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

Visible part of dominated self-affine sets in the plane

Eino Rossi [Eino Rossi] Department of Mathematics and Statistics, University of Helsinki
P.O. Box 68 (Pietari Kalmin katu 5)
00014 University of Helsinki
Finland
[email protected]
Abstract.

The dimension of the visible part of self-affine sets, that satisfy domination and a projection condition, is being studied. The main result is that the Assouad dimension of the visible part equals to 1 for all directions outside the set of limit directions of the cylinders of the self-affine set. The result holds regardless of the overlap of the cylinders. The sharpness of the result is also being discussed.

Key words and phrases:
Visible part, self-affine, weak tangent.
2010 Mathematics Subject Classification:
28A80
The author was funded by the Academy of Finland through project Nos. 314829(Frontiers of singular integrals) and 309365(Quantitative rectifiability in Euclidean and non-Euclidean spaces).

1. Introduction

For eS1e\in S^{1}, let (e)\ell(e) denote the half line starting from origin and propagating to direction ee. That is (e)={te:t0}\ell(e)=\{te:t\geq 0\}. For a compact set E2E\subset\mathbb{R}^{2} the visible part of EE in direction eS1e\in S^{1} is the set of points in x2x\in\mathbb{R}^{2} that satisfy

({x}+(e))E={x}.(\{x\}+\ell(e))\cap E=\{x\}.

This set is denoted by ViseE\operatorname{Vis}^{e}E. Let proje\operatorname{proj}^{e} denote the orthogonal projection along the direction ee. (Note that ViseE\operatorname{Vis}^{e}E may be different from ViseE\operatorname{Vis}^{-e}E, but projeE=projeE\operatorname{proj}^{e}E=\operatorname{proj}^{-e}E always.) Consider the Hausdorff dimension of the visible part of a compact set EE. If dimHE<1\operatorname{dim_{H}}E<1 then dimHprojeE=dimHE\operatorname{dim_{H}}\operatorname{proj}^{e}E=\operatorname{dim_{H}}E for almost all eS1e\in S^{1} by Marstrand’s projection theorem [16]. Since ViseEE\operatorname{Vis}^{e}E\subset E and projeViseE=projeE\operatorname{proj}^{e}\operatorname{Vis}^{e}E=\operatorname{proj}^{e}E, it follows that dimHViseE=dimHE\operatorname{dim_{H}}\operatorname{Vis}^{e}E=\operatorname{dim_{H}}E for almost all eS1e\in S^{1}. If dimHE1\operatorname{dim_{H}}E\geq 1, then still we have that 1dimHViseE1\leq\operatorname{dim_{H}}\operatorname{Vis}^{e}E for almost all eS1e\in S^{1}, but the upper bound dimHViseEdimHE\operatorname{dim_{H}}\operatorname{Vis}^{e}E\leq\operatorname{dim_{H}}E should no longer be optimal for most eS1e\in S^{1}. The visibility conjecture states that dimHViseE=1\operatorname{dim_{H}}\operatorname{Vis}^{e}E=1 for almost all eS1e\in S^{1}. Obviously one can not hope this to hold for all directions, since a graph of a function can have dimension greater than 11 for example. Further, an example of Davies and Fast [6] shows that dimHVise(K)=2\operatorname{dim_{H}}\operatorname{Vis}^{e}(K)=2 is possible for a dense GδG_{\delta} set of directions. This is the furthest one can go, since recently Orponen [20] showed that it is impossible to have dimHVise(K)=2\operatorname{dim_{H}}\operatorname{Vis}^{e}(K)=2 for set of directions of positive measure. It is rather easy to see that the visibility conjecture is false for the box counting dimension and thus for the Assouad dimension as well. This follows, since a countable set equals to its visible part for almost all directions and there exist compact countable sets with full box dimension. For example, one can simply consider K=A×AK=A\times A, where A={0}{(Sn)1}n=1A=\{0\}\cup\{(S_{n})^{-1}\}_{n=1}^{\infty} and Sn=k=1n1/kS_{n}=\sum_{k=1}^{n}1/k. For details, see Example 5.2.

The visibility conjecture has been confirmed in a few special cases: Järvenpää et.al. [8] proved the conjecture for quasi-circles, Arhosalo et al. [1] confirmed that for fractal percolation the conjecture holds almost surely, and Falconer and Fraser [7] showed that the conjecture holds for self-similar sets satisfying a projection condition and the open set condition so that the open set can be chosen to be convex. In all these cases, the authors actually verified the conjecture for the box dimension and for all directions eS1e\in S^{1}. See also the recent work of Järvenpää et.al. [9].

One obvious variant of the problem is to consider the visible set from a given point instead of a direction. O’Neil [18] showed that for compact connected subsets of 2\mathbb{R}^{2}, the Hausdorff dimension of the visible part from a point x2x\in\mathbb{R}^{2} is strictly less than the Hausdorff dimension of the original set, and it is uniformly bounded away from 22, for almost all viewpoints xx. An other related problem is to determine when ViseE=E\operatorname{Vis}^{e}E=E. Orponen [19] showed that if dimHE>1\operatorname{dim_{H}}E>1, then the set of directions for which ViseE=E\operatorname{Vis}^{e}E=E has Hausdorff dimension at most 2dimHE2-\operatorname{dim_{H}}E. On the other hand, it follows from the main result of [22], that if dimBE<1/2\operatorname{dim_{B}}E<1/2, then ViseE=E\operatorname{Vis}^{e}E=E holds outside a set of directions of box dimension 2dimBE2\operatorname{dim_{B}}E. For other related results, see for example [5, 4, 23].

In this paper I study the visible parts of self-affine sets. Domination and projection condition are standing assumptions throughout the paper. Theorem 2.4 is the main result and it says that the Assouad dimension of the visible part equals to 11 for all directions outside the set of the limit directions given by the affine dynamics. This theorem then has several corollaries. Corollary 4.3 says that for dominated self-affine carpets the Assouad dimension of the visible part equals to 11 for all but two exceptional directions (that span the same line). Corollary 4.4 says that if the self-affine system satisfies the strong cone separation, then the Assouad dimension of the visible part equals to 11 for almost all directions. These results can be seen as rather strong, considering how easily the Assouad dimension jumps up in different situations. For example, it is well known that the fractal percolation has equal Hausdorff and box dimension <2<2 but full Assouad dimension, and Assouad dimension also tends to be maximal in projections in a way that is impossible for Hausdorff or box dimension [21]. Corollary 4.5 studies the case where the limit directions of the cylinders do not overlap too much, and states that the Hausdorff dimension of the visible part equals to 11 for all directions in this case.

Acknowledgement.

I want to thank Balázs Bárány, Antti Käenmäki, and Tuomas Orponen for inspiring discussions on the topics of this paper. I also wish to thank the anonymous referee for the valuable comments on how to improve the quality of this paper.

2. Preliminaries and statement of the main result

The purpose of this section is only to fix the setting of the paper and state the main result. In the next sections, along the course of the proof, I give more insight by explaining the geometry behind the assumptions and the result.

Throughout the paper, a direction means a unit vector eS1e\in S^{1} and orientation is an element of the projective space 1\mathbb{P}^{1}, that is, the metric space of lines in 2\mathbb{R}^{2} that go through origin, and where the distance is measured by the angle between the lines. For a vector e2e\in\mathbb{R}^{2}, let e={te:t}\left<e\right>=\{te:t\in\mathbb{R}\} denote the corresponding element of the projective space. It is sometimes more intuitive to think S1S^{1} as a set of angles instead of unit vectors. This justifies the use of notations θ:={t(cosθ,sinθ):t}\langle\theta\rangle:=\{t(\cos\theta,\sin\theta):t\in\mathbb{R}\}, projθ:=proj(cosθ,sinθ)\operatorname{proj}^{\theta}:=\operatorname{proj}^{(\cos\theta,\sin\theta)}, Visθ:=Vis(cosθ,sinθ)\operatorname{Vis}^{\theta}:=\operatorname{Vis}^{(\cos\theta,\sin\theta)}, and (θ):=((cosθ,sinθ))\ell(\theta):=\ell((\cos\theta,\sin\theta)) for θ\theta\in\mathbb{R}.

Let A:22A\colon\mathbb{R}^{2}\to\mathbb{R}^{2} be an invertible linear map, so that A(B(0,1))A(B(0,1)) is not a ball. Then A(B(0,1))A(B(0,1)) is an ellipse whose semiaxes have different lengths. Let α1(A)\alpha_{1}(A) be the length of the longer one of the semiaxes and let α2(A)\alpha_{2}(A) be the length of the shorter one. Equivalently αk(A),k=1,2\alpha_{k}(A),k=1,2 are the square roots of the eigenvalues of ATAA^{T}A (ordered so that the larger is α1\alpha_{1}). Also, set ϑ1(A)1\vartheta_{1}(A)\in\mathbb{P}^{1} to be the orientation of the longer semiaxes of A(B(0,1))A(B(0,1)). That is, ϑ1(A)=Aη1(A)\vartheta_{1}(A)=\left<A\eta_{1}(A)\right>, where η1(A)\eta_{1}(A) is the normalized eigenvector of ATAA^{T}A associated to the eigenvalue α1(A)2\alpha_{1}(A)^{2}. Likewise, set ϑ2(A)=Aη2(A)\vartheta_{2}(A)=\left<A\eta_{2}(A)\right>, where η2(A)\eta_{2}(A) is the normalized eigenvector corresponding to α2(A)2\alpha_{2}(A)^{2}. It is a basic fact that ϑ1(A)ϑ2(A)\vartheta_{1}(A)\perp\vartheta_{2}(A) and η1(A)η2(A)\eta_{1}(A)\perp\eta_{2}(A).

Let {Ai}i=1κ\{A_{i}\}_{i=1}^{\kappa} be a collection of contractive invertible linear maps, let {ci}i=1κ\{c_{i}\}_{i=1}^{\kappa} be a collection of vectors in 2\mathbb{R}^{2}, and let φi(x)=Aix+ci\varphi_{i}(x)=A_{i}x+c_{i}, for all i{1,,κ}i\in\{1,\ldots,\kappa\}. It standard that there exists a unique compact set EE satisfying

E=i=1κφi(E).E=\bigcup_{i=1}^{\kappa}\varphi_{i}(E).

The set EE is called self-affine.

Set Σ=k{1,,κ}k\Sigma^{*}=\bigcup_{k\in\mathbb{N}}\{1,\ldots,\kappa\}^{k} and Σ={1,,κ}\Sigma=\{1,\ldots,\kappa\}^{\mathbb{N}}. Write Σn\Sigma^{n} for {1,,κ}n\{1,\ldots,\kappa\}^{n} even though this is abusing the notation. Let |𝚒||\mathtt{i}| denote the length of the word 𝚒\mathtt{i}. That is, |𝚒|=n|\mathtt{i}|=n whenever 𝚒Σn\mathtt{i}\in\Sigma^{n} and |𝚒|=|\mathtt{i}|=\infty, when 𝚒Σ\mathtt{i}\in\Sigma. For 𝚒,𝚓Σ\mathtt{i},\mathtt{j}\in\Sigma, let 𝚒𝚓\mathtt{i}\wedge\mathtt{j} be the longest common beginning of 𝚒\mathtt{i} and 𝚓\mathtt{j} and define a distance function ϱ\varrho in Σ\Sigma by setting ϱ(𝚒,𝚓)=2|𝚒𝚓|\varrho(\mathtt{i},\mathtt{j})=2^{-|\mathtt{i}\wedge\mathtt{j}|}, with the interpretation 2=02^{-\infty}=0. This makes (Σ,ϱ)(\Sigma,\varrho) a compact metric space.

