Visible part of dominated self-affine sets in the plane
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:
28A801. Introduction
For , let denote the half line starting from origin and propagating to direction . That is . For a compact set the visible part of in direction is the set of points in that satisfy
This set is denoted by . Let denote the orthogonal projection along the direction . (Note that may be different from , but always.) Consider the Hausdorff dimension of the visible part of a compact set . If then for almost all by Marstrand’s projection theorem [16]. Since and , it follows that for almost all . If , then still we have that for almost all , but the upper bound should no longer be optimal for most . The visibility conjecture states that for almost all . Obviously one can not hope this to hold for all directions, since a graph of a function can have dimension greater than for example. Further, an example of Davies and Fast [6] shows that is possible for a dense set of directions. This is the furthest one can go, since recently Orponen [20] showed that it is impossible to have 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 , where and . 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 . 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 , the Hausdorff dimension of the visible part from a point is strictly less than the Hausdorff dimension of the original set, and it is uniformly bounded away from , for almost all viewpoints . An other related problem is to determine when . Orponen [19] showed that if , then the set of directions for which has Hausdorff dimension at most . On the other hand, it follows from the main result of [22], that if , then holds outside a set of directions of box dimension . 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 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 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 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 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 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 and orientation is an element of the projective space , that is, the metric space of lines in that go through origin, and where the distance is measured by the angle between the lines. For a vector , let denote the corresponding element of the projective space. It is sometimes more intuitive to think as a set of angles instead of unit vectors. This justifies the use of notations , , , and for .
Let be an invertible linear map, so that is not a ball. Then is an ellipse whose semiaxes have different lengths. Let be the length of the longer one of the semiaxes and let be the length of the shorter one. Equivalently are the square roots of the eigenvalues of (ordered so that the larger is ). Also, set to be the orientation of the longer semiaxes of . That is, , where is the normalized eigenvector of associated to the eigenvalue . Likewise, set , where is the normalized eigenvector corresponding to . It is a basic fact that and .
Let be a collection of contractive invertible linear maps, let be a collection of vectors in , and let , for all . It standard that there exists a unique compact set satisfying
The set is called self-affine.
Set and . Write for even though this is abusing the notation. Let denote the length of the word . That is, whenever and , when . For , let be the longest common beginning of and and define a distance function in by setting , with the interpretation . This makes a compact metric space.
For quantities and , usually depending on , the notation means that there is a constant , that may depend on the self-affine set , so that . Further, means that and .
For , let and for the sake of brevity, write and for . Similarly, also write . The line is called the orientation of the cylinder , because the cylinder is “close” to being a line segment that is parallel to the line , at least when is large. This phenomena is examined in more detail in Proposition 3.4. As usual, let be the canonical projection defined by
From time to time, the set is also denoted by . The system is called dominated, or said to satisfy dominated splitting, if there are constants and , so that for all , with . Domination ensures the existence of the limit orientation for all symbols . The next lemma records this fact along with other useful properties of the limit orientations.
Lemma 2.1.
Let be a dominated self-affine set. Then
-
(1)
exists for all and the convergence is uniform.
-
(2)
The map is uniformly continuous.
-
(3)
contains the accumulation points of the set .
-
(4)
for all and .
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 . The proof there works for individual for which
In the setting of this paper, the domination implies the uniform bound 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 .
To prove (4) it suffices to show that converges to as , since is a diffeomorphism. Write
and for now let be a unit vector with for . Then it follows from domination that
as long as stays bounded away from zero. To show that it does, recall that . In particular,
where the left hand side is at least and the right hand side is at most
Thus the triangle inequality gives
and so the domination implies that for large . ∎
In addition to domination, a crucial assumption in this paper is the following projection condition.
Definition 2.2.
An affine IFS (or the invariant set ) satisfies the projection condition if and if for all with , there is so that is a non-trivial interval for all and .
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 is the smallest convex set containing .) That is, if is a line in and is the convex hull of and is an invertible affine map, then is also a line and is the convex hull of . Asking if is an interval, is equivalent to asking if is an iterval. Further, is an interval if and only if every line of the form that meets also meets .
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 , the box dimension by , and the Assouad dimension by . 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 for all sets .
The main result is the following theorem.
Theorem 2.4.
