Stable intersections of conformal regular Cantor sets with large Hausdorff dimensions
Abstract.
In this paper we prove that among pairs of conformal dynamically defined Cantor sets with sum of Hausdorff dimensions , there is an open and dense subset of such pairs verifying . This is motivated by the work [1], where Moreira and Yoccoz proved a similar statement for dynamically defined Cantor sets in the real line. Here we adapt their argument to the context of conformal Cantor sets in the complex plane, this requires the introduction of several new concepts and a more detailed analysis in some parts of the argument.
1. Introduction
The aim of this paper is to prove a complex version of the main theorem in [1]. We will prove that among pairs of conformal dynamically defined Cantor sets with sum of Hausdorff dimensions , there is an open and dense subset of such pairs verifying . The analogous statement for dynamically defined Cantor sets in the real line was proved by Moreira and Yoccoz in [1]. Here we adapt their argument to the context of conformal Cantor sets, this requires the introduction of several new concepts and a more detailed analysis in some parts of the argument. The main new difficulties come from the fact that smooth real maps are naturally conformal, in the sense that their derivative preserves angles, and this is not true for maps in dimension two. A rough version of our main theorem is the following.
Theorem 1.1.
There exists an open and dense subset of the space of pairs of conformal Cantor sets in the complex plane such that satisfying the following. If , then
is dense in
In particular, .
The work of Moreira and Yoccoz was motivated by a conjecture of Palis, according to which generic pairs of dynamically defined Cantor sets in the real line either verify that their arithmetic difference has zero Lebesgue measure or non-empty interior. Palis conjecture emerged from his work with Takens ([2], [3]), where, in their study of homoclinic bifurcations for surface diffeomorphisms, they used the crucial fact that if , then has zero Lebesgue measure. Looking for a converse, Palis proposed its conjecture.
The study of homoclinic bifurcations has proved to be fruitful in the understanding of dynamics for surface diffeomorphisms. Complicated dynamical phenomena arise from them. For example, arbitrarily close to any diffeomorphism exhibiting a generic homoclinic tangency, there are open regions in which any diffeomorphism belonging to a residual set has an infinite number of sinks- this is the so called Newhouse phenomenon. Looking for analogous results and using similar ideas to those of Newhouse, Buzzard [4] proved the existence of an open set of automorphisms of with stable homoclinic tangencies. Buzzard’s strategy was very similar to the work of Newhouse [9], constructing a “very thick” horseshoe, such that the Cantor sets, this time living in the complex plane, associated to it would also be “very thick”. However, the concept of thickness does not have a simple extension to this complex setting and so the argument to guarantee intersections between the Cantor sets after a small pertubation is different.
Furthermore, Moreira and Yoccoz were able to use their solution to Palis conjecture in the study of homoclinic bifurcations for surface diffeomorphisms (see [5]). They proved that given a surface diffeomorphism with a homoclinic quadratic tangency associated to a horseshoe with dimension larger than one, the set of diffeomorphisms close to presenting a stable tangency has positive density at .
The authors of this paper have been trying to push all this theory to the context of Cantor sets in the complex plane and apply it to the study of homoclinic bifurcations of automorphisms of . In [6], Araújo and Moreira introduced the concept of conformal Cantor sets and also extended the concept of recurrent compact sets to this context, which serves as a tool to obtain stable intersection. More importantly, Araújo and Moreira showed that given a complex horseshoe of an automorphism of , one can associate to it a conformal Cantor set. Using all this, they were able to recast Buzzard’s example in terms of the theory of conformal Cantor sets.
In the paper [7], Moreira and Zamudio proved a multidimensional version of the scale recurrence lemma for conformal Cantor sets. The real version of this lemma was a key step in the solution, by Moreira and Yoccoz, of Palis conjecture. Moreover, Moreira and Zamudio used the multidimensional conformal version of the scale recurrence lemma to prove a dimension formula relating the Hausdorff dimension of the image of a product of conformal Cantor sets , where is a function, and the sum of the Hausdorff dimensions .
In this paper we intend to use the results in [6], [7] and obtain a version of Palis conjecture for conformal Cantor sets. In a future paper we plan to use the concepts and ideas laid down in this paper to study homoclinic bifurcations for automorphisms of . We plan to obtain results analogous to the ones in [5]. We expect that conformal regular Cantor sets in will play a role in the study of homoclinic bifurcations of automorphisms of , similar to regular Cantor sets in in the study of homoclinic bifurcations of surface diffeomorphisms.
The paper is organized as follows. In section 2 we fix the notation, present basic concepts and results which will be used later. Most of the proofs are omitted and references are given to the works [6], [7], [8] and [9]. We also reduce our work to the proof of theorem 2.22; its proof is long and contained in the remaining sections. In section 3 we define a random family of Cantor sets and analyze some of its geometrical properties; it is from this family that we will find the pair of Cantor sets in the conclusion of 2.22. In section 4 we prove theorem 2.22, assuming propositions 4.1 and 4.2. The proof of these propositions is postponed to the last two sections (5 and 6) of this work.
2. Notation and preliminaries
In this section we give the definitions of the objects appearing in this paper and recall some results regarding them that have appeared on previous works [6], [7], [8]. We also present new results and reduce the proof of the main theorem to the proof of theorem 2.22.
2.1. Notations
Here we introduce some of the notations we will use along the text:
-
•
We will work with the space of complex numbers . We identify it with in the usual way. In this space we will use the Euclidean metric, given by the norm .
-
•
We will also work with the space of non-zero complex numbers . We will sometimes identify with , where , through the map . Note that has the structure of a commutative group. We will endow with the metric coming from the inclusion .
-
•
Given a linear map we denote its norm by , its minimum norm by . They are given by
We will say that is conformal if .
-
•
We will use to denote the identity matrix, sometimes we will use the same symbol for the identity function. Each case will be clear from the context.
-
•
Let be a metric space and . We will use the notation for the open neighborhood around , i.e. . For a point and a positive real number we use the notation .
-
•
We will have to deal with many inequalities and several parameters. In order to reduce the number of constants introduced along the text we will use the following notations: Given expressions and depending in the parameter , when we write this means that there is a constant such that for all . The constant can only depend on other constants when those have already been fixed. This ensures that we will not get any contradiction between the different constants that will appear. will mean , will mean and both hold.
-
•
Derivative of a function, we use two notations or , both mean derivative of order of at the point . They are -linear functions.
-
•
The uniform norm of functions , where and are subsets of normed vector spaces, will be denoted by . In many occasions, will be the derivative of order of a function.
-
•
Given a real number we will denote by its integer part, this is .
2.2. The space of conformal regular Cantor Sets
We begin by the very concept of conformal regular Cantor set. First we remember that a regular Cantor set, also called dynamically defined Cantor set, is given by the following data.
-
•
A finite set of letters and a set of admissible pairs.
-
•
For each a compact connected set .
-
•
A map , for , defined on an open neighbourhood of .
These data must verify the following assumptions:
-
•
The sets , , are pairwise disjoint.
-
•
implies , otherwise .
-
•
For each , the restriction can be extended to a diffeomorphism from an open neighborhood of onto its image such that for some constant , where is the minimum norm of the linear operator on .
-
•
The subshift induced by , called the type of the Cantor set,
, is topologically mixing.
Once we have all these data we can define a Cantor set (i.e. a totally disconnected, perfect compact set) on the complex plane:
We say that such a set is conformal if, for all , the derivative of at , denoted by , is a conformal linear operator.
Notice that we always consider the degree of differentiability of the map , , to be a real number larger than one. This means that has derivatives up to order and is Hölder with exponent . This hypothesis, as we will precise later in this section, allows us to control the geometry of small parts of the Cantor set . All Cantor sets in this paper will be conformal regular Cantor sets.
Besides, as we mentioned on the introduction, an important family of dynamically defined sets are contained in the class of conformal regular Cantor sets. Let be an automorphism of exhibiting a horseshoe and be a hyperbolic periodic point in it. Then, there is a subset , open in the topology of as an immersed manifold, and some sufficiently small such that is, after some parametrization, a conformal regular Cantor set. See Theorem A of [6].
