On multifractal analysis and large deviations
of singular hyperbolic attractors
Abstract
In this paper we study the multifractal analysis and large derivations for singular hyperbolic attractors, including the geometric Lorenz attractors. For each singular hyperbolic homoclinic class whose periodic orbits are all homoclinically related and such that the space of ergodic probability measures is connected, we prove that: (i) level sets associated to continuous observables are dense in the homoclinic class and satisfy a variational principle; (ii) irregular sets are either empty or are Baire generic and carry full topological entropy. The assumptions are satisfied by -generic singular hyperbolic attractors and -generic geometric Lorenz attractors . Finally we prove level-2 large deviations bounds for weak Gibbs measures, which provide a large deviations principle in the special case of Gibbs measures. The main technique we apply is the horseshoe approximation property.
1 Introduction
Ergodic theorems appear as cornerstones in ergodic theory and dynamical systems, as they allow to describe long time behavior of points in full measure sets with respect to invariant probability measures. Given this starting point, a particularly important topic of interest is to characterize level sets, the velocity of convergence to time averages and the set of points for which time averages do not exist, often called irregular points. In this paper we will be interested in the multifractal analysis and large deviations for flows with singularities, whose concepts we will recall.
Let , , denote the space of -vector fields on a closed smooth Riemannian manifold endowed with the -topology. Given a vector field and a compact invariant subset of the -flow generated by , we denote by the space of continuous functions on . For any , Birkhoff’s ergodic theorem ensures that for any , the time average exists for -almost all point . Defining, for each , the -level set
and the -irregular set by
one obtains the multifractal decomposition
of the flow with respect to the observable . The properties of the entropy, dimension or genericity of level sets and irregular sets have been much studied in the recent years. For uniformly hyperbolic systems, both diffeomorphisms and vector fields, a rigorous mathematical theory of multifractal analysis is available (see [9, 8, 10, 23, 39, 40, 56] and references therein). In rough terms, for uniformly hyperbolic dynamical systems each level set carries all ergodic information and their topological entropy satisfies a variational principle using the invariant measures supported on it, while the irregular set carries full topological entropy and, under some conformality assumption, it has full Hausdorff dimension. The multifractal analysis of non-uniformly hyperbolic systems had a few contributions (see e.g. [7, 17, 41, 60]) but the theory still remains quite incomplete, especially in the time-continuous setting.
A second and related topic of interest concerns the theory of large deviations which, in the dynamical systems framework, addresses on the rates of convergence in the ergodic theorems. This gives a finer description of the behavior inside the level sets of the multifractal decomposition described above. More precisely, given a reference probability measure on (possibly non-invariant) one would like to provide sharp estimates for the -measure the deviation sets
for all and all real numbers . Large deviations in dynamical systems are often used to measure the velocity of convergence to a certain probability. For uniformly hyperbolic and certain non-uniformly hyperbolic systems, both diffeomorphisms and vector fields, large deviation have been well-studied (see e.g. [4, 12, 21, 31, 67] and references therein). From the technical viewpoint, in most situations large deviations principles often rely on the following mechanisms: (i) differentiability of the pressure function; (ii) some gluing orbit property and weak Gibbs estimates; (iii) existence of Young towers with exponential tails modeling the dynamical system; or (iv) entropy-denseness on the approximation by horseshoes. We refer the reader to [18, 36, 47, 61] for more details on each of these approaches.