Let be a self-affine set satisfying the projection condition and the dominated splitting. Then for all with .
The proof goes via weak tangents, defined as follows. Let be a sequence of compact sets in . Say that converges to a compact set in , if
and
For and we write for the magnification function that shifts to origin and scales with factor . That is
Let be compact. Then is said to be a weak tangent of , and written , if converges to in for some sequences and . 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 for any fixed or for all 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 with . By the projection condition, it holds that is an interval for some , so . Thus the task is to prove the upper bound. Proposition 4.1 says that for all weak tangents of (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 , were is a weak tangent of with maximal Hausdorff dimension. ∎
Considering the proof of Proposition 4.1, it is obvious that if is a weak tangent of with in , then (by passing to a subsequence if necessary) it also holds that converges to a weak tangent, say , of . Of course , but unfortunately, it is not generally true that or . In particular, one can not just take the weak tangent of of maximal dimension and expect it to have anything to do with the weak tangent of 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 can be covered by graphs of few well behaving functions and that can be covered by a countable collection of lines. The arguments about visibility rely heavily on the projection condition.
The visibility conjecture asks if for almost all . So, the remaining step to confirm the conjecture in some special case, is to show that is of measure zero (or technically, that the set 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 , 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 by rectangles. The domination ensures that uniformly as . Therefore the approximation of can be done with a “very narrow” rectangle if is large. This motivates the following definition.
Definition 3.1.
For define the approximating rectangle to be the smallest closed rectangle that includes and has sides parallel to and . For any approximating rectangle , the length of the sides parallel to is denoted by and the length of the sides parallel to is denoted by . The orientation of cylinder is also called the orientation of the approximating rectangle .
Lemma 3.2.
Let be a self-affine set satisfying the projection condition and the dominated splitting. Let be a weak tangent of with in and let .
Then there exists a sequence of finite words and a sequence of approximating rectangles so that and and .
Proof.
Since , there exists sequences and so that and . Furthermore, can be chosen so that . By setting it is obvious that . Note also that and . By domination, there exist , so that for large . Thus,
and
∎
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 and fix for all . A set of the form
is called a Kakeya type set. The collection is called the direction set of the Kakeya type set.
Proposition 3.4.
Let be a self-affine set satisfying the projection condition and the dominated splitting, and let be a weak tangent of . Then , where is a Kakeya type set with direction set , that satisfies for all .
Proof.
Let converge to in and fix . By Lemma 3.2, there is a sequence of approximating rectangles with and so that . Recall that has orientation . Since , at least one of the shorter sides of is outside . One can now fix a direction that points to this side that is far away. To put this precise, choose a short side of that does not meet and extend this line segment to an infinite line from both ends and call this line for now (if there are two choices for the short side, then it does not matter which one is chosen). Note also that is perpendicular to . Then choose so that and meets for some . By passing to a sub-sequence, one can also assume that converges to some and by Lemma 2.1 part (3) it holds that . By the projection condition there is at least one so that is an interval whenever is large. Further by compactness of , the approximating rectangles , with large , have orientation bounded away from . Thus, by the projection condition and the choices made above, it is clear that
Trivially . So, by taking union over all , it then follows that
(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 . In particular, if then also for all small at least. If then the union doesn’t need to be over at all. Note however that even tough is on a line it may be that due to overlap of cylinders in the original self-affine set .
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 be a self-affine set satisfying the projection condition and the dominated splitting and let so that . Then for all .
As mentioned earlier the strategy is to cover with graphs of nice functions and a collection of vertical lines. With this in mind, recall some basic facts. For , let denote the graph of . That is, . A function satisfying for some and for all is called semi-decreasing. Also, is said to be semi-increasing if is semi-decreasing and 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 be semi-monotone. Then . Further the set of discontinuity points of 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 is semi-decreasing and that the involved constant is . Define by and consider . Since is monotone and is Lipschitz, the second claim follows. Also, since is monotone. On the other hand, , where is defined by . Clearly , since . On the other hand, it is easy to see that is bi-Lipschitz, so ∎
Proof of Proposition 4.1.
