Hausdorff Dimension, Lagrange and Markov Dynamical Spectra for Geometric Lorenz Attractors

By Carlos Gustavo T. Moreira, Maria José Pacifico, and Sergio Romaña Ibarra

Abstract

In this paper, we show that geometric Lorenz attractors have Hausdorff dimension strictly greater than . We use this result to show that for a “large” set of real functions, the Lagrange and Markov dynamical spectrum associated to these attractors has persistently nonempty interior.

1. Introduction

In 1963 the meteorologist E. Lorenz published in the Journal of Atmospheric Sciences Reference Lor63 an example of a parametrized polynomial system of differential equations,

as a very simplified model for thermal fluid convection, motivated by an attempt to understand the foundations of weather forecasting. Numerical simulations for an open neighborhood of the chosen parameters suggested that almost all points in phase space tend to a strange attractor, called the Lorenz attractor. However Lorenz’s equations proved to be very resistant to rigorous mathematical analysis and also presented very serious difficulties to rigorous numerical study.

A very successful approach was taken by Afraimovich, Bykov, and Shil’nikov Reference ABS77 and by Guckenheimer and Williams Reference GW79 independently: they constructed the so-called geometric Lorenz models for the behavior observed by Lorenz (see section 2 for a precise definition). These models are flows in three-dimensions for which one can rigorously prove the coexistence of an equilibrium point accumulated by regular orbits. Recall that a regular solution is an orbit where the flow does not vanish. Most remarkably, this attractor is robust: it cannot be destroyed by a small perturbation of the original flow. Taking into account that the divergence of the vector field induced by system Equation 1 is negative, it follows that the Lebesgue measure of the Lorenz attractor is zero. Henceforth, it is natural to ask about its Hausdorff dimension. Numerical experiments give that this value is approximately equal to 2.062 (cf. Reference Vis04) and also, for some parameter, the dimension of the physical invariant measure lies in the interval (cf. Reference GN16). In this paper we address the problem to prove that the Hausdorff dimension of a geometric Lorenz attractor is strictly greater that . In Reference AP83 and Reference Ste00, this dimension is characterized in terms of the pressure of the system and in terms of the Lyapunov exponents and the entropy with respect to a good invariant measure associated to the geometric model. But, in both cases, the authors prove that the Hausdorff dimension is greater than or equal to , but it is not necessarily strictly greater than . A first attempt to obtain the strict inequality was given in Reference ML08, where the authors achieve this result in the particular case that both branches of the unstable manifold of the equilibrium meet the stable manifold of the equilibrium. But this condition is quite strong and extremely unstable. One of our goals in this paper is to prove the strict inequality for the Hausdorff dimension for any geometric Lorenz attractor; see Figure 1. Thus, our first result is the following.

Theorem A.

The Hausdorff dimension of a geometric Lorenz attractor is strictly greater than .

To achieve this, since it is well known that the geometric Lorenz attractor is the suspension of a skew product map with contracting invariant leaves, defined in a cross-section, we start studying the one-dimensional map induced in the space of leaves. We are able to prove the existence of an increasing nested sequence of fat (Hausdorff dimension almost ) regular Cantor sets of the one-dimensional map (Theorem 1). This fact implies that the maximal invariant set for the skew product (or, to first return map associated to the flow) has Hausdorff dimension strictly greater than , and this, in its turn, implies that the Hausdorff dimension of a geometric Lorenz attractor is strictly greater than . In another words, Theorem A is a consequence of the following result.

Theorem 1.

There is an increasing family of regular Cantor sets for such that

The proof of this theorem, although nontrivial, is relatively elementary, and it combines techniques of several subjects of mathematics, such as ergodic theory, combinatorics, and dynamical systems (fractal geometry).

To announce the next goal of this paper, let us recall the classical notions of Lagrange and Markov spectra (see Reference CF89 for further explanation and details).

The Lagrange spectrum is a classical subset of the extended real line, related to Diophantine approximation. Given an irrational number , the first important result about upper bounds for Diophantine approximation is Dirichlet’s approximation theorem, stating that for all , has an infinite number of solutions .