For quantities xx and yy, usually depending on 𝚒\mathtt{i}, the notation xyx\lesssim y means that there is a constant C>1C>1, that may depend on the self-affine set EE, so that xCyx\leq Cy. Further, xyx\approx y means that xyx\lesssim y and yxy\lesssim x.

For 𝚒=(i1,i2,,in)Σ\mathtt{i}=(i_{1},i_{2},\ldots,i_{n})\in\Sigma^{*}, let A𝚒=Ai1Ai2AinA_{\mathtt{i}}=A_{i_{1}}A_{i_{2}}\dots A_{i_{n}} and for the sake of brevity, write αk(𝚒)=αk(A𝚒)\alpha_{k}(\mathtt{i})=\alpha_{k}(A_{\mathtt{i}}) and ϑk(𝚒)=ϑk(A𝚒)\vartheta_{k}(\mathtt{i})=\vartheta_{k}(A_{\mathtt{i}}) for k=1,2k=1,2. Similarly, also write φ𝚒=φi1φi2φin\varphi_{\mathtt{i}}=\varphi_{i_{1}}\circ\varphi_{i_{2}}\circ\dots\circ\varphi_{i_{n}}. The line ϑ1(𝚒)\vartheta_{1}(\mathtt{i}) is called the orientation of the cylinder φ𝚒(E)\varphi_{\mathtt{i}}(E), because the cylinder φ𝚒(E)\varphi_{\mathtt{i}}(E) is “close” to being a line segment that is parallel to the line ϑ1(𝚒)\vartheta_{1}(\mathtt{i}), at least when |𝚒||\mathtt{i}| is large. This phenomena is examined in more detail in Proposition 3.4. As usual, let π:ΣE\pi\colon\Sigma\to E be the canonical projection defined by

{π𝚒}=n=1φ𝚒|n(E).\{\pi\mathtt{i}\}=\bigcap_{n=1}^{\infty}\varphi_{\mathtt{i}|_{n}}(E).

From time to time, the set φ𝚒|n(E)\varphi_{\mathtt{i}|_{n}}(E) is also denoted by E𝚒|nE_{\mathtt{i}|_{n}}. The system {Ai}i=1κ\{A_{i}\}_{i=1}^{\kappa} is called dominated, or said to satisfy dominated splitting, if there are constants τ>1\tau>1 and n0n_{0}\in\mathbb{N}, so that α1(𝚒)>τ|𝚒|α2(𝚒)\alpha_{1}(\mathtt{i})>\tau^{|\mathtt{i}|}\alpha_{2}(\mathtt{i}) for all 𝚒Σ\mathtt{i}\in\Sigma^{*}, with |𝚒|n0|\mathtt{i}|\geq n_{0}. Domination ensures the existence of the limit orientation for all symbols 𝚒Σ\mathtt{i}\in\Sigma. The next lemma records this fact along with other useful properties of the limit orientations.

Lemma 2.1.

Let EE be a dominated self-affine set. Then

  1. (1)

    ϑ1(𝚒)=limnϑ1(𝚒|n)\vartheta_{1}(\mathtt{i})=\lim_{n\to\infty}\vartheta_{1}(\mathtt{i}|_{n}) exists for all 𝚒Σ\mathtt{i}\in\Sigma and the convergence is uniform.

  2. (2)

    The map ϑ1:Σ1\vartheta_{1}\colon\Sigma\to\mathbb{P}^{1} is uniformly continuous.

  3. (3)

    ϑ1(Σ)\vartheta_{1}(\Sigma) contains the accumulation points of the set {ϑ1(𝚒):𝚒Σ}\{\vartheta_{1}(\mathtt{i}):\mathtt{i}\in\Sigma^{*}\}.

  4. (4)

    A𝚒ϑ1(𝚓)=ϑ1(𝚒𝚓)A_{\mathtt{i}}\vartheta_{1}(\mathtt{j})=\vartheta_{1}(\mathtt{i}\mathtt{j}) for all 𝚒Σ\mathtt{i}\in\Sigma^{*} and 𝚓Σ\mathtt{j}\in\Sigma.

Proof.

The proof of (1) is a direct modification of [10, Lemma 2.1], where the existence of the limit in question is showed for almost all 𝚒\mathtt{i}. The proof there works for individual 𝚒Σ\mathtt{i}\in\Sigma for which

lim infn1nlogα2(𝚒|n)α1(𝚒|n)>0.\liminf_{n\to\infty}-\frac{1}{n}\log\frac{\alpha_{2}(\mathtt{i}|_{n})}{\alpha_{1}(\mathtt{i}|_{n})}>0.

In the setting of this paper, the domination implies the uniform bound logτ>0\log\tau>0 for the above liminf. The uniform bound also implies that the convergence is uniform. The part (2) follows from (1), and (3) follows from (2) and compactness of Σ\Sigma.

To prove (4) it suffices to show that A𝚒1ϑ1(𝚒𝚓|n)A_{\mathtt{i}}^{-1}\vartheta_{1}(\mathtt{i}\mathtt{j}|_{n}) converges to ϑ1(𝚓)\vartheta_{1}(\mathtt{j}) as nn\to\infty, since A𝚒A_{\mathtt{i}} is a diffeomorphism. Write

η1(𝚒𝚓|n)=tnη1(𝚓|n)+snη2(𝚓|n),\eta_{1}(\mathtt{i}\mathtt{j}|_{n})=t_{n}\eta_{1}(\mathtt{j}|_{n})+s_{n}\eta_{2}(\mathtt{j}|_{n}),

and for now let θk(𝚓|n)S1\theta_{k}(\mathtt{j}|_{n})\in S^{1} be a unit vector with θk(𝚓|n)=ϑk(𝚓|n)\left<\theta_{k}(\mathtt{j}|_{n})\right>=\vartheta_{k}(\mathtt{j}|_{n}) for k=1,2k=1,2. Then it follows from domination that

A𝚒1ϑ1(𝚒𝚓|n)\displaystyle A_{\mathtt{i}}^{-1}\vartheta_{1}(\mathtt{i}\mathtt{j}|_{n}) =A𝚒1A𝚒A𝚓|nη1(𝚒𝚓|n)=tnα1(𝚓n)θ1(𝚓|n)+snα2(𝚓n)θ2(𝚓|n)\displaystyle=\left<A_{\mathtt{i}}^{-1}A_{\mathtt{i}}A_{\mathtt{j}|_{n}}\eta_{1}(\mathtt{i}\mathtt{j}|_{n})\right>=\left<t_{n}\alpha_{1}(\mathtt{j}_{n})\theta_{1}(\mathtt{j}|_{n})+s_{n}\alpha_{2}(\mathtt{j}_{n})\theta_{2}(\mathtt{j}|_{n})\right>
=θ1(𝚓|n)+snα2(𝚓|n)tnα1(𝚓|n)θ2(𝚓|n)ϑ1(𝚓)\displaystyle=\left<\theta_{1}(\mathtt{j}|_{n})+\frac{s_{n}\alpha_{2}(\mathtt{j}|_{n})}{t_{n}\alpha_{1}(\mathtt{j}|_{n})}\theta_{2}(\mathtt{j}|_{n})\right>\to\vartheta_{1}(\mathtt{j})

as long as tnt_{n} stays bounded away from zero. To show that it does, recall that |A𝚒𝚓|nη1(𝚒𝚓|n)|=maxvS1|A𝚒𝚓|nv||A_{\mathtt{i}\mathtt{j}|_{n}}\eta_{1}(\mathtt{i}\mathtt{j}|_{n})|=\max_{v\in S^{1}}|A_{\mathtt{i}\mathtt{j}|_{n}}v|. In particular,

|A𝚒𝚓|nη1(𝚓|n)||A𝚒𝚓|nη1(𝚒𝚓|n)|,|A_{\mathtt{i}\mathtt{j}|_{n}}\eta_{1}(\mathtt{j}|_{n})|\leq|A_{\mathtt{i}\mathtt{j}|_{n}}\eta_{1}(\mathtt{i}\mathtt{j}|_{n})|,

where the left hand side is at least α2(𝚒)α1(𝚓|n)\alpha_{2}(\mathtt{i})\alpha_{1}(\mathtt{j}|_{n}) and the right hand side is at most

α1(𝚒)|tA𝚓|nη1(𝚓|n)+sA𝚓|nη2(𝚓|n)|α1(𝚒)|tnα1(𝚓|n)+snα2(𝚓|n)|.\alpha_{1}(\mathtt{i})|tA_{\mathtt{j}|_{n}}\eta_{1}(\mathtt{j}|_{n})+sA_{\mathtt{j}|_{n}}\eta_{2}(\mathtt{j}|_{n})|\leq\alpha_{1}(\mathtt{i})|t_{n}\alpha_{1}(\mathtt{j}|_{n})+s_{n}\alpha_{2}(\mathtt{j}|_{n})|.

Thus the triangle inequality gives

α2(𝚒)α1(𝚒)|tn|+|sn|α2(𝚓|n)α1(𝚓|n)\frac{\alpha_{2}(\mathtt{i})}{\alpha_{1}(\mathtt{i})}\leq|t_{n}|+|s_{n}|\frac{\alpha_{2}(\mathtt{j}|_{n})}{\alpha_{1}(\mathtt{j}|_{n})}

and so the domination implies that |tn|21α2(𝚒)α1(𝚒)1|t_{n}|\geq 2^{-1}\alpha_{2}(\mathtt{i})\alpha_{1}(\mathtt{i})^{-1} for large nn. ∎

In addition to domination, a crucial assumption in this paper is the following projection condition.

Definition 2.2.

An affine IFS {φi}\{\varphi_{i}\} (or the invariant set EE) satisfies the projection condition if 1{ϑ1(𝚓):𝚓Σ}\mathbb{P}^{1}\setminus\{\vartheta_{1}(\mathtt{j}):\mathtt{j}\in\Sigma\}\neq\emptyset and if for all eS1e\in S^{1} with e1{ϑ1(𝚓):𝚓Σ}\langle e\rangle\in\mathbb{P}^{1}\setminus\{\vartheta_{1}(\mathtt{j}):\mathtt{j}\in\Sigma\}, there is n0n_{0} so that projeφ𝚒(E)\operatorname{proj}^{e}\varphi_{\mathtt{i}}(E) is a non-trivial interval for all 𝚒Σn\mathtt{i}\in\Sigma^{n} and nn0n\geq n_{0}.

Remark 2.3.