Fix a direction as in the claim and let be sequences so that in . After passing to subsequence if necessary, it can also be assumed that converges to some weak tangent in . By Proposition 3.4, the weak tangent is a Kakeya type set, so let and , so that . By assumption for all . Without loss of generality, assume that . Let and . Compactness of ensures that is strictly positive. Still, without loss of generality, assume that , so it suffices to estimate the dimension of . Set . The reason of focusing on this smaller ball inside , is merely a technicality and there is no need for the reader to worry about this too much. In a nutshell, if is a line or half line that meets , then may be different from . The choice of ensures that if is a line or half line included in the weak tangent, and it meets , then equals to .
First divide into three sets that each have nice enough geometry. Recall that consists of line segments, and only the lines that hit are meaningful. If is a collection of lines and half lines so that , and consists of those elements that meet , then set
For the lines in there are two possibilities. According to Lemma 3.2, for , it may be that there exists a sequence of approximating rectangles converging to in . In this case set . If this is not the case, then, since every approximating rectangle can have at most one short side meeting , there are two sequences of approximating rectangles, say and , so that in . (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 does not meet and the right side of does not meet . By passing to subsequences, one can assume that there are so that converges to in , with and that converges to in , with . In this case, set and .
Finally, set
and . The closures are taken to ensure that is a compact set for all . Obviously . Note that despite the closures, the sets are still collections of lines, since the closure of any set of lines in is still a set of lines in . Most importantly, for each line in , there still exists a sequence of approximating rectangles with , since they existed for all lines of .
Now it is time to estimate . Trivially, so it suffices to consider for separately. Of course, some of the sets may be empty, but at least one of them is nonempty since is nonempty.
Start with . Note that for some . Consider the function defined by . Let denote the strip . Obviously . Consider , with . Since is on a line segment , with , it is clear that , so is semi-decreasing, and by Lemma 4.2.
Similarly, define by setting . Again, . Consider , with . Since is on a line segment , with , it is clear that . Thus is semi-increasing, and by Lemma 4.2.
Finally, define by and note that is Lipschitz. All in all, the above considerations show that .
Then it is time to estimate . The aim is to show that 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 , b) the boundary points of . Considering the first case, Lemma 4.2 showed that a semi-monotone function can have only countably many points of discontinuity. Hence, set
For the second case, let be the endpoints of the intervals , and set
The final step is to show that . If this is not the case, then there exists a point . Since and is not visible, there exists with and for some .
Assume first that . Then, by the choices made above, there is a sequence of approximating rectangles, with side lengths and , converging to in . Let be a sequence so that and . (Recall that .) Let be so large that and for all large . This is possible since in both inequalities the left hand side converges to zero and converges to . Now, by the projection condition, there exists a point so that implying that , which is a contradiction.
Then assume that . Assume also that and that is continuous at , since otherwise is covered by or . Thus there exists with and and . Again, there is a sequence of approximating rectangles and points with . Let be a sequence so that and . When is so large that
(4.1) |
the projection condition implies that for all large implying that . See Figure 1 for clarification. The case is symmetric to the case .
The conclusion now is that and each element in the union has Hausdorff dimension , so the proof is finished. ∎
Considering the visibility conjecture, there is still the question whether . For dominated self-affine carpets, this is true. A self-affine set is called a carpet if all are diagonal matrixes. If a carpet is dominated, then 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 be a self-affine carpet satisfying the projection condition and the dominated splitting. Then holds for all except possibly one and its opposite .
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 so that and , and for all and that
(4.2) |
for all . As is intuitive, a cone is a union of set of lines through origin in that have bounded angle from some fixed line. In what follows, the cone is understood as a subsets of or depending on the situation, and this should not cause any confusion. So, equivalently, a cone is an interval in the projective space . As discussed in the proof of [13, Lemma 4.1] it follows that and for all . Further, without loss of generality, assume that is uniformly separated from for all independently of the length . This follows simply by choosing to be the minimal cone that includes and then applying the previous deduction to the cones . (Note that the strong cone separation holds with the cone as well.)
Corollary 4.4.
Let be a self-affine satisfying the projection condition and the strong cone separation. Then for almost all .
Proof.
The strong cone separation implies domination [3, Theorem B], so by Theorem 2.4 it is enough to show that . This would certainly follow form and this in turn follows if is porous. Recall that a subset of some metric space is porous if there are constants so that for all and , there is so that . For the connection between Assouad dimension and porosity, see for example [14].