We will usually write only to represent all the data that defines a particular dynamically defined Cantor set. Of course, the compact set can be described in multiple ways as a Cantor set constructed with the objects above, but whenever we say that is a Cantor set we assume that one particular set of data as above is fixed. In this spirit, we may represent the Cantor set by the map that defines it as described above, since all the data can be inferred if we know .
Notice that in our definition we did not require the pieces to have non-empty interior. To circumvent this, we introduce the following sets.
Lemma 2.1.
There is sufficiently small such that the sets satisfy:
-
(i)
is open and connected.
-
(ii)
and can be extended to an open neighbourhood of , such that it is a embedding (with inverse ) from this neighbourhood to its image and .
-
(iii)
The sets , , are pairwise disjoint.
-
(iv)
implies , and implies .
The sets could substitute the pieces in our definition as to make the hypothesis of open interiors be true. These changes do not enlarge the Cantor set. To see this, we introduce more notation and a previous result.
Associated to a Cantor set we define the sets
Given , , and , we write:
-
•
if , ;
-
•
if ,
-
•
if and , , in which for all and .
-
•
Define the distance between and by .
-
•
if starts with , we define as the unique finite word such that .
For we say that it has length and define:
and the function by:
Notice that . Furthermore, we can consider the sets defined in the same way
but using the version of the pieces and consider the function to be defined in the larger set having image equal to .
Now we have the following lemma.
Lemma 2.2.
Let be a dynamically defined Cantor set and the sets defined above. There exists a constant such that:
As a consequence of this lemma we can see that
since and , and so the Cantor set has not been enlarged. Another consequence is that is contained in the interior of the union of the pieces . From now on we will work with the assumption that the sets have non-empty interior and that they contain in the interior of their union. We keep the definition as before because it will be useful in the definition of limit geometries.
The following definition will be useful in the future.
Definition 2.3.
For every Cantor set we define the homeomorphism
that carries each point to its itinerary along the pieces , that is
if and only if for all .
Our main result concerns generic Cantor sets. So now we fix the topology on the space of Cantor sets. We remind that any Cantor set we are considering is given by a map that is, at least, for some .
Definition 2.4.
(The space ) For a fixed symbolic space and real number (we also allow m=). The set of all conformal regular Cantor sets with the type is defined as the set of all conformal Cantor sets described as above whose set of data includes the alphabet and the set of admissible pairs used in the construction of . We denote it by .
The topology on is generated by a basis of neighbourhoods where is any Cantor set in and . The neighborhood is the set of all conformal regular Cantor sets given by such that , (that is, the pieces are close in the Hausdorff topology) and the restrictions of and to are close in the metric. The topology on is the one such that a sequence of Cantor sets converges to if and only if the sequence converges to in the topology of for every .
We also consider the union , the topology in is the finest topology such that the inclusions are continuous maps, the so called inductive limit topology. Thus, a set is open if and only if is open in for all . It is not difficult to prove that an open set can be written as a union , where each is open in and if .
2.3. Limit geometries
To study the geometry of small parts of our Cantor sets, we introduce more objects. For each , denote by the set . For each , fix a point . We will refer to these points as base points. Define by
Additionally, given we write and .
Given and , define as the unique map in
such that
The maps act as a normalization of small parts of the Cantor set . For that purpose, we define the maps by
Through them we have the first result that allows control over the sets .
In what follows we consider some fixed and being a Cantor set in the space .
Proposition 2.5.
(Limit Geometries) For each the sequence of embeddings converges in the topology to a embedding . The convergence is uniform over all and in a small neighbourhood of in .
The defined for any are called the limit geometries of .
Remark 2.6.
Define and consider in this set the topology given by the metric . Likewise, for , let be the space of embeddings from to with inverse equipped with the topology given by the metric
For fixed and a Cantor set , the map is -Hölder, if we consider the metrics described above for both spaces. In case the Cantor set is , for , then there is a constant such that . The constant can be chosen uniformly in a neighborhood of the Cantor set.
Since the convergence is uniform with respect to and in a neighborhood of , the limit geometries depend continuously in and the Cantor set .
We also remark that the derivative is conformal for all .
Remark 2.7.
It is important to observe that limit geometries depend on the choice of base points. Nonetheless, different choice of base points do not change the resultant limit geometries by much, only by an affine transformation that is bounded by some constant depending on . Here we mean that such transformations are given by maps , where . This bound is, as before, uniform for Cantor sets sufficiently close to . See the paragraph after Corollary 3.2 of [6].
For reasons that will become more clear in the future, from now on we assume that for each the corresponding base point is a pre-periodic point.
Before proceeding, we fix some more notation. For and we write
Furthermore, to establish stable intersections, we are going to analyse very small parts of the Cantor sets, whose size will be controlled by a real number . This number should be regarded as a variable that we are going to assume in various instances to be very small, as to make the various estimates we are going to find in the future to fit all together. This being said, let be a sufficiently large constant.
Definition 2.8.
For , the set is defined as the set of words such that
We say that the set has an approximate size .
Using standard techniques (see [2] and [8]), one can prove that there is a constant depending only in the Cantor set and the parameter such that
(1) |
where is the Hausdorff dimension of .
Remark 2.9.
Notice that, because the set of limit geometries represent a compact subset of , every piece of approximate size also contains the ball for some depending only on . The result remains valid for perturbations sufficiently close to . Even more, by maybe enlarging a little bit, because of Corollaries 3.3 and 3.4 of [6], it follows that for any
(2) |
This allows us to control the approximate size of the sets through the derivative of the map at .
2.4. Recurrent compact criterion for stable intersections
Our next objects are called configurations. They are a way of moving a Cantor set in the plane without changing its internal structure.
Definition 2.10.
A -configuration of a piece of a Cantor set is a , , diffeomorphism
The space of all configurations of a piece is denoted by and we equip it with the topology. The space of all configurations is denoted by and we equip it with the inductive limit topology. This is, is open if and only is open in the topology of , for all .
If is an affine map, we call it an affine configuration. Observe that a limit geometry is a configuration of a piece. Configurations of the type , where and , are of great importance to our work and so are called affine configurations of limit geometries.
The renormalization operators represent a way of looking into smaller parts of the Cantor set.
Definition 2.11.
Let and be two Cantor sets. Choose any pair of words and . Then, the renormalization operator acts on any pair of configurations and by
The notation above clearly indicates that we can consider the operators and as separate, each acting on configurations of and respectively. In a very similar way to the proposition 2.5, the one defining limit geometries, one can show (see Lemma 3.11 of [6]) that the set of affine configurations of limit geometries form an attractor in the space of configurations under the action of renormalizations. Even more, see lemma 3.8 of [6], the renormalization operators act in a very simple manner over limit geometries:
Lemma 2.12.
For any and , , there is an affine transformation such that
Moreover, this transformation can be calculated by
Now we properly establish the notion of stable intersection between Cantor sets. Given two Cantor sets and any pair of configurations we say that it is:
-
•
linked whenever .
-
•
intersecting whenever .
-
•
has stable intersections whenever for any pairs of Cantor sets in a small neighbourhood of and any configuration pair that is sufficiently close to in the topology of .
The set in the statement of theorem 1.1 represents the set of all such that is a pair of configurations having stable intersections in the sense just described, where is the translation by on .
The main theorem in the introduction is a consequence of theorem 2.23, because it guarantees stable intersections for affine configurations of Cantor sets. In turn, theorem 2.23 is a consequence of theorem 2.22. The statement of this theorem requires that we recall some more concepts. First, notice that the space of affine configurations of limit geometries of a Cantor set can be seen as the image of the continuous association
Definition 2.13.
The space of relative affine configurations of limit geometries will be denoted by . It is the quotient of by the action of the affine group by composition on the left, that is, , where ranges in .
The concepts of linking, intersection and stable intersection were well defined for pairs of affine configurations of limit geometries, and since they are invariant by the action of , they are also defined for relative configurations in .
