2}, λ=Φ\lambda=\Phi, ρ=1\rho=1, and η=Φ2\eta=\Phi^{2}, where

(5.1) Φ=512.\Phi=\frac{\sqrt{5}-1}{2}.

In this case the the space 𝒩λ<=×{0}\mathcal{N}_{\lambda}^{<}=\mathbb{N}\times\{0\}\cong\mathbb{N} since λ=[0;1¯]\lambda=[0;\bar{1}] and so λk=1\lambda_{k}=1 for all kk\in\mathbb{N}. Thus, wΦ(m,n)=wΦ(m)=mw_{\Phi}(m,n)=w_{\Phi}(m)=m. Also due to λk=1\lambda_{k}=1 for each kk\in\mathbb{N}, the semiconvergents Pm/QmP_{m}/Q_{m} simply coincide with the convergents pm/qmp_{m}/q_{m}, and the convergents are in this case defined by

p0\displaystyle p_{0} =0,\displaystyle=0, q0=1,\displaystyle q_{0}=1,
p1\displaystyle p_{1} =1,\displaystyle=1, q1=1,\displaystyle q_{1}=1,
pm\displaystyle p_{m} =pn1+pn2,\displaystyle=p_{n-1}+p_{n-2}, qm=qm1+qm2.\displaystyle q_{m}=q_{m-1}+q_{m-2}.

It is thus clear that pm=Fibmp_{m}=\operatorname{Fib}_{m} and qm=Fibm+1q_{m}=\operatorname{Fib}_{m+1}, where (Fibm)m(\operatorname{Fib}_{m})_{m\in\mathbb{N}} is the Fibonacci sequence with Fib0=0\operatorname{Fib}_{0}=0 and Fib1=1\operatorname{Fib}_{1}=1.

The first return times (hm)m(h_{m})_{m\in\mathbb{N}} in the case of λ=Φ\lambda=\Phi are given by

hm=(qm1)+(pm1)+1=Fibm+1+Fibm1=Fibm+21h_{m}=(q_{m}-1)+(p_{m}-1)+1=\operatorname{Fib}_{m+1}+\operatorname{Fib}_{m}-1=\operatorname{Fib}_{m+2}-1

for all mm\in\mathbb{N}. Observe that hm=hm1+hm2+1h_{m}=h_{m-1}+h_{m-2}+1 for all m2m\geq 2, and h0=2h_{0}=2, h1=4h_{1}=4. Now note that η=Φ2=1Φ=1λ\eta=\Phi^{2}=1-\Phi=1-\lambda, and thus we have

m0=max{m:pm<1 or qm<1}=0.\displaystyle m_{0}=\max\{m\in\mathbb{N}:p_{m}<1\text{ or }q_{m}<1\}=0.

The errors of the convergents Δm(Φ)\Delta_{m}(\Phi) are given by

(5.2) Δm(Φ)=qmλpm=Fibm+1ΦFibm.\Delta_{m}(\Phi)=q_{m}\lambda-p_{m}=\operatorname{Fib}_{m+1}\Phi-\operatorname{Fib}_{m}.
Refer to caption
Figure 9. A plot of 1500 iterates of 500 uniformly chosen points within the box [1,λ]×[0,1.1][-1,\lambda]\times[0,1.1] under the TCE with parameters α=(π/20.6,0.5,0.7,π0.6)\alpha=(\pi/2-0.6,0.5,0.7,\pi-0.6), τ:12,21\tau:1\mapsto 2,2\mapsto 1, λ=Φ\lambda=\Phi, η=Φ2\eta=\Phi^{2}, and ρ=1\rho=1. The first 400 points of each orbit are omitted to remove transients.
Proposition 5.1.

We have

Δm(Φ)=(Φ)m+1\Delta_{m}(\Phi)=-(-\Phi)^{m+1}
Proof.