Let be the cone from (4.2). Since is a union of lines through origin, it can also be considered as a subset of the projective space. Recall that is a metric space where the distance is measured by , the angle between the corresponding lines in (which lies in the interval as usual). An invertible linear mapping can naturally be interpreted as a mapping on the projective space as well since by linearity. Thus the mappings naturally generate collections of nested compact sets inside , since for all . To be more precise, I claim that setting , the collection satisfies the conditions
-
(M1)
for all and
-
(M2)
as for all .
Part (M1) is obviously true, but verifying part (M2) requires some work. If these conditions hold, then the collection is called a Moran construction and there is a unique compact set
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 for all , it follows that equals to .
Now, to see that also (M2) holds, fix , and . Further, fix unit vectors and so that
It immediately follows that
Further, it is no restriction to assume that and are non-negative. Since , there exists , so that and . Moreover, can be chosen to be independent of and the level , since is uniformly separated from . Since is fixed, there exists so that
(4.3) |
for angles and . Therefore, it follows that
(4.4) |
and as , which proves (M2).
It now suffices to show that , satisfies the following bounded distortion property: there are constants and so that
for all and , where
If the bounded distortion holds, then let be a gap between two neighboring first level cylinders and . Note that exists due to the strong cone separation. Let and . Let be a finite word with but . Then, due to the separation of the cones, there is a gap and
which shows that is porous.
To show the bounded distortion property, fix , . There is a small annoying technicality that the angle between lines in may be realized “outside” if the opening angle of is larger than . Therefore, let be so large that for all lines when and assume .
As before, fix unit vectors and so that
where and are positive. Next, note that depending on if is positive or negative, and that a similar formula holds for . Combining this with (4.3) gives
(4.5) |
for lines with . If , then it could be that the angle is realized outside , and the last inequality in the above estimate may not hold. In this case recalling the choice of still gives
(4.6) |
Since by the choice of , the lower estimate can be treated as a single case. Again, relying on (4.3) gives
(4.7) |
Combining (4.5), (4.6), and (4.7) gives
(4.8) |
and this proves the bounded distortion with the constants and . ∎
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 be a self-affine satisfying the projection condition and the dominated splitting. Assume further that the sets have Hausdorff dimension at most for all . Then for all .
Proof.
Assume that for some , since otherwise the claim for follows from Theorem 2.4.
Divide the cylinders of into different classes according to the angle that the orientation of the cylinder has with . Set and
(4.9) |
From Lemma 2.1, it follows that , where . It also follows that all the elements of the union are compact sets. The sets are not exactly self-affine but each of them is a finite union of affine images of . In particular, if , then for all and, by Lemma 2.1, . Thus Theorem 2.4 gives that . Recalling that, is an affine image of , 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 for with , , and . The associated self-affine set is a Bedford-McMullen carpet, and it is well known that
and
see for example [15]. In particular, . It is easy to see that the system is dominated and that .
To verify the projection condition, fix with . Note that for all , so can be fixed so that the angle between the x-axis and is small, say smaller than , for all , with . By Remark 2.3, it is now enough to show that if is a line so that , and , where is the convex hull of , then also . So let be such a line. The projection of to arbitrary direction is not an interval, but since the angle between and -axis is small, it is easy tho see that for some . Since has even smaller angle with the -axis, it follows that implies for some . Continuing in this manner yields a word so that for all . Thus , and so the projection condition holds.
Since domination and projection condition are satisfied Theorem 2.4 gives that
for all . However, it is also easy to see that, outside the tri-adic points on the -axis, there is no vertical alignment of points of . Moreover, for each the tri-adic point , there are at most two points, say and , of with . Therefore, there exist a countable set , so that . Thus .
The weak tangent of that satisfies is obviously , where is the middle thirds Cantor set. However, and so . Therefore, if is the weak tangent of that has maximal dimension, then .
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 , where . Consider firs the box dimension of just . For , consired the index for which is closest to . Then, to cover with intervals of length , it is essentially enough cover all of and the rest can be neglected. Anyway, at least intervals are needed. On the other hand,
and it is an exercise to show that . Thus it follows that
which implies that . If one considers , essentially the same calculation shows that . Because is countable, for almost all directions, and thus for almost all . (For dimensions , one can of course consider , the fold product of .)
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.