Also, since the renormalization operator acts by composition on the right on , its action commutes with the multiplication on the left by affine transformations and so it can be naturally defined on . This space can be identified with by the identification and, in this manner, the topology on is the product topology on . The action of the renormalization operator on a relative configuration can be described by
and if , then
(3) |
It is more convenient to see the space through one more identification:
where . We will call the scale part of the relative configuration and the translation part. The equation (3) provides to us a formula for the renormalization under this identification if we analyse the scale and translation parts separately:
The space of relative scales is given by , where . We identify with through the map . It acts on by complex multiplication. The space of relative configurations projects to by the map
where means derivative of the affine map (which is an element in ). We trivialize in the following way: we map to such that and . In this sense we can think of as . The renormalization operators act on the space of relative scales by
Most of the time we will work with scales which are bounded away from zero and infinity, for this purpose we introduce the notation
where is a positive real number.
The object which we present in the following definition will play a central role in the proof of our main theorems. It is a useful tool to get stable intersection between pairs of Cantor sets.
Definition 2.14 (Recurrent compact).
Let and be a pair of Cantor sets. Let be a compact set in . We say that is recurrent (for the pair ) if for any relative affine configuration of limit geometries , there are finite words , such that satisfies , where the are renormalization operators associated to the pair of Cantor sets and .
If such a renormalization can be done using words and such that their total size combined is equal to one, we say that such a set is immediately recurrent.
Theorem B of [6] states that if belongs to a recurrent compact set associated to a pair of Cantor sets and , then it represents pairs of affine configurations of limit geometries of these Cantor sets that have stable intersections. For the convenience of the reader, we copy its statement below.
Theorem.
The following properties are true:
-
(1)
Every recurrent compact set is contained in an immediately recurrent compact set.
-
(2)
Given a recurrent compact set (resp. immediately recurrent) for , , for any , in a small neighbourhood of we can choose base points and respectively close to the pre-fixed and , for all and , in a manner that is also a recurrent compact set for and .
-
(3)
Any relative configuration contained in a recurrent compact set has stable intersections.
Remark 2.15.
For each pair of maps in the small neighbourhood of in the theorem above, let and be the corresponding homeomorphisms defined in 2.3. The base points and in the theorem above are chosen so that and for all and all , meaning that their itineraries under the action of and are the same for all pairs of maps in this neighbourhood. In subsequent contexts, the base points will be chosen in the same way.
2.5. Perturbation of Conformal Cantor Sets
Let be a conformal Cantor set defined by a map . We show that if is non-essentially real then arbitrarily close to , in the topology, there is a conformal Cantor set defined by a map that is holomorphic on a small open neighbourhood of . This is an important property that will allow us to perturb more freely the conformal Cantor sets and adapt the random perturbation argument from [1] to our context.
We begin with the following lemma:
Lemma 2.16.
Let be a conformal Cantor set given by . For consider the set
Assume that, for all , has two linearly independent vectors (over ). Then, for all and the -linear map is conformal, i.e. there is a complex number such that
The operation in the right hand side of the last equality corresponds to complex multiplication.
Proof.
Notice that the case is just the definition of conformality for the Cantor set. Now we proceed by induction, assume the result for . Let , then there are sequences and such that . Hence
This shows that the limit exists, denote it by . Moreover
If we take another vector , and using the symmetry of the operator , we would have
This shows that does not depend on , denote it by . Since we can choose generating , we conclude that
∎
To use this lemma we need to consider Cantor sets that are indeed two dimensional. This concept is precised by the following definition.
Definition 2.17.
We will say that a Cantor set is essentially real if there exists such that the limit Cantor set is contained in a straight line. Otherwise, we say it is non-essentially real.
It is not difficult to prove that is essentially real if and only if for every the limit Cantor set is contained in a straight line. Moreover, one can prove that being essentially real is equivalent to being contained in a one dimensional manifold embedded on the plane.
Lemma 1.4.1 from [8] can be adapted to our context and it can be used to prove that every point belonging to a non-essentially real Cantor set verifies that has two linearly independent vectors (over ). We now show that being non-essentially real is an open property.
Lemma 2.18.
Let be a non-essentially real conformal Cantor set. Every conformal Cantor set, close enough to in the topology, is also non-essentially real.
Proof.
If the lemma does not hold, we would have a sequence of conformal Cantor sets converging to and such that every is essentially real. Let , denote by the limit geometry associated to and the Cantor set . Since all are essentially real then, for all , is contained in a line passing through the origin. Taking a subsequence we can assume that, as goes to infinity, converges to a a set contained in a line passing through the origin. Using the fact that limit geometries depend continuously on the Cantor set, we conclude that is contained in a line and therefore is essentially real, contradicting the hypothesis in the lemma. ∎
Lemma 2.19.
Let be a non-essentially real conformal Cantor set. Arbitrarily close to , in the topology, we can construct a conformal Cantor set such that is holomorphic on a neighbourhood of .
Proof.
Since is non-essentially real then the -derivative , at a point in the Cantor set, is determined by a complex number, which we denote by , i.e.
In this situation, the Taylor approximation of at the point is
Hence, is approximated (close to ) by a complex polynomial, which is an holomorphic function. Now, to globally aproximate by a function which is holomorphic in a neighborhood of its Cantor set , we are going to take many of the previous polynomial approximations and glue them together.
Choose any real number larger than zero. Consider such that is a partition of . For each we choose a point and define the polynomial
We can also consider functions , , with the following properties:
-
•
for all .
-
•
, for a constant independent of .
-
•
, for all .
-
•
, for a constant independent of .
Indeed, to be able to construct these bump functions, we only need to show that given both in , the distance between the pieces and is at least for some constant independent of . For that, we can suppose that and for some , since this would be the worst scenario. If ends with , the distance between these sets is comparable to
Hence the existence of follows from the compactness of the space of limit geometries.
Now, let be and very close, in the topology, to . Define , with the same domain as , by
Notice that
Therefore, the norm of will be small provided
is small, for all , (remember that support of is contained in ). We already know that . On the other hand, Taylor approximation implies that
We conclude that taking small enough and close enough to , we get as close to as we want. Notice that thanks to the way in which we defined , it is and by lemma 2.18 we can suppose it is non-essentially real. Moreover, in the set the function is holomorpic. We can also guarantee that verifies the hypothesis necessary to define a dynamically defined Cantor set (with the same sets , ), we just need to take -close enough to . Even more, the Cantor set , generated by , is contained in . ∎
To prove our main theorems we will use the scale recurrence lemma (see subsection 2.7). To use this lemma we need that our Cantor sets are non-essentially affine. A Cantor set , with , is said to be non-essentially affine when there is a pair of limit geometries and in such that and a point such that
The following lemma allow us to perturb and get a non-essentially affine Cantor set.
Lemma 2.20.
Let be a non-essentially real conformal Cantor set. Arbitrarly close to , in the topology, there is a Cantor set which is non-essentially real, non-essentially affine and such that its expanding function is holomorphic in a neighborhood of .
Proof.
Let be the perturbed Cantor set from lemma 2.19. If is non-essentially affine we are done. Otherwise, choose a piece , , and let be the corresponding base point. As previously mentioned, it is pre-periodic.
Claim.
If is sufficiently small, we can chose ending with so that no word in appears more than once in . Given any word with this property, if is sufficiently small, there is ending with and such that does not appear elsewhere in . Furthermore, there is ending with such that no subword of belonging to appears in it.
Proof.
Since the shift is mixing over , there must be at least two sequences of distinct lengths (both larger than 1) such that both end and begin with and no other letter in these words is . Now construct as
If is sufficiently small, a suitable choice of , , and can be done so that no subword in appears more than once. This can be seen by analysing the behaviour of the distance between two consecutive letters in any two subwords of .
Now, choose a subword such that begins with . If never appears in , then we can make . Indeed, for to appear more than once, the word must be contained in but in another position. This implies that a subword in , corresponding to the beginning of for example, appears more than once in . But this does not happen.
Now, suppose we are given a word . We want to prove there is some such that never appears in . Choose a beginning of and an ending . Let be such that and never appear in it and also suppose that the first letter of is the same as its last letter. Then, similarly to the analysis before, we can make .
The existence of comes from a counting argument, in which we show that the words in which or appear do not account for all possible words . Remember that if is the Hausdorff dimension of , then .