To check the projection condition in a specific case, it may be useful to note that affinie mappings preserve lines and convex hulls. (The convex hull of a set EE is the smallest convex set containing EE.) That is, if \ell is a line in 2\mathbb{R}^{2} and KK is the convex hull of EE and AA is an invertible affine map, then AA\ell is also a line and AKAK is the convex hull of AEAE. Asking if projeφ𝚒(E)\operatorname{proj}^{e}\varphi_{\mathtt{i}}(E) is an interval, is equivalent to asking if projA𝚒1e(E)\operatorname{proj}^{A_{\mathtt{i}}^{-1}e}(E) is an iterval. Further, projA𝚒1e(E)\operatorname{proj}^{A_{\mathtt{i}}^{-1}e}(E) is an interval if and only if every line \ell of the form {y}+A𝚒1e\{y\}+\langle A_{\mathtt{i}}^{-1}e\rangle that meets KK also meets EE.

As said, the purpose is to study the dimension of the visible part. I assume that the reader is familiar with basic notions of dimension. The Hausdorff dimension is denoted by dimH\operatorname{dim_{H}}, the box dimension by dimB\operatorname{dim_{B}}, and the Assouad dimension by dimA\operatorname{dim_{A}}. The definitions of Hausdorff and box dimension one can find from almost any text book of fractal geometry or geometric measure theory (see for example [17]), and for Assouad dimension one can check for example [14]. If the reader is not interested in the Assouad dimension, then I just want to remark that the results are new also for the Hausdorff dimension and that the versions with Assouad dimension are just stronger since dimHKdimAK\operatorname{dim_{H}}K\leq\operatorname{dim_{A}}K for all sets KK.

The main result is the following theorem.

Theorem 2.4.

Let EE be a self-affine set satisfying the projection condition and the dominated splitting. Then dimHVise(E)=dimAVise(E)=1\operatorname{dim_{H}}\operatorname{Vis}^{e}(E)=\operatorname{dim_{A}}\operatorname{Vis}^{e}(E)=1 for all eS1e\in S^{1} with e{ϑ1(𝚒):𝚒Σ}\langle e\rangle\not\in\{\vartheta_{1}(\mathtt{i}):\mathtt{i}\in\Sigma\}.

The proof goes via weak tangents, defined as follows. Let FnF_{n} be a sequence of compact sets in 2\mathbb{R}^{2}. Say that FnF_{n} converges to a compact set F2F\subset\mathbb{R}^{2} in B(0,R)B(0,R), if

sup{dist(Fn,x):xFB(0,R)}0 as n\sup\{\operatorname{dist}\left(F_{n},x\right):x\in F\cap B(0,R)\}\to 0\text{ as }n\to\infty

and

sup{dist(x,F):xFnB(0,R)}0 as n.\sup\{\operatorname{dist}\left(x,F\right):x\in F_{n}\cap B(0,R)\}\to 0\text{ as }n\to\infty.

For x2x\in\mathbb{R}^{2} and r>0r>0 we write Mx,rM_{x,r} for the magnification function that shifts xx to origin and scales with factor r1r^{-1}. That is

Mx,r(y)=yxr.M_{x,r}(y)=\frac{y-x}{r}.

Let X2X\subset\mathbb{R}^{2} be compact. Then WB(0,1)W\subset B(0,1) is said to be a weak tangent of XX, and written WTan(X)W\in\operatorname{Tan}(X), if Mxn,rn(X)M_{x_{n},r_{n}}(X) converges to WW in B(0,1)B(0,1) for some sequences (xn)X(x_{n})\subset X and rn0r_{n}\searrow 0. It is typical to consider the weak tangents as subsets of the unit ball (or the unit square), but this is just a convenient choice. One could as well consider the convergence in B(0,R)B(0,R) for any fixed R>0R>0 or for all R>0R>0 to allow the weak tangents to be unbounded as well.

Assuming Proposition 4.1, which deals with the dimension of the weak tangents of the visible part, its proof of Theorem 2.4 is rather simple.

Proof of theorem 2.4.

Consider eS1e\in S^{1} with e{ϑ1(𝚒):𝚒Σ}\langle e\rangle\not\in\{\vartheta_{1}(\mathtt{i}):\mathtt{i}\in\Sigma\}. By the projection condition, it holds that projeE𝚒\operatorname{proj}^{e}E_{\mathtt{i}} is an interval for some 𝚒Σ\mathtt{i}\in\Sigma^{*}, so dimHViseE1\operatorname{dim_{H}}\operatorname{Vis}^{e}E\geq 1. Thus the task is to prove the upper bound. Proposition 4.1 says that dimHW1\operatorname{dim_{H}}W\leq 1 for all weak tangents WW of ViseE¯\overline{\operatorname{Vis}^{e}E} (the closure is needed since the visible part is not necessarily closed). Recalling that the Assouad dimension of a compact set equals to the maximum of the Hausdorff dimensions of its weak tangents [11, Proposition 5.8], it then follows that dimHViseEdimAViseE¯dimHW1\operatorname{dim_{H}}\operatorname{Vis}^{e}E\leq\operatorname{dim_{A}}\overline{\operatorname{Vis}^{e}E}\leq\operatorname{dim_{H}}W\leq 1, were WW is a weak tangent of ViseE¯\overline{\operatorname{Vis}^{e}E} with maximal Hausdorff dimension. ∎

Considering the proof of Proposition 4.1, it is obvious that if WW is a weak tangent of ViseE¯\overline{\operatorname{Vis}^{e}E} with Mxn,rnViseE¯WM_{x_{n},r_{n}}\overline{\operatorname{Vis}^{e}E}\to W in B(0,1)B(0,1), then (by passing to a subsequence if necessary) it also holds that (Mxn,rnE)B(0,1)(M_{x_{n},r_{n}}E)\cap B(0,1) converges to a weak tangent, say TT, of EE. Of course WTW\subset T, but unfortunately, it is not generally true that WViseT¯W\subset\overline{\operatorname{Vis}^{e}T} or WViseT¯W\supset\overline{\operatorname{Vis}^{e}T}. In particular, one can not just take the weak tangent TT of EE of maximal dimension and expect it to have anything to do with the weak tangent of ViseE¯\overline{\operatorname{Vis}^{e}E} of maximal dimension. See example 5.1. Instead, the strategy is to use the structure of the self-affine set and the weak tangents obtained in Section 3 to show that WViseT¯W\cap\overline{\operatorname{Vis}^{e}T} can be covered by graphs of few well behaving functions and that WViseT¯W\setminus\overline{\operatorname{Vis}^{e}T} can be covered by a countable collection of lines. The arguments about visibility rely heavily on the projection condition.

The visibility conjecture asks if dimHVisθ(E)=1\operatorname{dim_{H}}\operatorname{Vis}^{\theta}(E)=1 for almost all eS1e\in S^{1}. So, the remaining step to confirm the conjecture in some special case, is to show that ϑ1(Σ)\vartheta_{1}(\Sigma) is of measure zero (or technically, that the set e{eϑ1(Σ)}e\in\{\langle e\rangle\in\vartheta_{1}(\Sigma)\} is of measure zero). Theorem 4.4 deals with this in the case where the self-affine system also satisfies the strong cone separation. In section 5, I give an example where the visible part has large dimension in directions ϑ1(Σ)\vartheta_{1}(\Sigma), showing that Theorem 2.4 is sharp.

3. Weak tangents of dominated self-affine sets

This section deals with the structure of the weak tangent sets of self-affine sets satisfying the projection condition and dominated splitting. Recall that no separation conditions are required. The structure of tangents of self-affine sets under separation conditions has been studied in [2, 10, 15, 11] for example.

To study the local structure of self-affine sets it is convenient to approximate the cylinders φ𝚒(E)\varphi_{\mathtt{i}}(E) by rectangles. The domination ensures that α2(𝚒)/α1(𝚒)0\alpha_{2}(\mathtt{i})/\alpha_{1}(\mathtt{i})\to 0 uniformly as |𝚒||\mathtt{i}|\to\infty. Therefore the approximation of φ𝚒(E)\varphi_{\mathtt{i}}(E) can be done with a “very narrow” rectangle if |𝚒||\mathtt{i}| is large. This motivates the following definition.

Definition 3.1.

For 𝚒Σ\mathtt{i}\in\Sigma^{*} define the approximating rectangle R(x,r,𝚒)R(x,r,\mathtt{i}) to be the smallest closed rectangle that includes Mx,r(E𝚒)M_{x,r}(E_{\mathtt{i}}) and has sides parallel to ϑ1(𝚒)\vartheta_{1}(\mathtt{i}) and ϑ2(𝚒)\vartheta_{2}(\mathtt{i}). For any approximating rectangle RR, the length of the sides parallel to ϑ1(𝚒)\vartheta_{1}(\mathtt{i}) is denoted by h(R)h(R) and the length of the sides parallel to ϑ2(𝚒)\vartheta_{2}(\mathtt{i}) is denoted by v(R)v(R). The orientation of cylinder E𝚒E_{\mathtt{i}} is also called the orientation of the approximating rectangle R(x,r,𝚒)R(x,r,\mathtt{i}).

Lemma 3.2.

Let EE be a self-affine set satisfying the projection condition and the dominated splitting. Let WW be a weak tangent of EE with Mxn,rn(E)WM_{x_{n},r_{n}}(E)\to W in B(0,1)B(0,1) and let xWx\in W.

Then there exists a sequence 𝚒nΣ\mathtt{i}_{n}\in\Sigma^{*} of finite words and a sequence Rn:=Rn(xn,rn,𝚒n)R_{n}:=R_{n}(x_{n},r_{n},\mathtt{i}_{n}) of approximating rectangles so that hn:=h(Rn)h_{n}:=h(R_{n})\to\infty and vn:=v(Rn)0v_{n}:=v(R_{n})\to 0 and dist(Rn,x)0\operatorname{dist}(R_{n},x)\to 0.

Proof.

Since xWx\in W, there exists sequences (zn)B(0,1)(z_{n})\subset B(0,1) and (𝚒n)Σ(\mathtt{i}_{n})\subset\Sigma^{*} so that znMxn,rn(E𝚒n)z_{n}\in M_{x_{n},r_{n}}(E_{\mathtt{i}_{n}}) and znxz_{n}\to x. Furthermore, 𝚒n\mathtt{i}_{n} can be chosen so that α2(𝚒n)rn/n\alpha_{2}(\mathtt{i}_{n})\approx r_{n}/n. By setting Rn=R(xn,rn,𝚒n)R_{n}=R(x_{n},r_{n},\mathtt{i}_{n}) it is obvious that dist(Rn,x)0\operatorname{dist}(R_{n},x)\to 0. Note also that α1(𝚒n)rnhn\alpha_{1}(\mathtt{i}_{n})\approx r_{n}h_{n} and α2(𝚒n)rnvn\alpha_{2}(\mathtt{i}_{n})\approx r_{n}v_{n}. By domination, there exist τ>1\tau>1, so that α1(𝚒n)τnα2(𝚒n)\alpha_{1}(\mathtt{i}_{n})\geq\tau^{n}\alpha_{2}(\mathtt{i}_{n}) for large nn. Thus,

vnα2(𝚒n)/rn1/n0v_{n}\approx\alpha_{2}(\mathtt{i}_{n})/r_{n}\approx 1/n\to 0

and

hnα1(𝚒n)/rnτnα2(𝚒n)/rnτn/n.h_{n}\approx\alpha_{1}(\mathtt{i}_{n})/r_{n}\geq\tau^{n}\alpha_{2}(\mathtt{i}_{n})/r_{n}\approx\tau^{n}/n\to\infty.