In this paper we are interested in the multifractal analysis and large deviations of vector fields with singularities, including geometric Lorenz attractors (cf. Definition 2.5). The Lorenz attractor was observed by E. Lorenz [34] in 1963, whose dynamics sensitively depend on initial conditions. Later, J. Guckenheimer [26] and V. Aframovi-V. Bykov-L. Sil’nikov [3] introduced a geometric model for the Lorenz attractor, nowadays known as geometric Lorenz attractors. It is known that the space of vector fields () exhibiting a geometric Lorenz attractor is an open subset in (cf. [52]). In the study of robustly transitive flows, Morales, Pacífico and Pujals [38] introduced the concept of singular hyperbolic flows (see Definition 2.2), which include the geometric Lorenz attractors as a special class of examples. The coexistence of singular and regular behavior is known to present difficulties to both the geometric theory and ergodic theory of flows, and to present new and rich phenomena in comparison to uniformly hyperbolic flows. In order to illustrate this, let us mention that for each , there exist a Baire residual subset and a dense subset in such that for a geometric Lorenz attractor of , the space of ergodic probability measures supported on is connected, while for a geometric Lorenz attractor of , the space of ergodic probability measures supported on is not connected; and a similar statement holds for -singular hyperbolic attractors in higher dimension [52, Theorem A&B]. In rough terms, the underlying idea is that while the set
of probability measures which give zero weight to the singularity set of a singular hyperbolic attractor is formed by hyperbolic measures, which inherit a good approximation by periodic orbits, Dirac measures at singularities can be either approximated or not (in the weak∗ topology) by periodic measures depending on the recurrence of the singular set to itself, measured in terms of proximity to vector fields displaying homoclinic loops (see [52] for the precise statements).
Here we use the horseshoe approximation technique to study the multifractal analysis and large deviations for singular hyperbolic attractors, including geometric Lorenz attractors. The classical approach to describe level sets involves the uniqueness of equilibrium states for Hölder continuous observables . However this is still an open problem for singular hyperbolic attractors. On the other hand, if the observable is merely continuous one knows that uniqueness of equilibrium states fails even for subshifts of finite type [29]. Another obstruction appears when one considers the irregular set . Indeed, while most constructions of fractal sets with high entropy involve the use of some specification-like property, the presence of hyperbolic singularities constitutes an obstruction for specification (see e.g. [55, 65]). In this direction, a standard argument is to establish the variational principle for saturated sets of generic points. However, up to now it is still unknown whether an invariant non-ergodic probability measure (for example, convex sum of infinite periodic measures not supported on a same horseshoe) has generic points. Similar obstructions occur in the study of large deviations. The drawback of looking for the differentiability of the pressure function is that, even in the hyperbolic context, it demands one to consider the space of Hölder continuous observables. The latter relies ultimately on the uniqueness of equilibrium states for Hölder continuous potentials, a question which in such generality remains widely open. Moreover, singular hyperbolic attractors seem not to display any gluing orbit property, as hinted by [11, 55, 65]. We first show that the horseshoe approximation technique, valid for -generic geometric Lorenz attractors and -generic singular hyperbolic attractors, is enough to show that the level sets and irregular sets of such singular hyperbolic attractors inherit the properties of the corresponding objects for special classes of horseshoes approximating them. This property plays a crucial role in the proof of level-2 large deviations lower bounds for weak Gibbs measures, ie, lower bounds on the measure of points whose empirical measures belong to some weak∗ open set of probability measures on the attractor . On the other hand, level-2 large deviations upper bounds for weak Gibbs measures follow a more standard approach, exploring ideas from the proof of the variational principle. However, as one requires a very mild Gibbs property, the large deviations rate function takes into account certain tails of constants often associated to the loss of uniform hyperbolicity. We refer the reader to [61] for a discussion on the relation between the weak Gibbs property and the tail of hyperbolic times of local diffeomorphisms, and to Theorem 6.1 for the precise statements.
Organization of the paper
In Section 2, we present some concepts and known results. The main theorems (Theorem A,B,C) of this paper are presented in Section 3. In Section 4, we introduce the notions of entropy denseness and horseshoe approximation properties, and provide sufficient conditions for these properties to be verified. Section 5 is devoted to the multifractal analysis of Lorenz attractors/singular hyperbolic attractors and to the proof of Theorems A and B. Finally, in Section 6 we provide large deviation estimates for Lorenz attractors/singular hyperbolic attractors and prove Theorem C.