The number of words in ending with is . More than that, if we fix a starting position for the appearance of , such as it beginning in the letter of (remember is very small), then the same estimate remains true. Notice however that the number of letters of is , and so the number of words in that fail our requirements is , thus, for small enough, there must be that satisfies our requirements.
To construct , we can use the same argument, all we need to observe is that the number of subwords of in is also .
∎
By maybe shrinking even further, we can assume that the periodic part of the itinerary of is a word very small when compared to . The combinatorial conditions above imply that belongs to only when . Besides, never belongs to this set.
Let be an holomorphic map close to the identity, and suppose it fixes the point and has derivative equal to the identity at this point. Similar to the construction on lemma 2.19, define a new Cantor set given by an expanding map that is equal to outside a small neighborhood of and equal to in . Note that the perturbed base point is equal to , thanks to the pre-periodicity of and the combinatorial properties of . Moreover, the limit geometry corresponding to stays the same close to , that is, we can choose a neighbourhood of the base point , with sufficiently large, such that . On the other hand, for ,
since the affine reescalings stay the same. Notice that the map is still holomorphic in a neighbourhood of and, because of lemma 2.18, it is non-essentially real if is sufficiently close to the identity. However, we can still choose so that
Hence . This implies that
and so the new Cantor set is also non-essentially affine.
∎
Remark 2.21.
We observe that the Cantor set constructed in lemma 2.19 can be also assumed to be close to in the topology for all . All one needs to do is to choose close to in this topology. This can be done using a mollifier supported in a very small neighbourhood of the origin and making . Notice that the proximity between the maps and in comes from the fact that is and has diameter of order . Moreover, using the fact that is -Holder, for , one gets the improved estimate for all . This can be seen analysing the Taylor approximation of the derivatives of in this domain.
It follows that, given any Cantor set , essentially real or not, arbitrarily close to in the topology of , there is a Cantor set that satisfies the conclusion of lemma 2.20. Indeed, suppose that is for some . Then the only part of the argument (polynomial approximation in the topology) in lemma 2.19 which uses the non-essentially real hypothesis can be done using the conformality of the Cantor set. Moreover, one can also choose to be non-essentially real. To do so, we first suppose that the expanding function is already holomorphic in a neighborhood of . Then one chooses a periodic point , of period , and observes that if then can not be essentially real. Thus, if we are done, otherwise we perturb along the periodic orbit to get such property. One execution of this idea of perturbing along a periodic orbit can be found in the proof of theorem 2.23.
2.6. Main theorems
Here we state our main theorems. The proof of theorem 2.22 will be given throughout the remaining sections. Using this theorem we will prove theorem 2.23. In particular, we will get that there is an open and dense set, among pairs of conformal Cantor sets , with , such that all elements in this set verify . Before stating the theorems, we remark that the Hausdorff dimension varies continuously with the Cantor set. This is proven in [9] for Cantor sets in the real line and the argument there can be adapted to our context.
Theorem 2.22.
Given a pair of non-essentially real conformal Cantor sets in , such that . Arbitrarily close to , , in the topology, we can find conformal Cantor sets , in , respectively, such that , has a non empty recurrent compact set.
Define the set as the pairs of conformal Cantor sets in such that for every relative configuration , there is such that the configuration has stable intersection.
Theorem 2.23.
The set is open in and is dense, in the topology, in . Moreover, for any in and such that and are conformal for all , the set
is dense in
In particular, .
Proof.
The proof is very similar to the one for the corresponding result in [1], except for the use of lemma 2.25. For the openness of , one observes that, if is big enough, then any relative configuration can be transported, using a renormalization operator, to the set . Given in , from compactness of the set one sees that there is a neighborhood of such that for any pair in this neighborhood, we have implies there is such that has stable intersection. Thus the same happens for the whole , and the neighborhood is contained in .
Furthermore, in the same context of the previous paragraph, from compactness of the set , for each there is some such that if and are maps -close to the identity in the metric, then has stable intersections, where and . We will need this later.
For the denseness we use theorem 2.22. Given , with and both of them non-essentially real, arbitrarily close to it there is having a non empty recurrent compact set . Perturbing, we may assume that also has periodic points , , associated to finite words , , that is , , such that if we write
where is concatenation of with itself times, similarly for , and , then the set is dense in (see lemma 2.25).
More precisely, in order to get the density of , we need to perturb such that have the property in lemma 2.25, where , are the periods of , , respectively. To do this, we define a family of conformal Cantor sets given by expanding functions depending in complex parameters , , such that the pairs form an open set. Choose a word such that , is small enough such that there are not more points of the periodic orbit of contained in and is holomorphic in (by lemma 2.19 we may assume this). For small enough one has that . We choose a function such that
where is in the ball of center and radius in . Define , notice that we can take as close as we want to in the topology by choosing small enough. If we choose small enough then we can guarantee that the Cantor set , associated to , and the set are contained in and therefore , which implies that
is conformal for all . This proves that is a conformal Cantor set. Moreover, notice that is still a periodic point of with the same period , and . Doing the same construction for one sees that for some value of , the pair satisfy the hypothesis of lemma 2.25.
Using equation (3), the denseness of the set and the fact that has non-empty interior, one concludes that for any there is , , in , , respectively, and such that . Therefore .
For the final part, let be such that is intersecting and take any . Then there is at least one pair of words and sufficiently large such that are still intersecting and the diameters of the sets and are smaller than . Indeed, if it was not the case, the sets and would be disjoint for all and sufficiently large, a contradiction with the intersecting hypothesis. We will prove that if is small there exists such that and has stable intersections.
Now, following definition 3.10 of [6], we can scale these pairs of configurations by normalizing in the second coordinate, obtaining another pair with intersection. More precisely, we choose such that
and consider now the pair of configurations . This pair of configurations is intersecting, because this property is clearly preserved under composition on the left by affine transformations.
Let be such that and are both . Reasoning as in the proof of lemma 3.11 of [6] (see claim 3.12), we observe that this pair of configurations can be written as
where: ends with ; ends with ; the maps and are close to the identity in the topology, and is a bounded affine transformation in . More precisely, if we set111Notice that if were not conformal for all we could not guarantee that .
and choose and with appropriate lengths, then and there is some constant (independent from ) such that and the distance of the maps and to the identity is bounded by and
Consider now the relative configuration . By the previous part, there is some such that, writing , the pair of configurations has stable intersections for every pair of maps -close to the identity in the metric. Notice that, since is intersecting, is bounded by . Therefore, by compactness of the set of all pairs of limit geometries , we may enlarge , still independently of , so that is bounded by .
Now we choose such that ; it follows that
where , , and the distances of and to the identity are bounded from above by and . Choosing sufficiently small, and so and very big, these distances to the identity are less than . This implies that have stable intersections. Notice, finally, that
Hence, making very small, we approximate by that has stable intersections, completing the proof. ∎
Remark 2.24.
Notice that thanks to remark 2.21, the set is dense, with respect to the topology of , inside the set .
Lemma 2.25.
Let , and consider the subgroup . Let be representatives of , respectively. Then is dense if and only there is not such that
Moreover, if is dense and then is dense. The set of pairs such that is dense is a countable intersection of open and dense sets.
Proof.
The lemma is proved using Kronecker’s theorem. It states that a vector generates a dense subgroup if and only if there is not such that .
Let be the canonical projection and choose such that , . Note that is dense if and only if the set
generates a dense subgroup in .
Moreover, this last property is invariant under invertible linear transformations in . Let be the linear map such that and . Then, is dense if and only if the set
generates a dense subgroup in . It is clear that this happens if and only if the projection of to generates a dense subgroup in . Thus, using Kronecker’s theorem we see that is dense if and only if there is not such that
Moreover, it is not difficult to see that . This proves the first part of the lemma.
Notice that the set of pairs such that is dense corresponds to the intersection, varying , of the sets
and each one of these sets is open and dense.