Observe that

Δm(Φ)\displaystyle\Delta_{m}(\Phi) =Fibm+1ΦFibm\displaystyle=\operatorname{Fib}_{m+1}\Phi-\operatorname{Fib}_{m}
=FibmΦ+Fibm1ΦFibm\displaystyle=\operatorname{Fib}_{m}\Phi+\operatorname{Fib}_{m-1}\Phi-\operatorname{Fib}_{m}
=Fibm1Φ(1Φ)Fibm\displaystyle=\operatorname{Fib}_{m-1}\Phi-(1-\Phi)\operatorname{Fib}_{m}
=Φ(FibmΦFibm1)\displaystyle=-\Phi(\operatorname{Fib}_{m}\Phi-\operatorname{Fib}_{m-1})
=ΦΔm1(Φ).\displaystyle=-\Phi\Delta_{m-1}(\Phi).

Recall that p0=0p_{0}=0 and q0=1q_{0}=1. Then a simple inductive argument shows us that

Δm(Φ)=(Φ)mΔ0(Φ)=(Φ)m(q0Φp0)=(Φ)m+1\Delta_{m}(\Phi)=(-\Phi)^{m}\Delta_{0}(\Phi)=(-\Phi)^{m}(q_{0}\Phi-p_{0})=-(-\Phi)^{m+1}

Refer to caption
Figure 10. A partition of CcC_{c} in the case where α=(1,0.5,π2.5,1)\alpha=(1,0.5,\pi-2.5,1), τ:12,21\tau:1\mapsto 2,2\mapsto 1, λ=Φ\lambda=\Phi, η=Φ2\eta=\Phi^{2}, and ρ=1\rho=1. A cascading pattern towards the origin can be seen, but its geometric structure becomes clearer after we apply the cone exchange EE.

Noting that the recurrence relations for pmp_{m} and qmq_{m} give us p1=1p_{-1}=1 and q1=0q_{-1}=0 and so we can set Δ1=1=(Φ)0\Delta_{-1}=-1=-(-\Phi)^{0}, which remains consistent with (5.2). With this proposition in mind, for mm\in\mathbb{N} we can determine the sets SmS_{m} as

Sm\displaystyle S_{m} ={(C0Δm)Cc(CcΔm+1)(Cd+1Δm1), if m is even,(C0Δm1)(CcΔm+1)Cc(Cd+1Δm), if m is odd.\displaystyle=\begin{cases}(C_{0}-\Delta_{m})\cap C_{c}\cap(C_{c}-\Delta_{m+1})\cap(C_{d+1}-\Delta_{m-1}),&\text{ if $m$ is even,}\\[4.2679pt] (C_{0}-\Delta_{m-1})\cap(C_{c}-\Delta_{m+1})\cap C_{c}\cap(C_{d+1}-\Delta_{m}),&\text{ if $m$ is odd.}\end{cases}
={(C0+(Φ)m+1)Cc(Cc+(Φ)m+2)(Cd+1+(Φ)m), if m is even,(C0+(Φ)m)(Cc+(Φ)m+2)Cc(Cd+1+(Φ)m+1), if m is odd.\displaystyle=\begin{cases}(C_{0}+(-\Phi)^{m+1})\cap C_{c}\cap(C_{c}+(-\Phi)^{m+2})\cap(C_{d+1}+(-\Phi)^{m}),&\text{ if $m$ is even,}\\[4.2679pt] (C_{0}+(-\Phi)^{m})\cap(C_{c}+(-\Phi)^{m+2})\cap C_{c}\cap(C_{d+1}+(-\Phi)^{m+1}),&\text{ if $m$ is odd.}\end{cases}

These are rhombi, as can be seen in figure 11, and as can be deduced from the discussion around (3.16) since α0=αd+1\alpha_{0}=\alpha_{d+1}. It is also clear to see that for all mm\in\mathbb{N}.

(5.3) Sm+2=Φ2Sm.S_{m+2}=\Phi^{2}S_{m}.

In this case, it is simple to find a partition for the entirity of the middle cone CcC_{c} for the map RκR_{\kappa}. In particular, define the sets XX and YY to be

(5.4) X=Cc(Cc(λη))(Cd+1+η)=Cc(CcΦ3)(Cd+1+Φ2),X=C_{c}\cap(C_{c}-(\lambda-\eta))\cap(C_{d+1}+\eta)=C_{c}\cap(C_{c}-\Phi^{3})\cap(C_{d+1}+\Phi^{2}),

and

(5.5) Y=Cc(Cc+η)=Cc(Cc+Φ2).Y=C_{c}\cap(C_{c}+\eta)=C_{c}\cap(C_{c}+\Phi^{2}).