2 Preliminaries
Before the statements of the main theorems, one would like to give some preliminaries in this section. Let , , denote the space of -vector fields on a closed smooth Riemannian manifold . For , denote by or for simplicity the -flow generated by and denote by the tangent map of . Moreover, given any -invariant set we denote by the set of singularities for the vector field in . Let be the space of all probability measures supported on endowed with the weak*-topology. Let be a metric on the space compatible with the weak∗ topology which can be defined as follows, see for instance [62, Section 6.1]: take (and fix) a countable dense subset of where is not the zero function for every , and for any set
where denotes the -norm on . The set of invariant (resp. ergodic) probability measures of supported on is denoted by (resp. ). We denote by the metric entropy of the invariant probability measure , defined as the metric entropy of with respect to the time-1 map of the flow.
2.1 Geometric Lorenz attractor and singular hyperbolicity
We recall the definitions of hyperbolic and singular hyperbolic sets.
Definition 2.1.
Given a vector field , a compact -invariant set is hyperbolic if admits a continuous -invariant splitting , where denotes the one-dimensional linear space generated by the vector field, and (resp. ) is uniformly contracted (resp. expanded) by , that is to say, there exist constants and such that for any and any ,
-
•
, for any ; and
-
•
, for any .
for any and . If is transitive, ie it admits a dense orbit, then the dimension of the stable subbundle is constant and is called the index of the hyperbolic splitting.
The concept of singular hyperbolicity was introduced by Morales-Pacifico-Pujals [38] to describe the geometric structure of Lorenz attractors and these ideas were extended to higher dimensional cases in [32, 35]. More explorations can be found in for instance [50] which establishes SRB measures for singular hyperbolic attractors. Let us recall this notion.
Definition 2.2.
Given a vector field , a compact and invariant set is singular hyperbolic if it admits a continuous -invariant splitting and constants such that, for any and any ,
-
•
is a dominated splitting : , and
-
•
is uniformly contracted by : for any ;
-
•
is sectionally expanded by : for any 2-dimensional subspace .
Remark 2.3.
The following properties hold:
-
1.
Given , it follows from the definition of singular-hyperbolicity that all hyperbolic periodic orbits of a singular-hyperbolic set have the same index.
-
2.
Given and an invariant compact set which contains regular points (i.e. points that are not singularities), if is hyperbolic, then it must contain no singularities. On the other hand, if , then is hyperbolic if and only if is singular hyperbolic for or for (cf. [38]).
-
3.
Singular hyperbolicity is a -robust property. More precisely, if is a singular hyperbolic invariant compact set of associated with splitting and constants , then there exists an open neighborhood of and a neighborhood of such that for any , the maximal invariant set of in is a singular hyperbolic set for with the same stable dimension and constants (cf. [37, Proposition 1]).
We give the definition of homoclinic class of a hyperbolic periodic orbit and homoclinically related property between hyperbolic periodic orbits. First recall that for a hyperbolic periodic point of , the stable/unstable manifolds associated to its orbit is defined as follows:
The classical theory on invariant manifolds by [28] ensures that and are submanifolds of .
Definition 2.4.
Let and assume are two hyperbolic periodic points of . The homoclinic class of is defined as
that is, the closure of the set of transversal intersections between the stable and unstable manifolds of the periodic orbit of . The two hyperbolic periodic orbits and are homoclinically related if the stable manifold of has non-empty transverse intersections with the unstable manifold of and vice versa. A homoclinic class is called non-trivial if it is not reduced to a single hyperbolic periodic orbit.
Finally, we give the definition of geometric Lorenz attractors following Guckenheimer and Williams [26, 27, 66] for vector fields on a closed 3-manifold . Recall that for a hyperbolic singularity of a vector field , its local stable manifold of size is
When we do not specify the size of the local stable manifold, we just write as .
Definition 2.5.
We say () exhibits a geometric Lorenz attractor , if has an attracting region such that is a singular hyperbolic attractor and satisfies:
-
•
There exists a unique singularity with three exponents , which satisfy and .