Finally, will be dense if and only if
is dense in . If is dense, then the projection of to is dense. If we also have , then from the expression
it is not difficult to see that is dense in this case. ∎
2.7. Scale Recurrence Lemma
In this section we will see how to adapt the scale recurrence lemma from [7] to our context. First we note that limit geometries in [7] were defined slightly different, they were defined as the limit of the function
where is the unique affine function satisfying , and . Denote those limit geometries by . It is clear that there is a complex number such that . It is not difficult to prove that one can go to the limit and find a complex number such that . Moreover, is uniformly bounded from above and below, i.e there is such that . One can also show that depends Lipschitz in in the sense that there is a constant such that
We will denote by the affine function defined by
This affine function can be written in terms of the numbers , and by the formula
The maps and are related by the equations
Now we assume we have two Cantor sets, and , and discuss how to go from the renormalization operators in [7] to ours. Define and by
Remember that we identify with through , and define given by
From the previous equations, one easily proves that the renormalization operators of [7], which are given by
and act on , are related to our renormalization operators by the “semiconjugacy” , i.e. . Using , it is not difficult to transport the scale recurrence lemma from [7] to one for :
Lemma 2.26.
Suppose that , are non-essentially affine and non-essentially real. If are conveniently large, there exist with the following properties: given , and a family of subsets of , , such that
there is another family of subsets of satisfying:
-
(i)
For any , is contained in the -neighborhood of .
-
(ii)
Let , ; there exist at least pairs (with , starting with the last letter of , ) such that, if , end respectively with , and
the -neighborhood of is contained in .
-
(iii)
for at least half of the .
Remark 2.27.
Here is the unique measure in giving measure to and invariant by translations. We notice that the lemma remains true if we change by Lebesgue measure in . Since all sets are in , one just would need to redefine the constants and . The same happens for the metric on , we prove the lemma with the metric
and ( is the class generated by ). In we can change this metric for the usual metric in .
We remark that it can be assumed the sets are closed. To do this we just have to redefine by taking their closure and increase the parameter .
Proof.
We are in the hypothesis of the scale recurrence lemma in [7]. Let , , , , and be the constants given by the lemma. We choose big enough and small enough such that
for any set with , where and is the Haar measure in such that . This can be done since is multiplication by a complex number, whose norm is uniformly bounded and away from zero. Indeed, if one chooses big such that and small enough such that , then using that one gets
We choose , and , the other constants will be chosen along the proof. Suppose we are given a family of sets as in the setting of the lemma, define a new family by
where the union is over all finishing in , respectively. Notice that thanks to the previous discussion one gets that . Then we can apply the scale recurrence lemma from [7] to get a new family , which satisfies the properties given in [7]. Now we go back to the space , define a new family by
where the union is over all pairs ending in , respectively. We will prove that the family satisfies the desired properties:
-
(i)
Note that is Lipschitz with constant . Enlarging the constants , we can suppose that , are Lipschitz with constant and . Using this we get
Taking gives the desired property.
-
(ii)
Let , then for some ending in . Let such that . Let be one of the pairs, associated to , given by the scale recurrence lemma in [7]. If we write
we know that the ball is contained in . This implies
Thus it is enough to take .
-
(iii)
Let such that and ending in , we have
The fact that for at least half of the implies immediatly the desired property with big enough.
∎
3. Random Perturbations of conformal Cantor sets
3.1. Random perturbations
From now on we will focus on the proof of theorem 2.22. Let be a pair of non-essentially real conformal regular Cantor sets such that . We first perturb, in the topology, the pair of Cantor sets so they satisfy the hypothesis of the Scale Recurrence Lemma and the map defining the Cantor set is holomorphic on a neighborhood of . All this can be done thanks to lemma 2.20. Applying now the scale recurrence lemma gives constants verifying the conclusions of the lemma. With the aim of reducing the number of constants, we will also assume, without loss of generality, that the diameters of the sets are all less than one. This can be achieved by changing the metric. To prove theorem 2.22, we will now only perturb the Cantor set , leaving unaltered.
Notice that a neighbourhood in contains a neighbourhood in for some integer . So from now on we fix this integer . The desired perturbation for will be picked by a probabilistic argument out of a family of random perturbations that we will now construct.
The following constructions and arguments are made having a parameter in mind. All constants from now on are independent of this parameter, and everything fits together in the end by choosing sufficiently small.
We first pick a subset of such that
is a partition of into disjoint cylinders.
We then define as the subset of formed of the words such that no word in appears twice in . To see that check the claim in the proof of lemma 2.20.
Let be a constant222This constant corresponds to in [1], since we already used this symbol in the scale recurrence lemma then we changed it to . sufficiently close to to have the following: let
for ; then the , , are pairwise disjoint.
For each we choose a smooth function satisfying:
Notice that, since , we can choose these functions in a way that , for all and some constant independent of (but not from ).
The probability space underlying the family of random perturbations is , where is the unitary disk in , equipped with the normalized Lebesgue measure.
For , we define to be the time-one map of the vector field
where is a conveniently large constant, to be chosen later. Finally, we define to be .
By our previous estimative on we have that is .
Since , for any , is close to the identity in the -topology, then is close to . Taking small enough we can suppose that generates a Cantor set (with the same family of sets , ), which we denote by . Moreover, taking small it can be proven that this Cantor set is in fact a conformal Cantor set. Indeed, let be the open set containing where the function is holomorphic. If is sufficiently small then333This is consequence of lemma 3.2 part (ii), note that the proof of this part of the lemma does not use the conformality of at the points in the Cantor set. We can also get this from the fact that is an hyperbolic set for and use continuation of the hyperbolic set (see Theorem 7.8 in [10])., for any and , and . It follows that
which is a conformal linear transformation.
Our task will be to find such that the pair of Cantor sets determined by have a non empty recurrent compact set of relative configurations.
Remark 3.1.
All the objects introduced in section 2 are well defined for the Cantor sets and we will denote them by adding a superscript indicating the corresponding value of , such as , , and for example. Notice however that these Cantor sets have the same type as , and therefore are close to in the topology. Besides, we consider for each the natural conjugation between the dynamical systems and
which carries each point to the sequence that satisfies . For each , we have a pre-periodic sequence that begins with , defined by . The set of base points for satisfies
for every . This is important for the study of limit geometries.
3.2. Some properties of the family
Let and be such that and begins with ; let . Any perturbed inverse branch is well defined in the neighborhood and for any
(4) |
Notice that , therefore
for small enough. This implies that
and
for all . This in turn immediately implies the formula for .
Lemma 3.2.
Let and be the homeomorphism defined in remark 3.1. If is sufficiently small,
-
(i)
for any , we have ;
-
(ii)
for any , we have ;
-
(iii)
for , we have
-
(iv)
for and a word with such that , we have
The constant is independent of , , , , and the size of the perturbation.
Proof.
(i): Let be the maximum distance between corresponding inverse branches of and . We will prove that by induction on . For we have , this is a direct consequence of .
Observe that covers all the pieces it intersects, therefore there exists sufficiently small such that if , then . Consequently, if , any point such that and the line segment joining and are contained in the extended domain of . In this domain, . Suppose . If is sufficiently small, then . Given a word , we write . Given a point ,
Of course, when writing this, we are assuming that belongs to the domain of . But this is the case when , which is true by hypothesis. More than that, because the segment joining the two points is inside this domain,
In this manner, choosing , we obtain , finishing this part.
(ii): Let . It follows that and . As the diameters of these sets converge exponentially to zero, the result follows from (i).
(iii): We now study the perturbed limit geometries. Notice that the base point used to define is not the same as the one for , but the estimate of (ii) gives us control over this displacement.
Fix and let . Let the base point be given by and the base point be given by for some fixed sequence . From (ii),
Write and ; and and for . Notice that
by (i). Likewise, seeing that ,
(5) |
by (i) again and the estimate for above ( is a contraction).
Remember that , where is an affine transformation, and . Hence
and similarly
The difference is thus equal to
(6) |
Let us analyze this expression for not very large. Define
for . If is such that (remember that was defined in eq. (2)), then, by the previous estimates, the first term of (6) is . On the other hand, for every ,
therefore
from which one can deduce, by induction on , that for any ,
(7) |
By (5), the fact that , and the fact that the maps are ,
Now write . It follows that . Then, making in (7) and dividing it by , we get that
(8) |
Let be the largest value such that . Thus and, again by induction on , for . Indeed, if it is true for all , then
if is sufficiently small. Plugging this estimate for with in (8) again yields for if is sufficiently small. We also know that and hence for . Hence the second term in (6) is .