Markov and Hurwitz improved this result by verifying that, for all irrational , the inequality has an infinite number of rational solutions , and is the best constant that works for all irrational numbers. Indeed, for , the gold number, Markov and Hurwitz also proved that, for every has a finite number of solutions in . Searching for better results for a fixed we are lead to define

Note that the results by Markov and Hurwitz imply that for all , and It can be proved that for almost every

We are interested in such that (which forms a set of Hausdorff dimension ).

Definition 1.

The Lagrange spectrum is the image of the map :

In , Perron gave an alternative expression for the map , as below. Write in continued fractions: . For each , define

Then

For a proof of equation Equation 2 see, for instance, Reference CM, Proposition 21.

Markov proved (Reference Mar80) that the initial part of the Lagrange spectrum is discrete: with for all .

In , Hall proved (Reference Hal47) that the regular Cantor set of the real numbers in in whose continued fraction only appear coefficients satisfies Using expression Equation 2 and this result by Hall it follows that . That is, the Lagrange spectrum contains a whole half-line, nowadays called a Hall’s ray.

Here we point out is a horseshoe for a local diffeomorphism related to the Gauss map, which has Hausdorff dimension . Hall’s result says that its image under the projection contains an interval. This is a key point to get nonempty interior in . In , Freiman proved (Reference Fre75) some difficult results showing that the arithmetic sum of certain (regular) Cantor sets, related to continued fractions, contain intervals, and he used them to determine the precise beginning of Hall’s ray (the biggest half-line contained in ) which is

Another interesting set related to Diophantine approximation is the classical Markov spectrum defined by

Notably, the Lagrange and Markov spectrum have a dynamical interpretation. Indeed, the expression of the map in terms of the continued fraction expression of given in Equation 2 allows one to characterize the Lagrange and Markov spectrum in terms of a shift map in a proper space. Let be the set of bi-infinite sequences of integer numbers and consider the shift map , and define

where and .

The Lagrange and the Markov spectra are characterized as (cf. Reference CF89 for more details)

These characterizations lead naturally to a natural extension of these concepts to the context of dynamical systems.

For our purposes, let us consider a more general definition of the Lagrange and Markov spectra. Let be a smooth manifold, let or , and let be a discrete-time () or continuous-time () smooth dynamical system on ; that is, are smooth diffeomorphisms, , and for all .

Given a compact invariant subset and a function , we define the dynamical Markov (resp., Lagrange) spectrum (resp., ) as

where

It can be proved that (cf. Reference RM17). In the discrete case, we refer to Reference RM17, where it was proved that for typical hyperbolic dynamics (with Hausdorff dimension greater than ), the Lagrange and Markov dynamical spectra have nonempty interior for typical functions.

Moreira and Romaña also proved that Markov and Lagrange dynamical spectra associated to generic Anosov flows (including generic geodesic flows of surfaces of negative curvature) typically have nonempty interior (see Reference RM15 and Reference Rom16 for more details).

Now we are ready to state our next result. Let be the vector field that defines a geometric Lorenz attractor , and let be an open neighborhood of where is defined.

Theorem B.

Let be the geometric Lorenz attractor associated to . Then arbitrarily close to , there are a flow and a neighborhood of such that, if denotes the geometric Lorenz attractor associated to , there is an open and dense set such that for all , we have

where denotes the interior of .

1.1. Organization of the text

This paper is organized as follows. In Section 2, we describe informally the construction of a geometric Lorenz attractor and announce the main proprieties used in the text. In Section 3 we prove the first main result in this paper, Theorem 1 and its consequences, Corollary C and Theorem A. In Section 4 we proof our last result, Theorem B.

2. Preliminary results: geometrical Lorenz model

In this section we present informally the construction of the geometric Lorenz attractor, following Reference GP10Reference AP10, where the interested reader can find a detailed exposition of this construction.

Let be a vector field in the cube , with a singularity at the origin . Suppose the eigenvalues , satisfy the relations

Consider and , and , with .