After the previous lemma, it is intuitive that the weak tangent contains lines and half lines pointing in different directions. Due to the obvious connection, it is natural to call such sets Kakeya type sets.

Definition 3.3.

Let X2X\subset\mathbb{R}^{2} and fix θxS1\theta_{x}\in S^{1} for all xXx\in X. A set of the form

xX{x}+(θx)\bigcup_{x\in X}\{x\}+\ell(\theta_{x})

is called a Kakeya type set. The collection {θx}xX\{\theta_{x}\}_{x\in X} is called the direction set of the Kakeya type set.

Proposition 3.4.

Let EE be a self-affine set satisfying the projection condition and the dominated splitting, and let WW be a weak tangent of EE. Then W=DB(0,1)W=D\cap B(0,1), where DD is a Kakeya type set with direction set Λ\Lambda, that satisfies θϑ1(Σ)\langle\theta\rangle\in\vartheta_{1}(\Sigma) for all θΛ\theta\in\Lambda.

Proof.

Let Mxn,rnEM_{x_{n},r_{n}}E converge to WW in B(0,1)B(0,1) and fix xWx\in W. By Lemma 3.2, there is a sequence Rn=R(xn,rn,𝚒n)R_{n}=R(x_{n},r_{n},\mathtt{i}_{n}) of approximating rectangles with hnh_{n}\to\infty and vn0v_{n}\to 0 so that dist(x,Rn)0\operatorname{dist}(x,R_{n})\to 0. Recall that RnR_{n} has orientation ϑ1(𝚒n)\vartheta_{1}(\mathtt{i}_{n}). Since hnh_{n}\to\infty, at least one of the shorter sides of RnR_{n} is outside B(0,1)B(0,1). One can now fix a direction that points to this side that is far away. To put this precise, choose a short side of RnR_{n} that does not meet B(0,1)B(0,1) and extend this line segment to an infinite line from both ends and call this line ξ\xi for now (if there are two choices for the short side, then it does not matter which one is chosen). Note also that ξ\xi is perpendicular to ϑ1(𝚒n)\vartheta_{1}(\mathtt{i}_{n}). Then choose θn\theta_{n} so that θn=ϑ1(𝚒n)\langle\theta_{n}\rangle=\vartheta_{1}(\mathtt{i}_{n}) and {x}+tθn\{x\}+t\theta_{n} meets ξ\xi for some t>0t>0. By passing to a sub-sequence, one can also assume that θn\theta_{n} converges to some θxS1\theta_{x}\in S^{1} and by Lemma 2.1 part (3) it holds that θxϑ1(Σ)\langle\theta_{x}\rangle\in\vartheta_{1}(\Sigma). By the projection condition there is at least one eS1ϑ1(Σ)e\in S^{1}\setminus\vartheta_{1}(\Sigma) so that projeφ𝚒(E)\operatorname{proj}^{e}\varphi_{\mathtt{i}}(E) is an interval whenever |𝚒||\mathtt{i}| is large. Further by compactness of ϑ1(Σ)\vartheta_{1}(\Sigma), the approximating rectangles RnR_{n}, with large nn, have orientation bounded away from e\langle e\rangle. Thus, by the projection condition and the choices made above, it is clear that

({x}+(θx))B(0,1)W.\left(\{x\}+\ell(\theta_{x})\right)\cap B(0,1)\subset W.

Trivially WxW{x}W\subset\bigcup_{x\in W}\{x\}. So, by taking union over all xWx\in W, it then follows that

W=xW({x}+(θy))B(0,1),W=\bigcup_{x\in W}\left(\{x\}+\ell(\theta_{y})\right)\cap B(0,1), (3.1)

which is exactly what was claimed. ∎

Remark 3.5.

For sure, the union in (3.1) is not optimal, meaning that it is not necessary to take the union over all xWx\in W. In particular, if xWx\in W then also zt=x+t(θx)Wz_{t}=x+t\ell(\theta_{x})\in W for all small t>0t>0 at least. If (θzt)=(θx)\ell(\theta_{z_{t}})=\ell(\theta_{x}) then the union doesn’t need to be over ztz_{t} at all. Note however that even tough ztz_{t} is on a line {x}+(θx)\{x\}+\ell(\theta_{x}) it may be that (θzt)(θx)\ell(\theta_{z_{t}})\neq\ell(\theta_{x}) due to overlap of cylinders in the original self-affine set EE.

4. Proofs of the main results

In this section I finish the proof of Theorem 2.4. As discussed earlier, all that is left to do is to prove Proposition 4.1. After this it is time to focus on the Corollaries 4.3, 4.4, and 4.5, that deal with the size of the exceptional set of directions, verifying the visibility conjecture in different special cases.

Proposition 4.1.

Let EE be a self-affine set satisfying the projection condition and the dominated splitting and let eS1e\in S^{1} so that eϑ1(Σ)\langle e\rangle\not\in\vartheta_{1}(\Sigma). Then dimHW1\operatorname{dim_{H}}W\leq 1 for all WTan(Vise(E)¯)W\in\operatorname{Tan}(\overline{\operatorname{Vis}^{e}(E)}).

As mentioned earlier the strategy is to cover WW with graphs of nice functions and a collection of vertical lines. With this in mind, recall some basic facts. For f:f\colon\mathbb{R}\to\mathbb{R}, let 𝒢(f)\mathcal{G}(f) denote the graph of ff. That is, 𝒢(f)={(x,y)2:f(x)=y}\mathcal{G}(f)=\{(x,y)\in\mathbb{R}^{2}:f(x)=y\}. A function f:f\colon\mathbb{R}\to\mathbb{R} satisfying f(t)f(s)L(ts)f(t)-f(s)\leq L(t-s) for some L>0L>0 and for all tst\geq s is called semi-decreasing. Also, ff is said to be semi-increasing if f-f is semi-decreasing and ff is called semi-monotone if it is semi-decreasing or semi-increasing. The aim is to use graphs of semi-monotone functions for the coverings, so the first thing to do is to check that their graphs are nice enough.

Lemma 4.2.

Let f:f\colon\mathbb{R}\to\mathbb{R} be semi-monotone. Then dimH𝒢(f)=1\operatorname{dim_{H}}\mathcal{G}(f)=1. Further the set of discontinuity points of ff is at most countable.

Proof.

It is standard that the claim holds for monotone functions. By symmetry, it is enough to show the semi-decreasing case. So, assume that ff is semi-decreasing and that the involved constant is LL. Define φ:\varphi\colon\mathbb{R}\to\mathbb{R} by φ(t)=Lt\varphi(t)=L\cdot t and consider g=fφg=f-\varphi. Since gg is monotone and φ\varphi is Lipschitz, the second claim follows. Also, dimH𝒢(g)=1\operatorname{dim_{H}}\mathcal{G}(g)=1 since gg is monotone. On the other hand, 𝒢(g)=Ψ(𝒢(f))\mathcal{G}(g)=\Psi(\mathcal{G}(f)), where Ψ:22\Psi\colon\mathbb{R}^{2}\to\mathbb{R}^{2} is defined by Ψ(x,y)=(x,yφ(x))\Psi(x,y)=(x,y-\varphi(x)). Clearly 1dimH𝒢(f)1\leq\operatorname{dim_{H}}\mathcal{G}(f), since projπ2𝒢(f)=\operatorname{proj}^{-\frac{\pi}{2}}\mathcal{G}(f)=\mathbb{R}. On the other hand, it is easy to see that Ψ\Psi is bi-Lipschitz, so 1dimH𝒢(f)=dimHΨ(𝒢(f))=dimH𝒢(g)=11\leq\operatorname{dim_{H}}\mathcal{G}(f)=\operatorname{dim_{H}}\Psi(\mathcal{G}(f))=\operatorname{dim_{H}}\mathcal{G}(g)=1

Proof of Proposition 4.1.

Fix a direction ee as in the claim and let xn,rnx_{n},r_{n} be sequences so that Mxn,rn(Vise(E)¯)WM_{x_{n},r_{n}}(\overline{\operatorname{Vis}^{e}(E)})\to W in B(0,1)B(0,1). After passing to subsequence if necessary, it can also be assumed that Mxn,rn(E)M_{x_{n},r_{n}}(E) converges to some weak tangent TT in B(0,1)B(0,1). By Proposition 3.4, the weak tangent TT is a Kakeya type set, so let XTX\subset T and {θx}xXS1\{\theta_{x}\}_{x\in X}\subset S^{1}, so that T=xX({x}+(θx))T=\bigcup_{x\in X}(\{x\}+\ell(\theta_{x})). By assumption ±eθx\pm e\neq\theta_{x} for all xXx\in X. Without loss of generality, assume that e=(0,1)e=(0,-1). Let β=min{|(θx,±π/2)|}\beta=\min\{|\sphericalangle(\theta_{x},\pm\pi/2)|\} and θ=π/2β\theta=\pi/2-\beta. Compactness of ϑ1(Σ)\vartheta_{1}(\Sigma) ensures that β\beta is strictly positive. Still, without loss of generality, assume that dimHW=dimHWB(0,21cosθ)\operatorname{dim_{H}}W=\operatorname{dim_{H}}W\cap B(0,2^{-1}\cos\theta), so it suffices to estimate the dimension of W=WB(0,21cosθ)W^{\prime}=W\cap B(0,2^{-1}\cos\theta). Set γ:=21cosθ\gamma:=2^{-1}\cos\theta. The reason of focusing on this smaller ball inside B(0,1)B(0,1), is merely a technicality and there is no need for the reader to worry about this too much. In a nutshell, if \ell is a line or half line that meets B(0,1)B(0,1), then (proje)[1,1](\operatorname{proj}^{e}\ell)\cap[-1,1] may be different from proje(B(0,1))\operatorname{proj}^{e}(\ell\cap B(0,1)). The choice of γ\gamma ensures that if \ell is a line or half line included in the weak tangent, and it meets B(0,γ)B(0,\gamma), then (proje)[γ,γ](\operatorname{proj}^{e}\ell)\cap[-\gamma,\gamma] equals to [γ,γ]proje(B(0,1))[-\gamma,\gamma]\cap\operatorname{proj}^{e}(\ell\cap B(0,1)).