We are left with controlling the difference for . Notice that if , then the four points , , and belong to the same piece where . Thus,
This way, if is small enough so that the segments joining to and to belong to the domain of ,
Write
Notice that and are both conformal matrices. This happens because and belong, for every , to the domain in which is holomorphic, provided is sufficiently small. Besides, the difference between these two integrals is , because is and
Furthermore, by (4) and so
respectively. This implies that there exists some constant independent of such that
since the matrices , , , and commute (they are all conformal). Therefore, defining , for
For the minimum above is equal to , which decays exponentially. Up to such a value, the formula above implies that . Using the fact shown before that and choosing sufficiently small, it follows that the sequence is . Making we conclude (iii).
(iv) By lemma 2.12,
and the analogous relation is valid for the perturbed version. To show that it is thus sufficient to prove that
Notice that by (iii),
and so, as these diameters are uniformly bounded away from zero,
This way, we are left with analysing the derivatives of the affine maps and . Also from lemma 2.12,
To meet our objectives we need only to show that for all
(9) |
To each let be the largest integer such that . The analysis of in (iii) implies that, uniformly on ,
for all . This also implies that for all
Let us now show that these estimates remain valid for a much larger value of , that is, when is sufficiently small. Define for each and
Notice that for , the points and are always on the same piece , with , because of the definition of this number. Thus, in a neighborhood of , is just composed with a translation (see (4)), and therefore
because of (i) and the fact that the are with uniformly bounded derivatives. It follows that for every such that
Hence for all , because for all by the discussion above. Moreover, since , the same estimations imply that for all (this is the only part we use ).
For , let
Following this notation,
and the analogous relation is valid for the perturbed versions. However, remember that there is such that . This geometric control implies that there is a positive constant such that
for every . Since , if is sufficiently small, then , and so the estimate (9) is valid for all , concluding the first part of the proof.
∎
Remark 3.3.
Some remarks relating the perturbation:
-
•
Note that since then, supposing is small enough, we have .
-
•
Remember that we assume the base points , , are pre-periodic. From this, it is easy to prove that the base points do not depend on , i.e. . Indeed, let be the symbolic sequence associated to the points . Then we can write
where is any element in , and is the last letter of . Notice that if is big enough then contains a word of repeated twice, thus any containing can not be in . This implies that
for all . Making go to infinity we conclude that does not depend on .
-
•
If then and using
one sees that and then . On can arrive to a similar estimate if , in this case we can decompose as a concatenation of at most words in and use the fact that . Notice that with this approach the constants get worse if we increase the power of in which we are interested, will be enough for us.
-
•
Let and , suppose that
for all and all in a neighborhood of , and in a neighborhood of . Remember that limit geometries are defined by , where
By our assumption we have , for all and in a neighborhood of . From the proof of the existence of limit geometries (see [6]) one has that there is a constant such that
and
the same constant works for all Cantor sets , since they depend continuously on . It follows easily that there is a constant such that
and
for all in a neighborhood of . Notice that since , , and then .
4. Proof of theorem 2.22
In this section we will define the set of relative configurations , which will be a recurrent compact set for at least one of the Cantor sets in the family of random perturbations. We first give a primary description of and prove that assuming a probabilistic estimate, proposition 4.1, then we can prove theorem 2.22. The proof of the probabilistic estimate will be given in later sections.
4.1. The recurrent compact set
The set will depend on , but only the translation coordinate . The image of under the projection map: will be a subset of independent of .
We will choose a subset of with good combinatorial properties, this will be crucial to prove the estimate of proposition 4.1. First, let be the subset of formed by words such that:
-
(1)
no word appears twice in ;
-
(2)
if appears at the end of , then it does not appear elsewhere in .
We next define as the subset of formed by which end with a word in . This is an open and closed subset in .
A family of subsets of , for will be constructed in subsection 4.2, in relation to Marstrand’s theorem, and it will satisfy the hypothesis
Then, the Scale recurrence Lemma gives us another family , , with the properties indicated in the statement of the lemma.
The set is defined to be the subset of formed by the such that , and there exists , with and , ending with , respectively.
For every in , we will define in subsection 4.3, considering the properties given by Marstrand’s theorem, a non empty subset , depending on , of the fiber of over .
Let
consider next the -neighbourhood of in :
Fix . We define two subsets , of . First,
Second, is the set of such that there exists , , with , and the image ) satisfies:
-
(i)
for any with , we have ;
-
(ii)
.
The following crucial estimate will be proven in section 6.
Proposition 4.1.
Assume that is chosen conveniently large. Then there exists , such that, for any , one has
Using the previous estimate, we will prove that for some the pair has a recurrent compact set, thus obtaining theorem 2.22. We proceed as in [1]. We discretize the set , i.e. we choose subsets , , such that
and for all one has
It is not difficult to see that this can be done in such a way that is polynomial in . For each , choose . If is small enough, we have
Therefore, the set is not the whole and we can choose outside of it. Observe that for all we have that implies . Now define
We will prove that is a recurrent compact set for . First, notice that . Then is not empty. Now, let , we have that there is with the properties in the definition of . Since then (this is clear from the definition of in subsection 4.3 and equation (3)). Thus and for some . For we have that
Therefore and , this implies that and there is some pair such that satisfies the properties (i) and (ii) described above.
We will prove that is in the interior of . Write , using equation (3) we have
Therefore . Analogously one has
In this case we get . One also has and .
Thanks to property (ii), we know that . Moreover, for any such that
we have
To conclude that we only need to show that . Property (i) above implies that , by the definition of the set this means that and for a pair in such that ends in it. However, since then and also ends in . Therefore and , which shows that is in the interior of . From the fact that the sets are closed and the definition of the sets , it is not difficult to prove that is a compact set. Therefore, is a recurrent compact set for the pair .
4.2. Set of good scales
In this subsection we will define the sets which we use to construct . Let in the space of relative scales, and points , . Consider
Then is the unique relative configuration above such that
(where represents this relative configuration).
Remember that, for some previously fixed (given by the scale recurrence lemma),
and .
Let , be the Hausdorff dimension of , , respectively. We equip each set (resp. with the -dimensional (resp. -dimensional) Hausdorff measure (resp. ).
Then, for , we denote by the image under of on .
As in the theory of Cantor sets in the real line, there are constants such that, for , :
This can be proven using the results appearing in Zamudio’s thesis ([8]). Indeed, for a given conformal Cantor set of dimension , using equation (1), one can find a sequence of coverings with size converging to zero and -volume bounded by , namely the covering by the pieces of , showing that .
On the other hand, lemma 1.2.3 of [8] gives that there exist constants , , , independent of , such that for all
for all , . Using this lemma, we conclude that given a finite cover of by balls of radii , , each intersects at most pieces of if and are sufficiently small. Since is a cover and , summing for all yields:
and so (and ) is always bounded from zero. To obtain the statement just restrict the arguments to and and take their product.
Notice that the same lemma 1.2.3 implies that there is a constant such that for , the product measure in ,
for any ball of radius . If this condition implies that:
This way, the proof of Theorem 9.7 in Mattila’s book (a Marstrand-type theorem) [11] can be adapted444all one needs to verify is that for any points , , where is some constant depending only on . Notice that , where for and are diffeomorphisms that distort area in a uniformly bounded way. A simple manipulation shows that the measure is bounded above by . If then for some constant and the desired inequality follows choosing big enough. If , using that is in a bounded set, one sees that is bounded by , for some constant depending on . to our context to show that for fixed the measure is absolutely continuous with respect to the Lebesgue measure for Lebesgue almost every , with density in satisfying
where is independent of , .
When one controls , this gives, by Cauchy-Schwarz inequality, a lower bound for the Lebesgue measure of , being a subset of with positive -dimensional Hausdorff measure; indeed we have:
and therefore
(10) |
Fix in . Let , , with , . One has
and
We therefore have
with independent of , , , . On the other hand, one has
We conclude that
We now define, with conveniently large to be determined later:
For , , we define as the set of such that there exists , ending respectively with , such that .