As we will see soon, the collection {X,Y,Sn:n2}\{X,Y,S_{n}:n\geq 2\} forms a partition of CcC_{c}. We are interested in the pre-image of these sets under EE. In particular, define the partition

𝒞={E1(S)Cj:j{1,,d},S{Y,X,S2,S3,,}}.\mathcal{C}^{\prime}=\left\{E^{-1}(S)\cap C_{j}:j\in\{1,...,d\},S\in\{Y,X,S_{2},S_{3},...,\}\right\}.
Refer to caption
Figure 11. The same partition as in figure 10, but after an application of EE, which reveals an alternating pattern of rhombi.

This partition can be seen in figure 10. As a consequence of the next theorem, (𝒞,Rκ)(\mathcal{C}^{\prime},R_{\kappa}) is a PWI with a countably infinite number of atoms. We also define a separate family of sets

𝒬={Qn,j=E1(Sn)Cj:n,j{1,,d}},\mathcal{Q}=\left\{Q_{n,j}=E^{-1}(S_{n})\cap C_{j}:n\in\mathbb{N},j\in\{1,...,d\}\right\},

which includes only the rhombi, and thus forms a partition of only a subset of the middle cone. Note that since λm=1\lambda_{m}=1 for all mm\in\mathbb{N} and m0=0m_{0}=0. We see that the set U(κ)U(\kappa) from Lemma 4.6 is given by

U(κ)\displaystyle U(\kappa) ={0}m=0Sm\displaystyle=\{0\}\cup\bigcup_{m=0}^{\infty}S_{m}
=(C0Δ0)Cc(Cd+1Δ1)\displaystyle=(C_{0}-\Delta_{0})\cap C_{c}\cap(C_{d+1}-\Delta_{-1})
=(C0Φ)Cc(Cd+1+1).\displaystyle=(C_{0}-\Phi)\cap C_{c}\cap(C_{d+1}+1).

Also observe that by removing S0S_{0} and S1S_{1} we get

U2,0(κ)={0}m=2Sm=(C0Φ3)Cc(Cd+1+Φ2).U_{2,0}(\kappa)=\{0\}\cup\bigcup_{m=2}^{\infty}S_{m}=(C_{0}-\Phi^{3})\cap C_{c}\cap(C_{d+1}+\Phi^{2}).

From this, we notice that

U2(κ)X=Cc(Cd+1+Φ2)((CcΦ3)(C0Φ3)),U_{2}(\kappa)\cup X=C_{c}\cap(C_{d+1}+\Phi^{2})\cap\left((C_{c}-\Phi^{3})\cup(C_{0}-\Phi^{3})\right),

but since Cd+1Φ3Cd+1C_{d+1}-\Phi^{3}\subset C_{d+1}, we know that Cc(Cd+1Φ3)=C_{c}\cap(C_{d+1}-\Phi^{3})=\emptyset, so

U2(κ)X\displaystyle U_{2}(\kappa)\cup X =Cc(Cd+1+Φ2)((Cd+1Φ3)(CcΦ3)(C0Φ3))\displaystyle=C_{c}\cap(C_{d+1}+\Phi^{2})\cap\left((C_{d+1}-\Phi^{3})\cup(C_{c}-\Phi^{3})\cup(C_{0}-\Phi^{3})\right)
=Cc(Cd+1+Φ2)¯\displaystyle=C_{c}\cap(C_{d+1}+\Phi^{2})\cap\overline{\mathbb{H}}
=Cc(Cd+1+Φ2).\displaystyle=C_{c}\cap(C_{d+1}+\Phi^{2}).