First divide TT into three sets that each have nice enough geometry. Recall that TT consists of line segments, and only the lines that hit B(0,γ)B(0,\gamma) are meaningful. If LL is a collection of lines and half lines so that (L)B(0,1)=T\left(\bigcup_{\ell\in L}\ell\right)\cap B(0,1)=T, and LLL^{\prime}\subset L consists of those elements that meet B(0,γ)B(0,\gamma), then set

LT\displaystyle L_{T} ={L:(B(0,1))=2}\displaystyle=\{\ell\in L^{\prime}:\sharp(\ell\cap\partial B(0,1))=2\}
LR\displaystyle L_{R} ={L:={x}+(θx), with cosθx>0, and |x|<1}\displaystyle=\{\ell\in L^{\prime}:\ell=\{x\}+\ell(\theta_{x}),\text{ with }\cos\theta_{x}>0,\text{ and }|x|<1\}
LL\displaystyle L_{L} ={L:={x}+(θx), with cosθx<0, and |x|<1}\displaystyle=\{\ell\in L^{\prime}:\ell=\{x\}+\ell(\theta_{x}),\text{ with }\cos\theta_{x}<0,\text{ and }|x|<1\}

For the lines in LTL_{T} there are two possibilities. According to Lemma 3.2, for LT\ell\in L_{T}, it may be that there exists a sequence of approximating rectangles RnR_{n} converging to \ell in B(0,1)B(0,1). In this case set LTT\ell\in L_{TT}. If this is not the case, then, since every approximating rectangle can have at most one short side meeting B(0,1)B(0,1), there are two sequences of approximating rectangles, say RnR_{n} and SnS_{n}, so that RnSnR_{n}\cup S_{n}\to\ell in B(0,1)B(0,1). (There might be many more that could be chosen, but it is enough to consider these two.) Since a small neighborhood of the vertical orientation is excluded, it makes sense to talk about left and right sides of these rectangles, referring to the shorter sides that are most right and most left. Assume that the left side of RnR_{n} does not meet B(0,1)B(0,1) and the right side of SnS_{n} does not meet B(0,1)B(0,1). By passing to subsequences, one can assume that there are x,zx,z\in\ell so that RnR_{n} converges to x+(θx)x+\ell(\theta_{x}) in B(0,1)B(0,1), with cosθx<0\cos\theta_{x}<0 and that SnS_{n} converges to z+(θz)z+\ell(\theta_{z}) in B(0,1)B(0,1), with cosθz>0\cos\theta_{z}>0. In this case, set x+(θx)TTLx+\ell(\theta_{x})\in T_{TL} and z+(θz)TTRz+\ell(\theta_{z})\in T_{TR}.

Finally, set

TT=LTT¯,TR=LRLTR¯,TL=LLLTL¯,T_{T}=\overline{\bigcup_{\ell\in L_{TT}}\ell},\qquad T_{R}=\overline{\bigcup_{\ell\in L_{R}\cup L_{TR}}\ell},\qquad T_{L}=\overline{\bigcup_{\ell\in L_{L}\cup L_{TL}}\ell},

and T=TTTLTRT^{\prime}=T_{T}\cup T_{L}\cup T_{R}. The closures are taken to ensure that TiB(0,1)T_{i}\cap B(0,1) is a compact set for all i{T,R,L}i\in\{T,R,L\}. Obviously TB(0,γ)=TB(0,γ)T\cap B(0,\gamma)=T^{\prime}\cap B(0,\gamma). Note that despite the closures, the sets TiT_{i} are still collections of lines, since the closure of any set of lines in 2\mathbb{R}^{2} is still a set of lines in 2\mathbb{R}^{2}. Most importantly, for each line in TiT_{i}, there still exists a sequence RnR_{n} of approximating rectangles with RnTiR_{n}\to T_{i}, since they existed for all lines of LiL_{i}.

Now it is time to estimate dimH(ViseT)\operatorname{dim_{H}}(\operatorname{Vis}^{e}T^{\prime}). Trivially, ViseTViseTTViseTLViseTR\operatorname{Vis}^{e}T^{\prime}\subset\operatorname{Vis}^{e}T_{T}\cup\operatorname{Vis}^{e}T_{L}\cup\operatorname{Vis}^{e}T^{\prime}_{R} so it suffices to consider ViseTi\operatorname{Vis}^{e}T_{i} for i{T,L,R}i\in\{T,L,R\} separately. Of course, some of the sets TiT_{i} may be empty, but at least one of them is nonempty since TB(0,γ)T\cap B(0,\gamma) is nonempty.

Start with TRT_{R}. Note that [γ,γ]projeTR=[u,γ]=:IR[-\gamma,\gamma]\cap\operatorname{proj}^{e}T_{R}=[u,\gamma]=:I_{R} for some uu. Consider the function fR:IRf_{R}\colon I_{R}\to\mathbb{R} defined by f(x)=min{y:(x,y)TR}f(x)=\min\{y:(x,y)\in T_{R}\}. Let Γ\Gamma denote the strip [γ,γ]×[-\gamma,\gamma]\times\mathbb{R}. Obviously 𝒢(fR)=Vise(TRΓ)\mathcal{G}(f_{R})=\operatorname{Vis}^{e}(T_{R}\cap\Gamma). Consider s,tIRs,t\in I_{R}, with s<ts<t. Since (s,f(s))(s,f(s)) is on a line segment LR\ell\in L_{R}, with t[s,γ]projet\in[s,\gamma]\subset\operatorname{proj}^{e}\ell, it is clear that f(t)f(s)tanθ(ts)f(t)-f(s)\leq\tan\theta(t-s), so fRf_{R} is semi-decreasing, and dimHG(fR)1\operatorname{dim_{H}}G(f_{R})\leq 1 by Lemma 4.2.

Similarly, define fL:ILf_{L}\colon I_{L}\to\mathbb{R} by setting f(x)=min{y:(x,y)TL}f(x)=\min\{y:(x,y)\in T_{L}\}. Again, 𝒢(fL)=Vise(TLΓ)\mathcal{G}(f_{L})=\operatorname{Vis}^{e}(T_{L}\cap\Gamma). Consider s,tILs,t\in I_{L}, with s<ts<t. Since (t,f(t))(t,f(t)) is on a line segment LL\ell\in L_{L}, with s[γ,t]projes\in[-\gamma,t]\subset\operatorname{proj}^{e}\ell, it is clear that f(t)f(s)tanθ(ts)f(t)-f(s)\geq-\tan\theta(t-s). Thus fLf_{L} is semi-increasing, and dimHG(fL)1\operatorname{dim_{H}}G(f_{L})\leq 1 by Lemma 4.2.

Finally, define fT:[γ,γ]f_{T}\colon[-\gamma,\gamma]\to\mathbb{R} by f(x)=min{y:(x,y)TT}f(x)=\min\{y:(x,y)\in T_{T}\} and note that fTf_{T} is Lipschitz. All in all, the above considerations show that dimHWViseT1\operatorname{dim_{H}}W^{\prime}\cap\operatorname{Vis}^{e}T^{\prime}\leq 1.

Then it is time to estimate dimH(WViseT)\operatorname{dim_{H}}(W^{\prime}\setminus\operatorname{Vis}^{e}T^{\prime}). The aim is to show that WViseTW^{\prime}\setminus\operatorname{Vis}^{e}T^{\prime} can be covered by a countable collection of vertical line segments. More specifically, by two collections of vertical lines that are parametrized by a) the discontinuity pints of fif_{i}, b) the boundary points of IiI_{i}. Considering the first case, Lemma 4.2 showed that a semi-monotone function can have only countably many points of discontinuity. Hence, set

LD:=i=L,R,T{{s}×:fi is discontinuous at s}.L_{D}:=\bigcup_{i=L,R,T}\bigcup\{\{s\}\times\mathbb{R}:f_{i}\text{ is discontinuous at }s\}.

For the second case, let {ti}i=16\{t_{i}\}_{i=1}^{6} be the endpoints of the intervals Ik,k=T,R,LI_{k},k=T,R,L, and set

LB:=i{ti}×L_{B}:=\bigcup_{i}\{t_{i}\}\times\mathbb{R}

The final step is to show that WViseTLDLBW^{\prime}\setminus\operatorname{Vis}^{e}T^{\prime}\subset L_{D}\cup L_{B}. If this is not the case, then there exists a point ω=(ω1,ω2)W(ViseTLDLB)\omega=(\omega_{1},\omega_{2})\in W^{\prime}\setminus(\operatorname{Vis}^{e}T^{\prime}\cup L_{D}\cup L_{B}). Since projeWITIRIL\operatorname{proj}^{e}W^{\prime}\subset I_{T}\cup I_{R}\cup I_{L} and ω\omega is not visible, there exists x=(ω1,x2)ViseTx=(\omega_{1},x_{2})\in\operatorname{Vis}^{e}T^{\prime}\cap\ell with x2<w2x_{2}<w_{2} and Ti\ell\in T_{i} for some i=T,R,Li=T,R,L.

Assume first that i=Ti=T. Then, by the choices made above, there is a sequence RnR_{n} of approximating rectangles, with side lengths hnh_{n}\to\infty and vn0v_{n}\to 0, converging to \ell in B(0,1)B(0,1). Let ωn\omega^{n} be a sequence so that ωnω\omega^{n}\to\omega and ωnMxn,rn(ViseE)\omega^{n}\in M_{x_{n},r_{n}}(\operatorname{Vis}^{e}E). (Recall that Mxn,rn(ViseE¯)B(0,1)WM_{x_{n},r_{n}}(\overline{\operatorname{Vis}^{e}E})\cap B(0,1)\to W.) Let nn be so large that dist(Rn,x)<(ω2nx2)/2\operatorname{dist}(R_{n},x)<(\omega^{n}_{2}-x_{2})/2 and vn/cosθ<(ω2nx2)/2v_{n}/\cos\theta<(\omega^{n}_{2}-x_{2})/2 for all large nn. This is possible since in both inequalities the left hand side converges to zero and (ω2nx2)(\omega^{n}_{2}-x_{2}) converges to (ω2x2)>0(\omega_{2}-x_{2})>0. Now, by the projection condition, there exists a point znRnMxn,rn(E)z_{n}\in R_{n}\cap M_{x_{n},r_{n}}(E) so that zn{ωn}+(π/2)z_{n}\in\{\omega^{n}\}+\ell(-\pi/2) implying that ωnMxn,rn(ViseE)\omega^{n}\not\in M_{x_{n},r_{n}}(\operatorname{Vis}^{e}E), which is a contradiction.

Then assume that i=Ri=R. Assume also that ω1int(IR)\omega_{1}\in\operatorname{int}(I_{R}) and that fRf_{R} is continuous at ω1\omega_{1}, since otherwise ω\omega is covered by LDL_{D} or LBL_{B}. Thus there exists z=(z1,z2)ViseTRz=(z_{1},z_{2})\in\operatorname{Vis}^{e}T_{R} with z1<w1z_{1}<w_{1} and z2=fR(z1)z_{2}=f_{R}(z_{1}) and z2+|z1x1|tanθ<(ω2x2)/4z_{2}+|z_{1}-x_{1}|\tan\theta<(\omega_{2}-x_{2})/4. Again, there is a sequence RnR_{n} of approximating rectangles and points znRnz^{n}\in R_{n} with znzz^{n}\to z. Let ωn\omega^{n} be a sequence so that wnωw^{n}\to\omega and ωnMxn,rn(ViseE¯)\omega^{n}\in M_{x_{n},r_{n}}(\overline{\operatorname{Vis}^{e}E}). When nn is so large that

z2n+|z1nx1|tanθ<x2+(ω2x2)38,vncos(θ)1<|ω2x2|/8,ω2n|ω1nω1|tanθ>ω2|ω2x2|/4,\begin{split}&z_{2}^{n}+|z^{n}_{1}-x_{1}|\tan\theta<x_{2}+(\omega_{2}-x_{2})\frac{3}{8},\\ &v_{n}\cos(\theta)^{-1}<|\omega_{2}-x_{2}|/8,\\ &\omega^{n}_{2}-|\omega^{n}_{1}-\omega_{1}|\tan\theta>\omega_{2}-|\omega_{2}-x_{2}|/4,\end{split} (4.1)

the projection condition implies that {ωn}+(π/2)Rn\{\omega^{n}\}+\ell(-\pi/2)\cap R_{n}\neq\emptyset for all large nn implying that ωW\omega\not\in W^{\prime}. See Figure 1 for clarification. The case i=Li=L is symmetric to the case i=Ri=R.