Assume that is a transverse section to the flow so that every trajectory eventually crosses in the direction of the negative axis as in Figure 2. Consider also and put . For each the time such that is given by , which depends on only and is such that when . Hence we get (where for

Let be given by

It is easy to see that has the shape of a triangle without the vertex , which are cusps points of the boundary of each of these sets. From now on we denote by the closure of . Note that each line segment is taken to another line segment as sketched in Figure 2. Outside the cube, to imitate the random turns of a regular orbit around the origin and obtain a butterfly shape for our flow, we let the flow return to the cross section through a flow described by a suitable composition of a rotation , an expansion , and a translation . Note that these transformations take line segments into line segments as shown in Figure 2, and so does the composition This composition of linear maps describes a vector field in a region outside , such that the time-one map of the associated flow realizes as a map . We note that the flow on the attractor we are constructing will pass though the region between and in a relatively small time with respect the linearized region.

The above construction enables us to describe, for , the orbit for all : the orbit starts following the linear flow until and then it will follow coming back to and so on. Now observe that and so the orbit of all converges to . Let us denote by the set where this flow acts. The geometric Lorenz flow is the couple and the geometric Lorenz attractor is the set

where is the Poincaré map.

Composing the expression in Equation 4 with , , and and taking into account that points in are contained in , we can write an explicit formula for the Poincaré map by

and

where and are suitable affine maps, with and . Figure 3 displays the main features of and on and , respectively.

2.1. Properties of the one-dimensional map .

Here we specify the properties of the one-dimensional map described in Figure 3:

()

is discontinuous at with lateral limits and ;

()

is differentiable on and , where ;

()

the lateral limits of at are and .

The properties above imply another important feature for the map , as it is shown in Lemma 2.1 below. We will present the proof of R. Williams (cf. Reference Wil79, Proposition 1) for Lemma 2.1, which we will use to construct “almost locally eventually onto” avoiding the singularity of (cf. Section 2.3).

Lemma 2.1.

Put . If is a subinterval, then there is an integer such that . That is, f is locally eventually onto.

Proof.

Let , if ; otherwise let be the bigger of the two intervals into which splits . Similarly, for each such that is defined, set

Note that , where and denotes length. Thus unless is in both and , we have

But as , this last cannot always hold, say

Then contains and one end point of , so that is one “half” of . Note that contains the other half, and finally .

The next lemma gives us the following ergodic property for as above (Reference Via97, Corollary 3.4).

Lemma 2.2.

Let be a -function, satisfying properties in Section 2.1. Then has some absolutely continuous invariant probability measure (with respect to Lebesgue measure ). Moreover, if is any such measure, then where has bounded variation.

2.2. Properties of the map

By definition is piecewise and the following bounds on its partial derivatives hold:

(a)

For all (), we have . As and there is such that

(b)

For (), we have . Since and , we get .

We note that from the first item above it follows the uniform contraction of the foliation given by the lines The foliation is contracting in the following sense: there is a constant such that, for any given leaf of the foliation and for , then

We notice that the geometric Lorenz attractor constructed above is robust, that is, it persists for all nearby vector fields. More precisely, there exists a neighborhood in containing the attracting set , such that for all vector fields which are -close to , the maximal invariant subset in , , is still a transitive -invariant set. This is a consequence of the domination of the contraction along the -direction over the expansion along the -direction (see, e.g., Reference AP10, Session 3.3.4). Moreover, for every that is -close to , the associated Poincaré map preserves a contracting foliation with leaves. It can be shown that the holonomies along the leaves are in fact Hölder-; see Reference AP10. Moreover, if we have a strong dissipative condition on the equilibrium , that is, if for some (see the definitions of as functions of the eigenvalues of in Equation 3), it can be shown that is a -smooth foliation Reference SV16, and so the holonomies along the leaves of are -maps. In particular, for strongly dissipative Lorenz attractors with , the one-dimensional quotient map is -smooth away from the singularity (cf. Reference SV16).