Therefore, we can see that

U2(κ)XY=Cc((Cc+Φ2)(Cd+1+Φ2)),U_{2}(\kappa)\cup X\cup Y=C_{c}\cap\left((C_{c}+\Phi^{2})\cup(C_{d+1}+\Phi^{2})\right),

and a similar argument tells us that Cc(C0+Φ2)=C_{c}\cap(C_{0}+\Phi^{2})=\emptyset and so finally we get,

U2(κ)XY=Cc((C0+Φ2)(Cc+Φ2)(Cd+1+Φ2))=Cc¯=Cc.U_{2}(\kappa)\cup X\cup Y=C_{c}\cap\left((C_{0}+\Phi^{2})\cup(C_{c}+\Phi^{2})\cup(C_{d+1}+\Phi^{2})\right)=C_{c}\cap\overline{\mathbb{H}}=C_{c}.

Therefore, 𝒞\mathcal{C}^{\prime} is a partition of CcC_{c} up to a set of Lebesgue measure 0.

Note that for all (m,n)𝒩Φ<(m,n)\in\mathcal{N}_{\Phi}^{<}, wΦ(m,1)=wΦ(m+1,0)w_{\Phi}(m,1)=w_{\Phi}(m+1,0), and so by recalling that wΦ(m,0)=wΦ(m)=mw_{\Phi}(m,0)=w_{\Phi}(m)=m, the condition that wΦ(m,1)>wΦ(m0,0)w_{\Phi}(m,1)>w_{\Phi}(m_{0},0) is equivalent to the condition that

wΦ(m+1,0)>wΦ(m0,0),w_{\Phi}(m+1,0)>w_{\Phi}(m_{0},0),

which is itself equivalent to

m>m01=1.m>m_{0}-1=-1.

With this in mind, Theorem 4.5 tells us that for all mm\in\mathbb{N}, if zE1(Sm)z\in E^{-1}(S_{m}), then

h(z)=hm,1=hm+1,0=hm+1=Fibm+31,h(z)=h_{m,1}=h_{m+1,0}=h_{m+1}=\operatorname{Fib}_{m+3}-1,

and

Rκ(z)=E(z)+Δm+1=E(z)(Φ)m+2,R_{\kappa}(z)=E(z)+\Delta_{m+1}=E(z)-(-\Phi)^{m+2},
Refer to caption
Figure 12. The same partition of CcC_{c} as in figure 10, after an application of RκR_{\kappa}, which has shifted the rhombi alternately. Note that there is an overlap between the cone and ribbon (both of which are more clearly seen in figure 11) on the top of the figure, causing an unavoidable overlap of the colours.

Observe that λ=Φ\lambda=\Phi is a special case of irrational number within [0,1][0,1] in the sense that it is a fixed point of the Gauss map gg. Thus, g2(Φ)=Φg^{2}(\Phi)=\Phi and thus we can choose η=η=1Φ=Φ2\eta^{\prime}=\eta=1-\Phi=\Phi^{2} and Theorem 4.7 tells us that the first return map RκR_{\kappa} exhibits exact self-similarity within U(κ)U(\kappa). In particular, for all zU0,0(κ)=U(κ)z\in U_{0,0}(\kappa)=U(\kappa), we have the following conjugacy

Rκ(z)=1Δ1Rκ((Δ1)z)=1Φ2Rκ(Φ2z).R_{\kappa}(z)=\frac{1}{-\Delta_{1}}R_{\kappa}((-\Delta_{1})z)=\frac{1}{\Phi^{2}}R_{\kappa}(\Phi^{2}z).

One consequence of this is that if there exists a periodic point zSmz\in S_{m} of period kk, then zz is a periodic point of RκR_{\kappa} with period k/hm+1k/h_{m+1}. The self-similarity shows that for all nn\in\mathbb{Z} such that 2nm2n\geq-m, Φ2nz\Phi^{2n}z is a periodic point of RκR_{\kappa}, thus also a periodic point of FκF_{\kappa} whose period is an integer multiple of hm+2n+1h_{m+2n+1}. In particular, there is a sequence (zn)n(z_{n})_{n\in\mathbb{N}} given by

(5.6) zn=Φ2nmz,z_{n}=\Phi^{2n-m}z,

so that for all nn\in\mathbb{N}, znS2n+m~z_{n}\in S_{2n+\tilde{m}} and the period of znz_{n} is an integer multiple of h2n+m~+1h_{2n+\tilde{m}+1}, where {0,1}m~m(mod2)\{0,1\}\ni\tilde{m}\cong m\pmod{2}.