ω\omegaωn\omega^{n}xxzzRnR_{n}znz^{n}
Figure 1. This picture explains the formulas (4.1). When the approximating rectangle RnR_{n} is narrow, and ωn\omega^{n} is near ω\omega, and znz^{n} is near zz, the projection condition ensures that ωn\omega^{n} is not visible.

The conclusion now is that W𝒢(fT)𝒢(fR)𝒢(fL)LDLBW^{\prime}\subset\mathcal{G}(f_{T})\cup\mathcal{G}(f_{R})\cup\mathcal{G}(f_{L})\cup L_{D}\cup L_{B} and each element in the union has Hausdorff dimension 11, so the proof is finished. ∎

Considering the visibility conjecture, there is still the question whether 1(ϑ1(Σ))=0\mathcal{H}^{1}(\vartheta_{1}(\Sigma))=0. For dominated self-affine carpets, this is true. A self-affine set is called a carpet if all AiA_{i} are diagonal matrixes. If a carpet is dominated, then ϑ1(Σ)\vartheta_{1}(\Sigma) is a singleton - it is either the horizontal or the vertical orientation. Thus theorem 2.4 immediately implies that the visibility conjecture holds in this class.

Corollary 4.3.

Let EE be a self-affine carpet satisfying the projection condition and the dominated splitting. Then dimHVise(E)=dimAVise(E)=1\operatorname{dim_{H}}\operatorname{Vis}^{e}(E)=\operatorname{dim_{A}}\operatorname{Vis}^{e}(E)=1 holds for all except possibly one eS1e\in S^{1} and its opposite e-e.

To verify the visibility conjecture in a more general setting, consider the self-affine sets satisfying the “strong cone separation” introduced in [13]: assume that there is a cone X2X\subset\mathbb{R}^{2} so that Ai(X)int(X)A_{i}(X)\subset\operatorname{int}(X) and AiT(X)int(X)A_{i}^{T}(X)\subset\operatorname{int}(X), and for all ii and that

Ai(X)Aj(X)=A_{i}(X)\cap A_{j}(X)=\emptyset (4.2)

for all iji\neq j. As is intuitive, a cone is a union of set of lines through origin in 2\mathbb{R}^{2} that have bounded angle from some fixed line. In what follows, the cone XX is understood as a subsets of 2\mathbb{R}^{2} or 1\mathbb{P}^{1} depending on the situation, and this should not cause any confusion. So, equivalently, a cone is an interval in the projective space 1\mathbb{P}^{1}. As discussed in the proof of [13, Lemma 4.1] it follows that η1(A𝚒),ϑ1(A𝚒)X\eta_{1}(A_{\mathtt{i}}),\vartheta_{1}(A_{\mathtt{i}})\in X and η2(A𝚒),ϑ2(A𝚒)X\eta_{2}(A_{\mathtt{i}}),\vartheta_{2}(A_{\mathtt{i}})\not\in X for all 𝚒Σ\mathtt{i}\in\Sigma^{*}. Further, without loss of generality, assume that η1(𝚒)\eta_{1}(\mathtt{i}) is uniformly separated from XcX^{c} for all 𝚒\mathtt{i} independently of the length |𝚒||\mathtt{i}|. This follows simply by choosing XX^{\prime} to be the minimal cone that includes i(Ai(X)AiT(X))\cup_{i}(A_{i}(X)\cup A_{i}^{T}(X)) and then applying the previous deduction to the cones Ai(X)A_{i}(X^{\prime}). (Note that the strong cone separation holds with the cone XX^{\prime} as well.)

Corollary 4.4.

Let EE be a self-affine satisfying the projection condition and the strong cone separation. Then dimHVise(E)=dimAVise(E)=1\operatorname{dim_{H}}\operatorname{Vis}^{e}(E)=\operatorname{dim_{A}}\operatorname{Vis}^{e}(E)=1 for almost all eS1e\in S^{1}.

Proof.

The strong cone separation implies domination [3, Theorem B], so by Theorem 2.4 it is enough to show that 1(ϑ1(Σ))=0\mathcal{H}^{1}(\vartheta_{1}(\Sigma))=0. This would certainly follow form dimAϑ1(Σ)<1\operatorname{dim_{A}}\vartheta_{1}(\Sigma)<1 and this in turn follows if ϑ1(Σ)\vartheta_{1}(\Sigma) is porous. Recall that a subset YY of some metric space is porous if there are constants r0,α>0r_{0},\alpha>0 so that for all yYy\in Y and r<r0r<r_{0}, there is xB(y,(1α)r)x\in B(y,(1-\alpha)r) so that B(x,αr)Y=B(x,\alpha r)\cap Y=\emptyset. For the connection between Assouad dimension and porosity, see for example [14].

Let XX be the cone from (4.2). Since XX is a union of lines through origin, it can also be considered as a subset of the projective space. Recall that 1\mathbb{P}^{1} is a metric space where the distance is measured by (,)\sphericalangle(\cdot,\cdot), the angle between the corresponding lines in 2\mathbb{R}^{2} (which lies in the interval [0,π/2][0,\pi/2] as usual). An invertible linear mapping A:22A\colon\mathbb{R}^{2}\to\mathbb{R}^{2} can naturally be interpreted as a mapping on the projective space as well since A(v)=A(v)A(\langle v\rangle)=\langle A(v)\rangle by linearity. Thus the mappings {Ai}i=1κ\{A_{i}\}_{i=1}^{\kappa} naturally generate collections of nested compact sets inside XX, since Ai(X)XA_{i}(X)\subset X for all ii. To be more precise, I claim that setting E𝚒=A𝚒(X)E_{\mathtt{i}}=A_{\mathtt{i}}(X), the collection {E𝚒}𝚒Σ\{E_{\mathtt{i}}\}_{\mathtt{i}\in\Sigma^{*}} satisfies the conditions

  1. (M1)

    E𝚒nE𝚒|n1E_{\mathtt{i}_{n}}\subset E_{\mathtt{i}|_{n-1}} for all 𝚒Σ\mathtt{i}\in\Sigma and nn\in\mathbb{N}

  2. (M2)

    diam(E𝚒|n)0\operatorname{diam}(E_{\mathtt{i}|_{n}})\to 0 as nn\to\infty for all 𝚒Σ\mathtt{i}\in\Sigma.

Part (M1) is obviously true, but verifying part (M2) requires some work. If these conditions hold, then the collection {E𝚒}𝚒Σ\{E_{\mathtt{i}}\}_{\mathtt{i}\in\Sigma^{*}} is called a Moran construction and there is a unique compact set

Y=𝚒ΣnE𝚒|n.Y=\bigcup_{\mathtt{i}\in\Sigma}\bigcap_{n\in\mathbb{N}}E_{\mathtt{i}|_{n}}.

Further knowledge about Moran constructions is not needed, but the interested reader can check for example[12]. For this particular example, the key point is that by combining (M1), (M2), with Lemma 2.1 parts (3) and (4), and the fact that η1(A𝚒)X\eta_{1}(A_{\mathtt{i}})\in X for all 𝚒Σ\mathtt{i}\in\Sigma, it follows that YY equals to ϑ1(Σ)\vartheta_{1}(\Sigma).

Now, to see that also (M2) holds, fix 𝚒Σn\mathtt{i}\in\Sigma^{n}, and a,bXa,b\in X. Further, fix unit vectors tη1(𝚒)+sη2(𝚒)t\eta_{1}(\mathtt{i})+s\eta_{2}(\mathtt{i}) and uη1(𝚒)+vη2(𝚒)u\eta_{1}(\mathtt{i})+v\eta_{2}(\mathtt{i}) so that

a=tη1(𝚒)+sη2(𝚒)andb=uη1(𝚒)+vη2(𝚒).a=\langle t\eta_{1}(\mathtt{i})+s\eta_{2}(\mathtt{i})\rangle\quad\text{and}\quad b=\langle u\eta_{1}(\mathtt{i})+v\eta_{2}(\mathtt{i})\rangle.

It immediately follows that

A𝚒(a)\displaystyle A_{\mathtt{i}}(a) =A𝚒(tη1(𝚒)+sη2(𝚒))=tA𝚒η1(𝚒)+sA𝚒η2(𝚒)\displaystyle=\langle A_{\mathtt{i}}(t\eta_{1}(\mathtt{i})+s\eta_{2}(\mathtt{i}))\rangle=\langle tA_{\mathtt{i}}\eta_{1}(\mathtt{i})+sA_{\mathtt{i}}\eta_{2}(\mathtt{i})\rangle
A𝚒(b)\displaystyle A_{\mathtt{i}}(b) =A𝚒(uη1(𝚒)+vη2(𝚒))=uA𝚒η1(𝚒)+vA𝚒η2(𝚒)\displaystyle=\langle A_{\mathtt{i}}(u\eta_{1}(\mathtt{i})+v\eta_{2}(\mathtt{i}))\rangle=\langle uA_{\mathtt{i}}\eta_{1}(\mathtt{i})+vA_{\mathtt{i}}\eta_{2}(\mathtt{i})\rangle

Further, it is no restriction to assume that tt and uu are non-negative. Since η2(𝚒)X\langle\eta_{2}(\mathtt{i})\rangle\not\in X, there exists δ>0\delta>0, so that t,u>δt,u>\delta and |s|,|v|<1δ|s|,|v|<1-\delta. Moreover, δ\delta can be chosen to be independent of a,b,𝚒a,b,\mathtt{i} and the level nn, since η2(𝚒)\langle\eta_{2}(\mathtt{i})\rangle is uniformly separated from XX. Since δ>0\delta>0 is fixed, there exists M>1M>1 so that

|γβ||tanγtanβ|M|γβ||\gamma-\beta|\leq|\tan\gamma-\tan\beta|\leq M|\gamma-\beta| (4.3)

for angles γ,β[π/2+δ/2,π/2δ/2]\gamma,\beta\in[-\pi/2+\delta/2,\pi/2-\delta/2] and MδπδM\delta\geq\pi-\delta. Therefore, it follows that

(A𝚒(a),A𝚒(b))A𝚒η2(𝚒)A𝚒η1(𝚒)|stvu|=α2(𝚒)α1(𝚒)|stvu|α2(𝚒)α1(𝚒)MπτnMπ,\sphericalangle(A_{\mathtt{i}}(a),A_{\mathtt{i}}(b))\leq\frac{\|A_{\mathtt{i}}\eta_{2}(\mathtt{i})\|}{\|A_{\mathtt{i}}\eta_{1}(\mathtt{i})\|}\left|\frac{s}{t}-\frac{v}{u}\right|=\frac{\alpha_{2}(\mathtt{i})}{\alpha_{1}(\mathtt{i})}\left|\frac{s}{t}-\frac{v}{u}\right|\leq\frac{\alpha_{2}(\mathtt{i})}{\alpha_{1}(\mathtt{i})}M\pi\leq\tau^{-n}M\pi, (4.4)

and τn0\tau^{-n}\to 0 as nn\to\infty, which proves (M2).