We finish this section noting that putting together the observations above and the results proved in Reference AP10, Section 3.3.4, we easily deduce the following result.

Proposition 1.

There is a neighborhood such that for all , if is the quotient map associated to the corresponding Poincaré map , then the properties from Section 2.1 are still valid. Moreover, there are constants uniformly on a -neighborhood of such that if is the continuation of obtained for the initial flow it holds that

Furthermore, condition ensures that has enough expansion to easily prove that every is locally eventually onto for all close to .

2.3. Almost locally eventually onto

In this section we shall use an argument similar to the one given in Lemma 2.1, to achieve a property of fundamental for the construction of the family of Cantor sets in Theorem 1. Roughly speaking, we shall prove the existence of a number arbitrarily close to one, depending only on , such that for any interval , we have the following:

(1)

an interval such that and with size equal to a fixed proportion of the size of .

(2)

a number such that the restriction , is a diffeomorphism, where . Moreover, we obtain a control on the distortion at each step , for all .

To do that, we start with an auxiliary result.

Lemma 2.3.

There is a constant such that for all interval such that and , then

Proof.

We denote and the preimage of in each branch of , that is, . Consider also the two preimages of , , , , in and , respectively. As and , then for some , and thus we get that some of the intervals is contained in . Thus, taking , we finish the proof.

Recall that . Now we consider a number satisfying

For an interval , we will use the number satisfying equation Equation 7 to define an interval avoiding the singularity and that is obtained by cutting a small part of with length . In this direction, we proceed as follows.

Given any interval , we denote by the subinterval of cutting an interval of size on the closest side to zero, that is,

Note that, if , then and if , . It is clear that and .

When an interval or has size large enough (), we define the subinterval of cutting an interval of size of in the side of the point ; in other words,

It is clear that both kinds of intervals, and , avoid the singularity.

Recall that and are the preimages of , that is, , . For the next lemma assume that and . So, for sufficiently close to , we have that and . Therefore and . Denote and the bigger of two parts into which splits and , respectively.

The next lemma says that if is sufficiently close to , then it is easy to determine and explicitly.

Lemma 2.4.

Keeping the notation of above, if is close enough to , then

Proof.

Since and , we need only to prove that

for sufficiently close to . Let us prove the left-hand inequality; the other one is analogous. Note that , then by definition of we have that . In particular, for all , it holds that . Now consider the number , which only depends of . So, we can taken sufficiently close to such that

and therefore we conclude that , as we wished.

From now on, we will denote .

To prove the next lemma, we use the same idea as in the proof of Lemma 2.1, to get control on the number of iterations required to increase the size of any interval avoiding the singularity in each step.

Lemma 2.5.

If is a subinterval, then there are a subinterval and an integer such that is a diffeomorphism such that , , and

Proof.

Given , let if ; otherwise let , where is the biggest connected component of . Similarly, for each such that is defined, set

Note that , where . Thus, unless is in both and , we have

But as , this last inequality cannot always hold. Let be the minimum number such that

Thus equation Equation 9 implies that satisfies the hypothesis of Lemma 2.3. Therefore, , and as (see equation Equation 7), we define the interval . To finish the proof of the lemma, we have to consider two cases, depending on the relative position of in the connected components of .

Case 1.

Assume that . Thus by definition of we have that for some . Moreover, as , then , therefore as , then arguing as in the proof of Lemma 2.3, we get that or , consequently since , then , which implies by Lemma 2.4 that or equivalently

In this case, we define the following sequence of intervals and . Hence, by construction, the interval satisfies

Therefore, we conclude that

So, taking , we have that is a diffeomorphism and it is easy to see that , . This concludes the proof of Case 1.

Case 2.

Assume that . Then, . Thus by the same argument as in Case 1, we have that , and by Lemma 2.4 we have or, equivalently, , then, . Note that for sufficiently close to , and therefore . To conclude our arguments, we note that by Lemma 2.4

In this case, we define the following sequence of intervals and . Then using a similar argument as in Case 1, we have that the interval satisfies the condition , , for . The proof of Case 2 is complete.