One has, for any , :
therefore, provided that
we will have
for all , . This means that we can apply the Scale recurrence Lemma with the family of subsets of . The sets are then defined using this lemma (see section 2.7), we can assume they are closed, this is justified in the remark after the lemma.
4.3. Construction of
We now consider the family of random perturbations again and proceed to construct the sets , for . For , let be the open and closed subset of formed by the ending with . Choose a subset of such that
is a partition of .
For , define a subset of the subset (recall ), as the set of words in starting with . For , we also define .
Let . We write
and for such an , we set | ||||
This depends on , but nearby (with ) will belong to the same and give the same projection of .
For , the set will actually only depend (as far as is concerned) on the projection of associated to .
We will say that two words are independent if there is no word such that both and start with .
With conveniently small, to be chosen in the following, let
Let and .
We define to be the set of points in the fiber for which there exist pairs in , with , such that, if we set
the following hold:
-
(i)
the words are pairwise independent;
-
(ii)
for , ;
-
(iii)
for , and , ;
-
(iv)
for , .
We will use also a slightly smaller set ; it is defined in the same way than , but with (iii), (iv) replaced by:
-
(iii)’
for , and ,
-
(iv)’
for , .
In the next section, we will prove the following estimate.
Proposition 4.2.
If has been chosen sufficiently small, there exists such that, for any and any , the Lebesgue measure of is .
5. Proof of proposition 4.2
In this section we will prove proposition 4.2. First we prove some lemmas that are necessary for the proposition. We follow the same argument as [1] sections 4.8-4.12 with some modifications.
Fix , we will work with this triple throughout this section and at the end we will prove that .
Choose a subfamily of of words starting with such that
is a partition of . Similarly, choose a subfamily of of words starting with such that
There is a constant such that, for each , we have
Let and for and . Since is bounded above and below, we have
(11) |
if is sufficiently large. We assume that the previous relations involving hold for any other triples in and for any value of , choosing large enough this can be easily guaranteed.
Say is good if there are no more than pairs such that the distance between the points and is less than . Otherwise, say it is bad.
Lemma 5.1.
The number of bad pairs is less than .
Before begining the proof, we remember the Vitali covering lemma. Let be a finite collection of balls. For , denote by the ball with same center as but having radius three times larger. The lemma states that there exists a subcollection of balls with the Vitali property, this is
-
•
The balls are pairwise disjoint and
-
•
The union is contained in .
In the case that the balls are subsets of the complex plane and have the same radius , one can see that every point is covered by no more than of the balls . Indeed, consider the ball centered at with radius . It contains all the balls such that cover . However, the balls are pairwise disjoint, and so there are no more than of them inside , otherwise they would overlap.
Proof.
By construction of , there exists , , with , , such that
Because of remark 2.6, the distance between the points and is of order for every and the same is true for their versions. Thus, if is sufficiently large, for each bad pair there are more than pairs satisfying
From now on, we denote as for any .
For each bad pair , consider the disk of radius and center at . Then the corresponding sets are subsets of . This way,
Let be the union of all the disks corresponding to bad pairs and be the number of these pairs. Choose a subcover of as in the Vitali lemma, indexed by the pairs belonging to a subset of the set of bad pairs. It follows that
On the other hand, by Cauchy-Schwartz theorem,
for every bad pair . But the Vitali covering covers each point at most times, so
It follows that , concluding the proof. ∎
Now, we construct the pairs amongst which the pairs of 4.2 must be looked for. We make the following observation:
Lemma 5.2.
Let . The number of words with , such that is as , uniformly in .
Remember that . It follows from conclusion (ii) of the Scale recurrence Lemma (lemma 2.26) and the last observation that we can find at least pairs such that, writing , we have:
-
•
;
-
•
.
As again belongs to , we can for each find at least pairs (with the first letter of , being the last one of , respectively), such that writing , we have
-
•
;
-
•
.
Concatenation of the , and gives a family of words in with at least elements.
We now consider the perturbed operators. In this case and by lemma 3.2 the distance between and is of order . Similarly one has
and again the distance between and is of order .
Now we fix such that , , and .
Lemma 5.3.
If has been chosen sufficiently small, there are at least pairs which are good and satisfy
and such that at least pairs start with .
Proof.
The proof is the same as in [1], bearing in mind the different but similar bound in the number of bad pairs. ∎
We call the pairs verifying the properties of the previous lemma excellent pairs. The following general lemma will be used later to estimate the measure of the union of the perturbed version of the sets .
Lemma 5.4.
Let , , be families of sets in such that for some , and :
-
•
, , .
-
•
, .
Then
and
Proof.
Notice that . Let be a subset of such that the balls have the Vitali property. Thus
where in the passage from the second to the third line we use the fact that the sets , , are disjoint. This proves the first inequality. For the second, use again Vitali to find a subset of such that the balls have the Vitali property. Then
∎
Lemma 5.5.
Let be an excellent pair. Consider all the pairs described above which begin with , and for each pair consider the corresponding pairs . Define the sets
Then
and the constant can be chosen to be independent of .
Proof.
First we make the following observation:
(12) |
where is an affine function with , and this holds for any , for some , and any .
For an excellent pair we consider the associated pairs and the sets
We will prove that the measure of is at least of the order . To do this, we use lemma 5.4 to see that the measures of the sets , and are of the same order. It is clear from equation (11) that the sets , and are all contained in, and contain, balls with radius of order , and centered at the points , and respectively. We remark that
we also have
given that , and . All this allows us to conclude that we can use lemma 5.4.
Now we can estimate the measure of . In the following lines of equations we will be using: lemma 5.4 for the first three lines, observation in equation (12) for the fifth and sixth line, equation (10) in the seventh line, in the last line we use that is an excellent pair and the fact that is of order .
We now consider the sets
We will estimate the measure of . Notice that and if we are able to prove that , for some constant , then we can use lemma 5.4 with being and being to conclude that
To prove that there is such we will proceed similarly to what we did when estimating . Note that and then there exists such that , , and
(13) |
The sets
are all contained in, and contain, balls with radius of order , and centered at the points
respectively. By lemma 3.2 we know that
on the other hand we also have , and , thus we can conclude that the distance between any two centers is of order less than . Therefore, we can apply lemma 5.4 taking as one of the families
and as another of these families.
Using the previous analysis together with equations (12), (10), (13) and the fact that the number of is a positive proportion of we obtain
This guarantees the existence of the desired constant and finishes the proof of the lemma. ∎
Now we can prove proposition 4.2. Consider the function
where means indicator function of the set and the sum is over all excellent pairs. We want to estimate the measure of the set
where is a constant defined in the following way. Suppose that we have two excellent pairs with the same first coordinate , and such that . Then
for some , . Thus
which shows that
This implies that if we fix , then the number of possible pairs such that is bounded by a uniform constant, independent of and , we denote this constant by (this last statement is a consequence of lemma 1.2.3 in [8]).
Notice that since then is supported in a ball of radius proportional to centered at , thus there is a constant such that
Now we estimate from above and from below. By lemmas 5.5 and 5.3 there is a constant such that
(14) |
Let and excellent pairs , such that . Remember that
where the notation means that, once we have chosen , we can choose any and will hold provided is small enough. With this in mind we obtain
Given that , are excellent, we conclude that there can be no more than excellent pairs intersecting at . Since we get that
We are ready to bound , using equation (14) and the previous estimates
and from this we get
We fix small enough such that .
To finish the proof of the proposition we will show that . Let , then there are at least excellent pairs , each one with an associated pair which starts with and such that . By the definition of , we can extract from this family of excellent pairs a subfamily , , such that all firsts coordinates are different. For we denote the associated pair by . We will prove the pairs have the properties necessary to guarantee that . Write
then:
-
(i)
Since all firsts coordinates of the excellent pairs are different we conclude that are pairwise independent.
-
(ii)
By the way in which was defined we get that all .
-
(iii)’
By the scale recurrence lemma we know that . We also know that , then, if is small enough, we have
We conclude that if (remember that for every there is such that ).
-
(iv)’
Given that , there is such that
Since we conclude that .