Given a map f:XXf:X\rightarrow X, let 𝒪f+(x)\mathcal{O}_{f}^{+}(x) denote the forward orbit of xXx\in X under ff, that is

𝒪f+(x)={fn(x):n}.\mathcal{O}_{f}^{+}(x)=\{f^{n}(x):n\in\mathbb{N}\}.
Proposition 5.2.

Suppose there exists a periodic point zSmz\in S_{m} for some mm\in\mathbb{N}, and let (zn)n(z_{n})_{n\in\mathbb{N}} be the sequence of periodic points given by (5.6). Then the sequence (𝒪Fκ+(zn))n\left(\mathcal{O}_{F_{\kappa}}^{+}(z_{n})\right)_{n\in\mathbb{N}} of periodic orbits accumulates on the interval [1,Φ][-1,\Phi].

Proof.

Let nn\in\mathbb{N}. Note that

(5.7) {Fκj(zn):1jh(z)}𝒪Fκ+(zn).\{F_{\kappa}^{j}(z_{n}):1\leq j\leq h(z)\}\subset\mathcal{O}_{F_{\kappa}}^{+}(z_{n}).

Lemma 3.3 tells us that for all 1jh(zn)1\leq j\leq h(z_{n}),

Fκj(zn)=E(zn)+Fκj(0).F_{\kappa}^{j}(z_{n})=E(z_{n})+F_{\kappa}^{j}(0).

Therefore,

(5.8) |Fκj(zn)Fκj(0)|=|E(zn)|=|zn|.|F_{\kappa}^{j}(z_{n})-F_{\kappa}^{j}(0)|=|E(z_{n})|=|z_{n}|.

Let HH\in\mathbb{N}. Then there exists an NN\in\mathbb{N} such that h(zn)Hh(z_{n})\geq H for all nNn\geq N, and thus (5.8) holds for all 1jH1\leq j\leq H. Now let ε>0\varepsilon>0 be small. Then there exists an NN^{\prime}\in\mathbb{N} such that for all integers nNn\geq N^{\prime} such that

(5.9) |zn|<ε.|z_{n}|<\varepsilon.

Set N=max{N,N}N^{*}=\max\{N,N^{\prime}\}. Then for all integers nNn\geq N^{*}, both (5.8) holds for all 1jH1\leq j\leq H and (5.9) holds. Hence, for all nNn\geq N^{*} we have

|Fκj(z)Fκj(0)|<ε,|F_{\kappa}^{j}(z)-F_{\kappa}^{j}(0)|<\varepsilon,

for all 1jH1\leq j\leq H. Since HH and ε\varepsilon are independent and arbitrary, we conclude that the sequence (𝒪Fκ+(zn))n\left(\mathcal{O}_{F_{\kappa}}^{+}(z_{n})\right)_{n\in\mathbb{N}} accumulates on the set 𝒪Fκ+(0)\mathcal{O}_{F_{\kappa}}^{+}(0).

By Proposition 3.1, FκF_{\kappa} is a 2-IET everywhere on the interval [1,λ][-1,\lambda] except on the preimages of 0, since Fκ(0)=η=1λF_{\kappa}(0)=-\eta=1-\lambda, contrary to Fκ(x)=x+λF_{\kappa}(x)=x+\lambda for x[1,0)x\in[-1,0) and Fκ(x)=x1F_{\kappa}(x)=x-1 for x(0,λ)x\in(0,\lambda).

Therefore FκF_{\kappa} is conjugate to an irrational rotation almost everywhere (with respect to one-dimensional Lebesgue measure), since λ=Φ\lambda=\Phi is irrational. In particular, since λ\lambda is irrational and η=1λ\eta=1-\lambda, we know, by for example Lemma 4.3, that Fκj(0)=Fκj1(η)F_{\kappa}^{j}(0)=F_{\kappa}^{j-1}(-\eta) is bounded away from 0 for all integers j>0j>0, so Fκj(0)0F_{\kappa}^{j}(0)\neq 0 for any j>0j>0.

Hence, the orbit of Fκ(0)=ηF_{\kappa}(0)=-\eta under FκF_{\kappa} is also the orbit under an irrational rotation, and thus the orbit of 0 is dense in the interval [1,λ][-1,\lambda], i.e.