To finish the proof of lemma, it is only left to estimate . For this, note that in any case, by construction and since , the estimate required for follows immediately.

Remark 1.

Note that we also proved the estimate

where are given by Equation 8 above.

The next corollary will be a fundamental tool for the proof of Theorem 1, more specifically, see Claim 4 in the proof of Theorem 1.

Corollary 1.

Let be a sequence such that . Then if with , we have the following.

There is a constant such that .

There are constants and such that for each hold that , where is as in Lemma 2.5.

Proof.

If , Lemma 2.5 implies that

where , and we finish the proof of item (a).

Let be the interval given by Lemma 2.5. Then Remark 1 provides

where are defined in Equation 8. The construction of gives . Thus

The next step is to estimate the derivative of . For this purpose, we use the inequalities Equation 10 and Equation 6 which provides that

We take and . This concludes the proof.

3. Fat Cantor sets for and the proof of Theorem 1

The main goal in this section is to prove Theorem 1, that is, that there are infinitely many regular Cantor sets for the one-dimensional map associated to a geometric Lorenz attractor, with Hausdorff dimension () very close to .

Before we announce precisely this result, let us recall the definition of the Hausdorff dimension of a Cantor set and the notion of a regular Cantor set. We refer the reader to the book Reference PT93, Chapter 4 for a nice exposition of the main properties of these kinds of Cantor sets. We proceed as follows.

Let be a Cantor set, and let be a finite covering of by open intervals in . We define the diameter as the maximum of , where denotes the length of . Define Then the Hausdorff -measure of is

One can show that there is an unique real number, the Hausdorff dimension of , which we denote by , such that for , and for , .

Definition 2.

A dynamically defined (or regular) Cantor set is a Cantor set , together with the following:

(i)

disjoint compact intervals such that and the boundary of each is contained in ;

(ii)

there is a expanding map defined in a neighborhood of such that, for each , is the convex hull of a finite union of some of these intervals ; moreover, satisfies

for each and sufficiently big, ;

.

We say that is a Markov partition for and that is defined by .

A classical example of regular Cantor set in is the ternary Cantor set of the elements of which can be written in base using only digits and . The set is a regular Cantor set, defined by the map given by

There is a class of examples of regular Cantor sets, given by a nontrivial basic set associated to a -diffeomorphism of a -manifold , which appear in the proof of Corollary A. Recall that a basic set is a compact hyperbolic invariant transitive set of which coincides with the maximal invariant set in a neighborhood of it. “Nontrivial” means that it does not consist of finitely many periodic orbits. These types of regular Cantor sets, roughly speaking, are given by the intersections and , where and are the stable and unstable manifolds of , respectively. We denote by the stable Cantor set and by the unstable Cantor set (cf. Reference PT93, chap. 4 or Reference RM15, Appendix).

If is a basic set associated to a -diffeomorphism defined in a surface, then it is locally the product of two regular Cantor sets and (cf. Reference PT93, Appendix 2). We shall use the following properties of a regular Cantor set, whose proofs can be found in Reference PT93:

Proposition 2 (Reference PT93, Proposition   4).

The Hausdorff dimension of a basic set satisfies

Proposition 3 (Reference PT93, Proposition   7).

If is a regular Cantor set, then

Proof of Theorem 1.

We construct the Cantor sets inductively. Denote and , and pick any interval . Lemma 2.1 implies that there is an iterate of such that is a diffeomorphism. Let be the complementary intervals in of . Again Lemma 2.1 implies that there are , , , and such that is a diffeomorphism.

Let be the complementary intervals of in , and let be the complementary intervals in .

Continuing with this process, in the th step, we obtain intervals such that, for each , there is so that is a diffeomorphism.

Now let be the complementary intervals of in , and let be an the invariant measure given by Lemma 2.2, which is absolutely continuous w.r.t. Lebesgue, and thus there is a constant such that

for any interval . Take , and put the integer part of , that is, .