6. Proof of Proposition 4.1
In this section we will prove proposition 4.1. Given , remember the decomposition where and . Recall that the set is given by the words in starting with the same word, in , in which finishes. In the same way as in [1], one uses Fubini’s theorem to reduce the proof of proposition 4.1 to proving
(15) |
where we have fixed such that and , , is normalized Lebesgue measure in .
The fact that means that and there is for which
and . Notice that , moreover if and only if for any .
Next, means that there are pairs in such that if we set
then:
-
(i)
the words are pairwise independent;
-
(ii)
;
-
(iii)
if ;
-
(iv)
.
Let be the word in in which ends, for each define in given by the concatenation of and a word at the beginning of , in such a way that (this can be done since ). Notice that the independence of the words imply that the words are all different.
Now we consider the decomposition of as , where and is the component of corresponding to . We use again Fubini’s theorem to reduce the proof of equation (15) to a similar formula in a smaller space. For fixed, we will prove that the set of such that has measure .
To prove the desired inequality, we will prove that for each there is a set with positive measure such that whenever is in this set we have (no matter the value of , ). More precisely, we will prove that if is in this set then , verify that if we set
then:
-
(i)
if ;
-
(ii)
.
The first property can be easily obtained. In fact, we already know (from (iii) above) that if . Notice that since , then and the fiber of over is the same as the one over , thus we only need to estimate . Using that , , and lemma 3.2 one gets that , then choosing sufficiently small gives the desired property (for any value of ).
For the second property we choose such that (then ends with ) and such that it does not contain anywhere else (this is possible since ). Set
We will prove the following lemmas:
Lemma 6.1.
Once has been fixed, the number only depends on , not in or for . Moreover, if is big enough then there is a constant such that
where .
Lemma 6.2.
If then .
Proof of lemma 6.1:.
From equation (3) we know that
The dependency on is on the term . Let such that , note that does not contain the word , let be the last letter of , we have
(16) |
We will prove that, assuming is fixed, this expression only depends on and not in for or , and it depends in a very specific way. Remember, see remark 3.3, that the base points were chosen to be pre-periodic points and that in fact they do not depend on , i.e. . Notice that and , then and appears only once in . We will study the dependency on for the different terms in equation (16):
-
•
Let be a finite word at the end of strictly shorter than . It is easy to prove by induction that does not depend on . Indeed, suppose that and are two such consecutive words. Assume does not depend on , we have
If the word is shorter than some word in , then such that and contains a word in repeated (remember that is pre-periodic, with a uniform bounded period). Thus cannot belong to and using equation (4) one gets that does not depend on (in fact, in this case it does not depend on ).
If the word is longer than any word in then it begins with a word in which can not be . This implies that such that , and the word is not in . Thus does not depend on .
-
•
Let be the first two letters of and define by . Define , we already proved that does not depend on . We have
-
•
We now study . By definition
Using the same arguments as before we see that, since does not contain and is pre-periodic, does not depend on . This also happens for in a neighborhood of , hence is independent of . Again, the same arguments prove that if then . We conclude that
for all .
-
•
We now treat . One has that
where is the last letter in . Given that and is pre-periodic we conclude that and are independent of . On the other hand, given that and does not contain we obtain that does not depend on . We conclude that is independent of .
From the previous analysis we get that only depends on and not in for or . Moreover, we have
taking derivative respect to we get
Observe that since belongs to the Cantor set then the matrix is conformal. Moreover
therefore defines a holomorphic function, which we will denote by . From the previous formulas it is not difficult to see that .
Now we show that is uniformly bounded. We already know that , using this inequality together with , , , and equation (3) one gets that . Therefore, choosing big enough one guarantees that the image of contains (see lemma 6.3). Having chosen this way, the fact that and proposition 4.2 gives that there is a constant such that
∎
In the previous proof we used the following lemma, it is proven using standard arguments in analysis, for completeness we present the proof.
Lemma 6.3.
Let be on and such that for all . Given , if is big enough, depending on , then .
Proof.
Redefining as and taking as , one can assume . Since is and then is a local diffeomorphism, hence its image is open. Let be a point in which is closest to , thus for all . We can cover by a finite number of open sets such that in each one of these sets is a diffeomorphism onto its image. We can use this cover to lift curves in (In fact, one can prove that is a covering map). Consider given by . Denote by a lifting of , i.e. , such that . The curve is and , otherwise this would contradict the choice of . Therefore
If then and the desired result follows. ∎
Proof of lemma 6.2:.
Let , then there exists pairs , , such that if we write
where is obtained from by setting the value in the coordinates belonging to , then
-
(i)
,…, are pairwise disjoint.
-
(ii)
.
-
(iii)’
implies .
-
(iv)’
.
We will prove that . To do this, we will prove that if write
where is obtained from by setting the value in the coordinates belonging to , then
-
(iii)
implies .
-
(iv)
.
First, notice that by lemma 3.2 we have that . To obtain and we applied the same renormalizations, with the same limit geometries but with different values of the perturbation parameter, therefore . We conclude that, taking small enough, item (iii)’ implies item (iii).
Now, to prove that (iv)’ implies (iv) we need stronger estimates. We have
We will compare the corresponding terms:
-
•
Using that , one has
-
•
Notice that and only differ at their values in the coordinates . Using that one sees that is not contained in . Then, from the arguments used in lemma 6.1, we see that . Moreover, since ends in and then
for all such that is strictly shorter that . For the same reasons we also have
for all such that is strictly shorter than . Both equalities still hold for in a neighborhood of either of the points. Thus we can use remark 3.3 and lemma 3.2 to obtain that
-
•
Notice that and only differ in the values associated to the coordinates . We will use again. Consider
and
Since one gets and . Thus we only need to estimate . Remember which verified , write
then
Using the analysis and notation from lemma 6.1 we see that: does not depend on , for all (in particular and ). We also have
and . Therefore
Here we used that limit geometries are and the norm of can be uniformly bounded. Using that for any word that ends with the letter , one can conclude that
And from this
therefore and .
From the previous estimates we conclude that , then if is small enough (iv)’ implies (iv). ∎
References
- [1] C. Moreira and J.-C. Yoccoz. Stable intersections of regular cantor sets with large hausdorff dimensions. Annals of Mathematics, 154(1):45–96, 2001.
- [2] Jacob Palis and Floris Takens. Cycles and measure of bifurcation sets for two-dimensional diffeomorphisms. Inventiones mathematicae, 82(3):397–422, 1985.
- [3] Jacob Palis and Floris Takens. Hyperbolicity and the creation of homoclinic orbits. Annals of Mathematics, 125(2):337–374, 1987.
- [4] Gregery T. Buzzard. Infinitely many periodic attractors for holomorphic maps of 2 variables. Annals of Mathematics, 145(2):389–417, 1997.
- [5] C. G. Moreira and J.-C. Yoccoz. Tangences homoclines stables pour des ensembles hyperboliques de grande dimension fractale. Annales scientifiques de l’École Normale Supérieure, 4e série, 43(1):1–68, 2010.
- [6] H. Araújo and C. G. Moreira. Stable intersections of conformal Cantor sets. arXiv e-prints, page arXiv:1910.03715, Oct 2019.
- [7] C. G. Moreira and A. Zamudio. Scale Recurrence Lemma and Dimension Formula for Cantor Sets in the Complex Plane. arXiv e-prints, page arXiv:1911.03041, Nov 2019.
- [8] A. Zamudio. Complex Cantor sets: scale recurrence lemma and dimension formula. PhD thesis, Instituto Nacional de Matematica Pura e Aplicada, 2017.
- [9] J. Palis and F. Takens. Hyperbolicity and sensitive chaotic dynamics at homoclinic bifurcations: fractal dimensions and infinitely many attractors in dynamics. Number 35 in Cambridge Studies in Advanced Mathematics. Cambridge University Press, 1995.
- [10] Michael Shub. Global Stability of Dynamical Systems. Springer, 1987.
- [11] Pertti Mattila. Geometry of Sets and Measures in Euclidean Spaces: Fractals and Rectifiability. Cambridge Studies in Advanced Mathematics. Cambridge University Press, 1995.