𝒪Fκ+(0)¯=[1,λ].\overline{\mathcal{O}_{F_{\kappa}}^{+}(0)}=[-1,\lambda].

Therefore, the sequence (𝒪Fκ+(zn))n\left(\mathcal{O}_{F_{\kappa}}^{+}(z_{n})\right)_{n\in\mathbb{N}} accumulates on the interval [1,λ][-1,\lambda]. ∎

Remark 5.3.

Although extending Proposition 5.2 to periodic continued fractions λ\lambda should follow from a similar strategy to the proof used here, an extension to aperiodic continued fractions seems to require nothing short of assuming/proving that every TCE has at least one periodic point in its ‘renormalizable domain’ U(κ)U(\kappa).

6. Discussion

Translated cone exchanges, introduced first in [16] and investigated in [26, 28], are an interesting and largely unexplored family of parametrised PWIs. They contain an embedding of a simple IET on the baseline and as such they are an interesting tool to understand more general PWIs by gaining leverage from known results for IETs. In this paper we go beyond results in [16, 26, 28] to show that for a dense subset of an open set in the parameter space of TCEs there is a mapping (κκ\kappa\mapsto\kappa^{\prime} in Theorem 4.7) that determines a renormalization scheme for the first return map RκR_{\kappa} of FκF_{\kappa} to the vertex 0 of the middle cone CcC_{c}. This helps us describe the small-scale, long-term behaviour of FκF_{\kappa} near the baseline [1,λ][-1,\lambda] via the large-scale, short-term behaviour of Fκ′′F_{\kappa^{\prime\prime}} with κ′′=(α,τ,λ′′,η′′,1)\kappa^{\prime\prime}=(\alpha,\tau,\lambda^{\prime\prime},\eta^{\prime\prime},1), λ′′=g2k(λ)\lambda^{\prime\prime}=g^{2k}(\lambda), a large enough integer k>0k>0 and some suitably chosen η′′\eta^{\prime\prime} described by (4.9). Proposition 5.2 is an example of this, where a periodic disk of small period for FκF_{\kappa} and the periodicity of the continued fraction coefficients of λ=Φ\lambda=\Phi give rise to an countable collection of periodic disks of arbitrarily high period clustering on [1,λ][-1,\lambda], through the renormalizability established by Theorem 4.7.

Although these results give a glimpse into the dynamics for orbits close to the baseline, there remains a lot more to do to understand the dynamics of these TCEs near general points in exceptional set, but this seems to be a far more complex task to undertake, especially as the dynamics near the baseline primarily consists of horizontal translations, whereas in general the effect of the rotations will be inextricably linked to translations.

Acknowledgements

We thank Pedro Peres and Arek Goetz for discussions about this research. NC and PA thank the Mittag-Leffler Institute for their hospitality and support to visit during the “Two Dimensional Maps” programme of early 2023.

Funding Acknowledgement

This work was supported by the Engineering and Physical Sciences Research Council.

Data Access Statement

For the purpose of open access, the authors have applied a Creative Commons Attribution (CC BY) licence to any Author Accepted Manuscript version arising from this submission.

No new data were generated or analysed during this study. The figures in this study were produced by python programming code written by NC. This code, namely the ‘pyTCE’ program, is publicly available at the following link:

https://github.com/NoahCockram/pyTCE/tree/main

References

  • [1] Masur, H. (1982). Interval Exchange Transformations and Measured Foliations, Ann. of Math. 115, no. 1, 169–200.
  • [2] Boshernitzan, M. D. and Carroll, C. R. (1997). An extension of Lagrange’s theorem to interval exchange transformations over quadratic fields. C.R. J. Anal. Math. 72: 21.
  • [3] Zorich, A. (1996). Finite Gauss measure on the space of interval exchange transformation. Lyapunov exponents, Ann. Inst. Fourier, Grenoble, 46, 325-370.
  • [4] Marchese, L. (2011). The Khinchin Theorem for interval-exchange transformations, Journal of Modern Dynamics 5, no. 1, 123-183.
  • [5] Ferenczi S. and Zamboni L. Q., (2008). Languages of k-interval exchange transformations, Bulletin of the London Mathematical Society 40, no. 4, 705–714.
  • [6] Avila, A. and Giovanni F., (2007). Weak Mixing for Interval Exchange Transformations and Translation Flows, Ann. of Math. 165, no. 2, 637–64.
  • [7] Katok, A. B. (1980). Interval exchange transformations and some special flows are not mixing, Israel J. Math. 35, 301-310.
  • [8] Buzzi, J. (2001). Piecewise isometries have zero topological entropy. Ergodic Theory and Dynamical Systems, 21, 1371–1377.
  • [9] Goetz, A. (2000). A self-similar example of a piecewise isometric attractor, Dynamical Systems: From Crystal to Chaos, 248-258.
  • [10] Goetz, A. and Poggiaspalla, G. (2004). Rotations by π/7\pi/7. Nonlinearity, 17, 1787–1802.
  • [11] Adler, R., Kitchens, B. and Tresser, C. (2001). Dynamics of non-ergodic piecewise affine maps of the torus. Ergodic Theory and Dynamical Systems, 21, 959–999.
  • [12] Lowenstein J. H., and Kouptsov, K. L. and Vivaldi, F. (2003). Recursive tiling and geometry of piecewise rotations by π/7\pi/7. Nonlinearity, 17, 371–395.
  • [13] Ashwin, P., and Goetz, A. (2005). Invariant Curves and Explosion of Periodic Islands in Systems of Piecewise Rotations. SIAM Journal on Applied Dynamical Systems, 4, 437–458.
  • [14] Ashwin, P., and Goetz, A. (2006). Polygonal invariant curves for a planar piecewise isometry. Trans. Amer. Math. Soc. 358 no. 1, 373-390.
  • [15] Poggiaspalla, G. (2006). Self-similarity in piecewise isometric systems. Dynamical Systems, 21, 147–189.
  • [16] Ashwin, P., Goetz, A., Peres, P. and Rodrigues, A. (2020). Embeddings of interval exchange transformations into planar piecewise isometries. Ergodic Theory and Dynamical Systems, 40, 1153–1179.
  • [17] Lynn, T. F., Ottino, J. M., Umbanhowar, P. B. and Lueptow, R. M. (2020). Identifying invariant ergodic subsets and barriers to mixing by cutting and shuffling: Study in a birotated hemisphere, 101, no. 1.
  • [18] Kahng, B. (2009). Singularities of two-dimensional invertible piecewise isometric dynamics, Chaos: An Interdisciplinary Journal of Nonlinear Science, 19, no. 2.
  • [19] Wilson, K. G. (1983). The renormalization group and critical phenomena, Reviews of Modern Physics, 55, no. 3, 583–600.
  • [20] Coullet, P. and Tresser, C. (1978). Itérations d’endomorphismes et groupe de renormalisation, Journal de Physique Colloques, 39, no. C5, 25–28.
  • [21] Feigenbaum, M. J. (1978). Quantitative universality for a class of nonlinear transformations, Journal of Statistical Physics, 19, 25–52.
  • [22] Feigenbaum, M. J. (1979). The universal metric properties of nonlinear transformations, Journal of Statistical Physics, 21, 669–706.
  • [23] McMullen, C. T. (1994), Complex dynamics and renormalization, Princeton University Press, no. 135.
  • [24] Rauzy, G. (1979). Échanges d’intervalles et transformations induites, Acta Arithmetica, 34, no. 4, 315–328.
  • [25] Veech, W. A. (1982). Gauss measures for transformations on the space of interval exchange maps, Ann. of Math. 115, 201-242.
  • [26] Peres, P. (2019). Renormalization in Piecewise Isometries, PhD thesis, University of Exeter.
  • [27] Bates, B., Bunder, M. and Tognetti, K. (2005). Continued Fractions and The Gauss Map. Academia Paedagogica Nyiregyhaziensis. Acta Mathematica, 21, 113–125.
  • [28] Peres, P. and Rodrigues, A. (2018). On the dynamics of Translated Cone Exchange Transformations. arXiv preprint arXiv:1809.05496.