Next, split each interval in intervals that are pairwise disjoint of equal -size. Then, for , we have

Consider the interval . Since is -invariant, inequality Equation 11 implies that

In what follows, given denotes the cardinality of .

Claim 1.

For any and any there is a set with , such that for each there is a point such that

Proof.

The idea of the proof is to count the number of intervals that does not satisfy this property. To do that, consider the set

We want show that . For this we proceed as follows. Put with , and let . Then, by the definition of , we obtain for all . Hence, equations Equation 12 and Equation 13 imply that

Hence we have

which implies that if is large enough (), then should be bigger than , i.e., .

Now implies that , and as , we get

Thus , and this concludes the proof of Claim 1.

Claim 2.

Consider the set . Then .

Proof.

As the intervals are pairwise disjoint, if , then

which implies a contradiction for large enough.

The above claim ensures that the set has cardinality .

Claim 3.

For all there is minimal, such that

Proof.

Let , then if for some , we are done. Otherwise, assume that there is such that for all . If for some , Claim 1 implies that there is such that for , and so we get , contradicting our hypothesis. Thus for all . Since acts as a diffeomorphism on with derivative and equation Equation 12 holds, we obtain

If , then we are done. Otherwise, if , reasoning as before, we get that . Since , then acts as a diffeomorphism on with derivative , which allows us to state that . Again, if , we are done. Otherwise, if and, reasoning as before, we get that , then

Using this argument recursively, if , then and it holds that

and

Thanks to inequality Equation 12 we conclude that

finishing the proof of Claim 3.

Now consider the sequence of intervals given by Claim 3. Since for all , we can apply Lemma 2.5 and Corollary 1 to get the following.

Claim 4.

For all , there is an interval and an integer such that is a diffeomorphism, for , , and .

Claim 5.

Let with , where is as in Claim 4. Then, there is a constant , depending only of , such that

Proof.

First note that the mean value theorem implies

It is enough to bound , since equality Equation 16 implies that inequality Equation 15 holds. For this reason, we proceed as follows. As is minimal satisfying Equation 14, we get

This implies, reasoning as in the proof of Claim 3, that for and hence is a diffeomorphism for .

Observe that by Claim 1, for each , there is such that , and so, if is the distance between sets, by equation Equation 17, we conclude that

Now we have

Recall that equation Equation 6 implies that , with , depending only of . Thus, since is a diffeomorphism for each , and satisfies property (see Section 2.1), we get

Making the change of variable , the last inequality provides

Using the inequality Equation 20 together with Equation 17 and replacing in the last term of equation Equation 19, we get that

Setting , we bound , and inequality Equation 16 follows, implying that inequality Equation 15 holds. The proof of Claim 5 is finished.

The next step is to construct the regular Cantor with Hausdorff dimension close to . For this sake, we consider the collection of surjective maps

Let be defined by , and let be the regular Cantor set defined by the intervals and , i.e.,

The final step is to show that as . For this, we use the same strategy given in Reference PT93, Theorem 3. In fact, consider the number

and define by

It is shown in Reference PT93, pp. 69–70 that . Therefore, we can estimate by computing . To do that, note that , and to simplify notation, denote and . Then

Corollary 1 gives that , where . To estimate the the supremum of , , in , we note that by the proof of Claim 5 the function has bounded distortion. Thus

where (see equation Equation 21). Since , the mean value theorem implies

The last two inequalities imply that

since

Therefore, since , inequalities Equation 22 and Equation 23 imply that

Hence

Thus, . Now we define which satisfies the condition of the theorem, finishing the proof of Theorem 1.

As an immediate consequence of Theorem 1, we have the following.

Corollary C.

The Hausdorff dimension of the bidimensional attractor for the Poincaré map , , is strictly greater than .

Proof.

For this, let , and let

be as in equation Equation 5. For each , let be the regular Cantor set given by Theorem 1 and define

Notice that by construction, each is a regular Cantor set (see comments after Definition 2) and so, for each , is a basic set for . Moreover, , and by Proposition 2

where and are the stable and unstable Cantor sets associated to the basic set As is a regular Cantor set, by Proposition 3, there is such that . Hence

Thus, Theorem 1 implies that for large enough. Since , this finishes the proof of Corollary C.