It now suffices to show that {E𝚒}𝚒Σ\{E_{\mathtt{i}}\}_{\mathtt{i}\in\Sigma^{*}}, satisfies the following bounded distortion property: there are constants k0k_{0}\in\mathbb{N} and D>1D>1 so that

d(𝚒)d(𝚒)D\frac{d^{*}(\mathtt{i})}{d_{*}(\mathtt{i})}\leq D

for all 𝚒Σk\mathtt{i}\in\Sigma^{k} and kk0k\geq k_{0}, where

d(𝚒)=supa,bXab(A𝚒(a),A𝚒(b))(a,b) and d(𝚒)=infa,bXab(A𝚒(a),A𝚒(b))(a,b)d^{*}(\mathtt{i})=\sup_{\genfrac{}{}{0.0pt}{}{a,b\in X}{a\neq b}}\frac{\sphericalangle(A_{\mathtt{i}}(a),A_{\mathtt{i}}(b))}{\sphericalangle(a,b)}\qquad\text{ and }\qquad d_{*}(\mathtt{i})=\inf_{\genfrac{}{}{0.0pt}{}{a,b\in X}{a\neq b}}\frac{\sphericalangle(A_{\mathtt{i}}(a),A_{\mathtt{i}}(b))}{\sphericalangle(a,b)}

If the bounded distortion holds, then let II be a gap between two neighboring first level cylinders Ai(X)A_{i}(X) and Aj(X)A_{j}(X). Note that II exists due to the strong cone separation. Let r>0r>0 and θY\theta\in Y. Let 𝚒\mathtt{i} be a finite word with A𝚒(Y)B(θ,r)A_{\mathtt{i}}(Y)\subset B(\theta,r) but |A𝚒(Y)|r|A_{\mathtt{i}}(Y)|\gtrsim r. Then, due to the separation of the cones, there is a gap G:=A𝚒(I)B(θ,r)G:=A_{\mathtt{i}}(I)\subset B(\theta,r) and

|G|r\displaystyle\frac{|G|}{r} d(𝚒)|I||A𝚒(Y)|d(𝚒)|I|d(𝚒)|Y|D1|I||Y|\displaystyle\gtrsim\frac{d_{*}(\mathtt{i})|I|}{|A_{\mathtt{i}}(Y)|}\geq\frac{d_{*}(\mathtt{i})|I|}{d^{*}(\mathtt{i})|Y|}\geq D^{-1}\frac{|I|}{|Y|}

which shows that YY is porous.

To show the bounded distortion property, fix 𝚒Σk\mathtt{i}\in\Sigma^{k}, a,bYa,b\in Y. There is a small annoying technicality that the angle between lines in XX may be realized “outside” XX if the opening angle of XX is larger than π/2\pi/2. Therefore, let k0k_{0} be so large that (A𝚒(x),A𝚒(y))δ\sphericalangle(A_{\mathtt{i}}(x),A_{\mathtt{i}}(y))\leq\delta for all lines x,yXx,y\in X when |𝚒|k0|\mathtt{i}|\geq k_{0} and assume kk0k\geq k_{0}.

As before, fix unit vectors tη1(𝚒)+sη2(𝚒)t\eta_{1}(\mathtt{i})+s\eta_{2}(\mathtt{i}) and uη1(𝚒)+vη2(𝚒)u\eta_{1}(\mathtt{i})+v\eta_{2}(\mathtt{i}) so that

a=tη1(𝚒)+sη2(𝚒)andb=uη1(𝚒)+vη2(𝚒).a=\langle t\eta_{1}(\mathtt{i})+s\eta_{2}(\mathtt{i})\rangle\quad\text{and}\quad b=\langle u\eta_{1}(\mathtt{i})+v\eta_{2}(\mathtt{i})\rangle.

where tt and uu are positive. Next, note that st=±tan(η1(𝚒),a)\frac{s}{t}=\pm\tan\sphericalangle(\langle\eta_{1}(\mathtt{i})\rangle,a) depending on if ss is positive or negative, and that a similar formula holds for vu\frac{v}{u}. Combining this with (4.3) gives

(A𝚒(a),A𝚒(b))A𝚒η2(𝚒)A𝚒η1(𝚒)|stvu|=α2(𝚒)α1(𝚒)|stvu|α2(𝚒)α1(𝚒)M(a,b).\sphericalangle(A_{\mathtt{i}}(a),A_{\mathtt{i}}(b))\leq\frac{\|A_{\mathtt{i}}\eta_{2}(\mathtt{i})\|}{\|A_{\mathtt{i}}\eta_{1}(\mathtt{i})\|}\left|\frac{s}{t}-\frac{v}{u}\right|=\frac{\alpha_{2}(\mathtt{i})}{\alpha_{1}(\mathtt{i})}\left|\frac{s}{t}-\frac{v}{u}\right|\leq\frac{\alpha_{2}(\mathtt{i})}{\alpha_{1}(\mathtt{i})}M\sphericalangle(a,b). (4.5)

for lines a,bYa,b\in Y with (a,b)δ\sphericalangle(a,b)\leq\delta. If (a,b)>δ\sphericalangle(a,b)>\delta, then it could be that the angle is realized outside XX, and the last inequality in the above estimate may not hold. In this case recalling the choice of MM still gives

(A𝚒(a),A𝚒(b))α2(𝚒)α1(𝚒)|stvu|α2(𝚒)α1(𝚒)M(πδ)α2(𝚒)α1(𝚒)M2(a,b).\sphericalangle(A_{\mathtt{i}}(a),A_{\mathtt{i}}(b))\leq\frac{\alpha_{2}(\mathtt{i})}{\alpha_{1}(\mathtt{i})}\left|\frac{s}{t}-\frac{v}{u}\right|\leq\frac{\alpha_{2}(\mathtt{i})}{\alpha_{1}(\mathtt{i})}M(\pi-\delta)\leq\frac{\alpha_{2}(\mathtt{i})}{\alpha_{1}(\mathtt{i})}M^{2}\sphericalangle(a,b). (4.6)

Since (A𝚒(a),A𝚒(b))δ\sphericalangle(A_{\mathtt{i}}(a),A_{\mathtt{i}}(b))\leq\delta by the choice of k0k_{0}, the lower estimate can be treated as a single case. Again, relying on (4.3) gives

(A𝚒(a),A𝚒(b))1MA𝚒η2(𝚒)A𝚒η1(𝚒)|stvu|=1Mα2(𝚒)α1(𝚒)|stvu|1Mα2(𝚒)α1(𝚒)(a,b).\sphericalangle(A_{\mathtt{i}}(a),A_{\mathtt{i}}(b))\geq\frac{1}{M}\frac{\|A_{\mathtt{i}}\eta_{2}(\mathtt{i})\|}{\|A_{\mathtt{i}}\eta_{1}(\mathtt{i})\|}\left|\frac{s}{t}-\frac{v}{u}\right|=\frac{1}{M}\frac{\alpha_{2}(\mathtt{i})}{\alpha_{1}(\mathtt{i})}\left|\frac{s}{t}-\frac{v}{u}\right|\geq\frac{1}{M}\frac{\alpha_{2}(\mathtt{i})}{\alpha_{1}(\mathtt{i})}\sphericalangle(a,b). (4.7)

Combining (4.5), (4.6), and (4.7) gives

M1α2(𝚒)α1(𝚒)(a,b)(A𝚒(a),A𝚒(b))α2(𝚒)α1(𝚒)M2(a,b)M^{-1}\frac{\alpha_{2}(\mathtt{i})}{\alpha_{1}(\mathtt{i})}\sphericalangle(a,b)\leq\sphericalangle(A_{\mathtt{i}}(a),A_{\mathtt{i}}(b))\leq\frac{\alpha_{2}(\mathtt{i})}{\alpha_{1}(\mathtt{i})}M^{2}\sphericalangle(a,b) (4.8)

and this proves the bounded distortion with the constants k0k_{0} and M3M^{3}. ∎

If there are not too many cylinders pointing to the same direction then it is possible to get rid of the exceptional directions, but this only works for Hausdorff dimension, since the argument uses countable stability.

Corollary 4.5.

Let EE be a self-affine satisfying the projection condition and the dominated splitting. Assume further that the sets {π𝚒:ϑ1(𝚒)=e}\{\pi\mathtt{i}:\vartheta_{1}(\mathtt{i})=\langle e\rangle\} have Hausdorff dimension at most 11 for all eS1e\in S^{1}. Then dimHVise(E)=1\operatorname{dim_{H}}\operatorname{Vis}^{e}(E)=1 for all eS1e\in S^{1}.

Proof.

Assume that ϑ1(𝚒)=e\vartheta_{1}(\mathtt{i})=\langle e\rangle for some 𝚒Σ\mathtt{i}\in\Sigma, since otherwise the claim for ee follows from Theorem 2.4.

Divide the cylinders of EE into different classes according to the angle that the orientation of the cylinder has with e\langle e\rangle. Set I(δ,k)={𝚒Σk:(ϑ1(𝚒𝚓),e)>δ for all 𝚓Σ}I(\delta,k)=\{\mathtt{i}\in\Sigma^{k}:\sphericalangle(\vartheta_{1}(\mathtt{i}\mathtt{j}),\langle e\rangle)>\delta\text{ for all }\mathtt{j}\in\Sigma\} and

E(δ,k)=𝚒I(δ,k)φ𝚒(E)E(\delta,k)=\bigcup_{\mathtt{i}\in I(\delta,k)}\varphi_{\mathtt{i}}(E) (4.9)

From Lemma 2.1, it follows that E=kE(k1,k)FeE=\bigcup_{k\in\mathbb{N}}E(k^{-1},k)\cup F_{e}, where Fe={π𝚒:ϑ1(𝚒)=e}F_{e}=\{\pi\mathtt{i}:\vartheta_{1}(\mathtt{i})=\langle e\rangle\}. It also follows that all the elements of the union are compact sets. The sets E(k1,k)E(k^{-1},k) are not exactly self-affine but each of them is a finite union of affine images of EE. In particular, if 𝚒I(k1,k)\mathtt{i}\in I(k^{-1},k), then eϑ1(𝚒𝚓)\langle e\rangle\not\in\vartheta_{1}(\mathtt{i}\mathtt{j}) for all 𝚓Σ\mathtt{j}\in\Sigma and, by Lemma 2.1, A𝚒1eϑ1(Σ)\langle A_{\mathtt{i}}^{-1}e\rangle\not\in\vartheta_{1}(\Sigma). Thus Theorem 2.4 gives that dimHVisA𝚒1eE=1\operatorname{dim_{H}}\operatorname{Vis}^{A_{\mathtt{i}}^{-1}e}E=1. Recalling that, Viseφ𝚒(E)\operatorname{Vis}^{e}\varphi_{\mathtt{i}}(E) is an affine image of VisA𝚒1eE\operatorname{Vis}^{A_{\mathtt{i}}^{-1}e}E, and that Hausdorff dimension is countably stable finishes the proof. ∎

5. Final remarks

This final section exhibits a few examples dealing with the sharpness of Theorem 2.4.

Example 5.1.