Proof of Theorem A.

Note that the geometric Lorenz attractor satisfies

Thus,

The proof of Theorem A is complete.

We finish this section by announcing a corollary of the proof of Theorem 1 that might be of interest to the reader.

Corollary D.

If is a -function that satisfies the properties described in Section 2.1 with , and also satisfies equation Equation 6, then there is an increasing family of regular Cantor sets for such that

4. Lagrange and Markov spectra: proof of Theorem B

In this section we prove Theorem B. For this, we first prove that small perturbations of the Poincaré map restricted to , with defined at Equation 24, can be realized as Poincaré maps of small perturbations of the initial geometric Lorenz flow (Lemma 4.1). Then, taking such that , we recover the properties described in Reference RM17 needed to apply Reference RM17, Main Theorem, obtaining nonempty interior in the Lagrange and Markov spectrum.

We start by giving the main theorem from Reference RM17, which is a fundamental tool for obtaining Theorem B. Given , denotes the interior of .

Theorem (Main Theorem at Reference RM17).

Let be a horseshoe associated to a -diffeomorphism such that . Then there is, arbitrarily close to , a diffeomorphism and a -neighborhood of such that, if denotes the continuation of associated to , there is an open and dense set such that for all , we have

The set is described by

where is the set of maximum points of in and are unit vectors in , respectively.

4.1. Perturbations of the Poincaré map

Fix with . By construction, there is small so that , where . Let be a -neighborhood of such that, if and is the hyperbolic continuation of , then .

The next lemma states that in a neighborhood of , we can recover as a Poincaré map associated to a geometric Lorenz flow , -close to .

Lemma 4.1.

Given , there is a geometric Lorenz flow , -close to , such that the restriction to of the Poincaré map associated to coincides with the restriction of to .

Proof.

For the proof we construct explicitly a flow , with the desired properties. For this, we proceed as follows. Let be a Markov partition of , and let be an open set with , for all , and such that if , then . The tubular flow theorem applied to gives local charts for satisfying

Put . Without loss of generality, we can assume that

We denote by and the maps and , respectively, in these coordinates.

Let be a -bump function such that for and for . Define the following flow on :

Note that

Consider the vector field on given by

By equation Equation 26, this vector field satisfies

Let be the vector field on defined by

By equations Equation 25 and Equation 28 we get that

Let be the open set , and consider the vector field given by . Finally, define the vector field by

Since , equation Equation 27 implies that is -close to . If is the flow associated to the vector field , equations Equation 26 and Equation 29 imply that the Poincaré map associated to restricted to is equal to restricted to . To finish the proof, note that for all , and thus, is a geometric Lorenz flow, as desired.

4.2. Regaining the spectrum

Recall that we are interested in studying the spectrum over a geometric Lorenz attractor , that is not a hyperbolic set, as well as . Thus, we cannot directly apply the techniques developed in the hyperbolic setting to analyze the spectrum in this case. So, the strategy we adopt is to profit from the fact that contains hyperbolic sets for the Poincaré map with Hausdorff dimension bigger than . Then we use similar arguments developed in Reference RM17 to show that the Lagrange and Markov dynamical spectrum has nonempty interior for a set of -real functions over the cross-section and with these functions regaining the spectrum over . In this direction, we proceed as follows.

The dynamical Lagrange and Markov spectra of and are related in the following way. Given a function , , let us denote by the function

where is the domain of and is such that and is a neighborhood of as in Theorem B.

Remark 2.

The map might be not in general.

For all , we have

In particular, if , we get

Remark 3.

It is worth noting that, given a vector field close to , the flow of still defines a Poincaré map defined in the same cross-sections where is defined.

Thus, the last equality reduces Theorem B to the following statement:

Theorem 4.1.

In the setting of Theorem B arbitrarily close to , there is an open set of -vector fields defined on such that for every there is a -open and dense subset , such that