Consider fi:[0,1]2[0,1]2f_{i}\colon[0,1]^{2}\to[0,1]^{2} for i=1,2,3i=1,2,3 with f1(x,y)=(31x,21y)f_{1}(x,y)=(3^{-1}x,2^{-1}y), f2(x,y)=(31x,21y)+(31,21)f_{2}(x,y)=(3^{-1}x,2^{-1}y)+(3^{-1},2^{-1}), and f3(x,y)=(31x,21y)+(231,0)f_{3}(x,y)=(3^{-1}x,2^{-1}y)+(2\cdot 3^{-1},0). The associated self-affine set EE is a Bedford-McMullen carpet, and it is well known that

dimHE=log2(2log32+1log32)=log2(2log32+1)\operatorname{dim_{H}}E=\log_{2}\left(2^{\log_{3}2}+1^{\log_{3}2}\right)=\log_{2}\left(2^{\log_{3}2}+1\right)

and

dimAE=log22+log32=1+log32,\operatorname{dim_{A}}E=\log_{2}2+\log_{3}2=1+\log_{3}2,

see for example [15]. In particular, 1<dimHE<dimAE1<\operatorname{dim_{H}}E<\operatorname{dim_{A}}E. It is easy to see that the system is dominated and that ϑ1(Σ)=π2\vartheta_{1}(\Sigma)=\left<\frac{\pi}{2}\right>.

To verify the projection condition, fix eS1e\in S^{1} with eπ2\langle e\rangle\neq\langle\frac{\pi}{2}\rangle. Note that A𝚒1(x,y)=(3nx,2ny)A_{\mathtt{i}}^{-1}(x,y)=(3^{n}x,2^{n}y) for all 𝚒Σn\mathtt{i}\in\Sigma^{n}, so n0n_{0}\in\mathbb{N} can be fixed so that the angle between the x-axis and A𝚒1(e)A_{\mathtt{i}}^{-1}(\langle e\rangle) is small, say smaller than π/8\pi/8, for all 𝚒Σn\mathtt{i}\in\Sigma^{n}, with nn0n\geq n_{0}. By Remark 2.3, it is now enough to show that if \ell is a line so that (,π)π/8\sphericalangle(\ell,\langle\pi\rangle)\leq\pi/8, and K\ell\cap K\neq\emptyset, where KK is the convex hull of EE, then also E\ell\cap E\neq\emptyset. So let \ell be such a line. The projection of KK to arbitrary direction is not an interval, but since the angle between \ell and xx-axis is small, it is easy tho see that φi1(K)\ell\cap\varphi_{i_{1}}(K)\neq\emptyset for some i1{1,2,3}i_{1}\in\{1,2,3\}. Since φi11\varphi_{i_{1}}^{-1}\ell has even smaller angle with the xx-axis, it follows that φi11K\varphi_{i_{1}}^{-1}\ell\cap K\neq\emptyset implies φi11φi2(K)\varphi_{i_{1}}^{-1}\ell\cap\varphi_{i_{2}}(K)\neq\emptyset for some i2{1,2,3}i_{2}\in\{1,2,3\}. Continuing in this manner yields a word 𝚒=(i1,i2,)Σ\mathtt{i}=(i_{1},i_{2},\ldots)\in\Sigma so that φ𝚒|n(K)\ell\cap\varphi_{\mathtt{i}|_{n}}(K) for all nn\in\mathbb{N}. Thus π𝚒E\pi\mathtt{i}\in\ell\cap E, and so the projection condition holds.

Since domination and projection condition are satisfied Theorem 2.4 gives that

1=dimHViseE=dimAViseE1=\operatorname{dim_{H}}\operatorname{Vis}^{e}E=\operatorname{dim_{A}}\operatorname{Vis}^{e}E

for all e±π2e\neq\pm\frac{\pi}{2}. However, it is also easy to see that, outside the tri-adic points on the xx-axis, there is no vertical alignment of points of EE. Moreover, for each the tri-adic point t[0,1]t\in[0,1], there are at most two points, say xx and yy, of EE with projπ2x=t=projπ2y\operatorname{proj}^{\frac{\pi}{2}}x=t=\operatorname{proj}^{\frac{\pi}{2}}y. Therefore, there exist a countable set HH, so that HVis±π2E=EH\cup\operatorname{Vis}^{\pm\frac{\pi}{2}}E=E. Thus 1<dimHE=dimHVis±π2E1<\operatorname{dim_{H}}E=\operatorname{dim_{H}}\operatorname{Vis}^{\pm\frac{\pi}{2}}E.

The weak tangent TT of EE that satisfies dimHT=dimAE\operatorname{dim_{H}}T=\operatorname{dim_{A}}E is obviously C×[0,1]C\times[0,1], where CC is the middle thirds Cantor set. However, Visπ2(T)=C×{0}\operatorname{Vis}^{-\frac{\pi}{2}}(T)=C\times\{0\} and so dimHVisπ2T=log32\operatorname{dim_{H}}\operatorname{Vis}^{-\frac{\pi}{2}}T=\log_{3}2. Therefore, if WW is the weak tangent of Visπ2E\operatorname{Vis}^{-\frac{\pi}{2}}E that has maximal dimension, then WVisπ2TW\neq\operatorname{Vis}^{-\frac{\pi}{2}}T.

In general the visibility conjecture is false for the Assouad and box dimensions as mentioned in the introduction. The following is a concrete counterexample.

Example 5.2.

Let A={0}{Sn1}n=1A=\{0\}\cup\{S_{n}^{-1}\}_{n=1}^{\infty}, where Sn=k=1n1/kS_{n}=\sum_{k=1}^{n}1/k. Consider firs the box dimension of just AA\subset\mathbb{R}. For δ>0\delta>0, consired the index nn for which δn:=(Sn)1(Sn+1)1\delta_{n}:=(S_{n})^{-1}-(S_{n+1})^{-1} is closest to δ\delta. Then, to cover AA with intervals of length δ\delta, it is essentially enough cover all of [0,Sn1][0,S_{n}^{-1}] and the rest can be neglected. Anyway, at least N(δ)δn1Sn1N(\delta)\approx\delta_{n}^{-1}S_{n}^{-1} intervals are needed. On the other hand,

δn=(Sn)1(Sn+1)1\displaystyle\delta_{n}=(S_{n})^{-1}-(S_{n+1})^{-1} =(n+1)1SnSn+11nSn2\displaystyle=\frac{(n+1)^{-1}}{S_{n}S_{n+1}}\approx\frac{1}{nS_{n}^{2}}

and it is an exercise to show that SnlognS_{n}\approx\log n. Thus it follows that

limδ0logN(δ)logδ\displaystyle\lim_{\delta\to 0}\frac{\log N(\delta)}{-\log\delta} =limnlog(δn1Sn1)logδn1=limn1logSnlogn+2logSn=1,\displaystyle=\lim_{n\to\infty}\frac{\log(\delta_{n}^{-1}S_{n}^{-1})}{\log\delta_{n}^{-1}}=\lim_{n\to\infty}1-\frac{\log S_{n}}{\log n+2\log S_{n}}=1,

which implies that dimBA=1\operatorname{dim_{B}}A=1. If one considers K=A×AK=A\times A, essentially the same calculation shows that dimBK=2\operatorname{dim_{B}}K=2. Because KK is countable, Vise(K)=K\operatorname{Vis}^{e}(K)=K for almost all directions, and thus dimBVise(K)=2\operatorname{dim_{B}}\operatorname{Vis}^{e}(K)=2 for almost all eS1e\in S^{1}. (For dimensions d>2d>2, one can of course consider K=AdK=A^{d}, the dd fold product of AA.)

References

  • [1] I. Arhosalo, E. Järvenpää, M. Järvenpää, M. Rams, and P. Shmerkin. Visible parts of fractal percolation. Proc. Edinb. Math. Soc. (2), 55(2):311–331, 2012.
  • [2] C. Bandt and A. Käenmäki. Local structure of self-affine sets. Ergodic Theory Dynam. Systems, 33(5):1326–1337, 2013.
  • [3] J. Bochi and N. Gourmelon. Some characterizations of domination. Math. Z., 263(1):221–231, 2009.
  • [4] M. Bond, I. Łaba, and J. Zahl. Quantitative visibility estimates for unrectifiable sets in the plane. Trans. Amer. Math. Soc., 368(8):5475–5513, 2016.
  • [5] M. Csörnyei. On the visibility of invisible sets. Ann. Acad. Sci. Fenn. Math., 25(2):417–421, 2000.
  • [6] R. O. Davies and H. Fast. Lebesgue density influences Hausdorff measure; large sets surface-like from many directions. Mathematika, 25(1):116–119, 1978.
  • [7] K. J. Falconer and J. M. Fraser. The visible part of plane self-similar sets. Proc. Amer. Math. Soc., 141(1):269–278, 2013.
  • [8] E. Järvenpää, M. Järvenpää, P. MacManus, and T. C. O’Neil. Visible parts and dimensions. Nonlinearity, 16(3):803–818, 2003.
  • [9] E. Järvenpää, M. Järvenpää, V. Suomala, and M. Wu. On dimensions of visible parts of self-similar sets with finite rotation groups. arXiv:2101.01017, 2021.
  • [10] A. Käenmäki, H. Koivusalo, and E. Rossi. Self-affine sets with fibred tangents. Ergodic Theory Dynam. Systems, 37(6):1915–1934, 2017.
  • [11] A. Käenmäki, T. Ojala, and E. Rossi. Rigidity of quasisymmetric mappings on self-affine carpets. Int. Math. Res. Not. IMRN, (12):3769–3799, 2018.
  • [12] A. Käenmäki and E. Rossi. Weak separation condition, Assouad dimension, and Furstenberg homogeneity. Ann. Acad. Sci. Fenn. Math., 41(1):465–490, 2016.
  • [13] A. Käenmäki and P. Shmerkin. Overlapping self-affine sets of Kakeya type. Ergodic Theory Dynam. Systems, 29(3):941–965, 2009.
  • [14] J. Luukkainen. Assouad dimension: antifractal metrization, porous sets, and homogeneous measures. J. Korean Math. Soc., 35(1):23–76, 1998.
  • [15] J. M. Mackay. Assouad dimension of self-affine carpets. Conform. Geom. Dyn., 15:177–187, 2011.
  • [16] J. M. Marstrand. Some fundamental geometrical properties of plane sets of fractional dimensions. Proc. Lond. Math. Soc. (3), 4:257–302, 1954.
  • [17] P. Mattila. Geometry of sets and measures in Euclidean spaces, volume 44 of Cambridge Studies in Advanced Mathematics. Cambridge University Press, Cambridge, 1995. Fractals and rectifiability.
  • [18] T. C. O’Neil. The Hausdorff dimension of visible sets of planar continua. Trans. Amer. Math. Soc., 359(11):5141–5170, 2007.
  • [19] T. Orponen. Slicing sets and measures, and the dimension of exceptional parameters. J. Geom. Anal., 24(1):47–80, 2014.
  • [20] T. Orponen. On the dimension of visible parts. J. Eur. Math. Soc.(JEMS), 2019. To appear.
  • [21] T. Orponen. On the Assouad dimension of projections. Proc. Lond. Math. Soc.(3), 122(2):317–351, 2021.
  • [22] E. Rossi and P. Shmerkin. Hölder coverings of sets of small dimension. J. Fractal Geom., 6(3):285–299, 2019.
  • [23] K. Simon and B. Solomyak. Visibility for self-similar sets of dimension one in the plane. Real Anal. Exchange, 32(1):67–78, 2006.