-
•
admits a -smooth cross section which is -diffeomorphic to , such that , and for every , there exists such that .
-
•
Up to the previous identification, the Poincaré map is a skew-product map
Moreover, it satisfies
-
–
for , and for ;
-
–
, and ;
-
–
the one-dimensional quotient map is -smooth and satisfies , , and for every .
-
–

It has been proved that geometric Lorenz attractor is a homoclinic class [5, Theorem 6.8]. Moreover, for , vector fields exhibiting a geometric Lorenz attractor forms a -open set in [52, Proposition 4.7] .
Proposition 2.6.
Let and . If exhibits a geometric Lorenz attractor with attracting region , then is a singular hyperbolic homoclinic class, and every pair of periodic orbits are homoclinic related. Moreover, there exists a -neighborhood of in , such that for every , is an attracting region of , and the maximal invariant set is a geometric Lorenz attractor.
2.2 Topological entropy on general subsets
Let be a compact metric space. The topological entropy of a continuous flow on a compact invariant set can be defined as the topological entropy of its time-1 map on . When fails to be invariant or compact, the topological entropy can be defined using Carathéodory structures as in the discrete-time framework, see [45], also see [57, Section 2.1]. Let us be more precise. Fix an arbitrary set . For any , for any and any , consider the -Bowen ball
Now, given a finite or countable collections of Bowen balls which covers and a fixed , we define:
For , define
where the infimum is taken over all finite or countable collections of Bowen balls which covers such that and , for all . Since is non-decreasing with respect to , the limit
does exist. It is not hard to check that
is well defined. Then the topological entropy of with respect to the flow is
Note that similarly with the definition of the metric entropy with respect to a flow, the topological entropy of with respect to the flow equals to that with respect to its time-one map .
2.3 Topological pressure and weak Gibbs measure
In this subsection, we always assume and is an invariant compact set. Recall that we denote by the flow generated by . One first recalls the notion of topological pressure associated to a potential and equilibrium state.
Definition 2.7.
Given a continuous function , here called a potential, the topological pressure is defined as
An invariant probability measure attaining the previous supremum is called an equilibrium state of with respect to .
Now we define the notion of weak Gibbs measure associated to which we will study the large deviations bounds.
Definition 2.8.
Let and be an invariant compact set. Given a Hölder continuous potential , an invariant measure is called a weak Gibbs measure with respect to if there exists a full -measure subset and so that the following holds: for any , and , there exists a constant such that and
(2.1) |
for any dynamic Bowen ball .
Such a weak Gibbs property in Definition 2.8 holds for large classes of non-uniformly hyperbolic dynamical systems and conformal probability measures with Hölder continuous Jacobians, in which case may be chosen as the set of points with infinitely many instants of hyperbolicity (see e.g. [61] for more details).
Remark 2.9.
Consider the extension defined by for every . It is a standing assumption in the sense that, for each , the map is lower semi-continuous, i.e. for each the set
A special case of weak Gibbs property occurs when can be chosen as a constant independent of . In case the constants can be chosen independently of both and and the probability measure is called a Gibbs measure.
There is a strong connection between equilibrium state and the notion of Gibbs measures in the context of uniformly hyperbolic dynamical systems: the space of equilibrium states coincides with the space of Gibbs measures. Most importantly, the quantitative description of dynamic balls make Gibbs measures extremely useful to compute the speed of convergence in the ergodic theorem (see e.g. [4, 12, 21, 67]).
2.4 Suspension flows
Let be a homeomorphism on a compact metric space and consider a continuous roof function . We define the suspension space to be
where the equivalence relation identifies with , for all . Let denote the quotient map from to . We define the flow on the quotient space by
For any function , we associate the function by . Since the roof function is continuous, is continuous as long as is. Moreover, to each invariant probability we associate the measure given by
Observe that since is bounded away from zero, the measure is well-defined and -invariant, i.e. for all and measurable sets . Moreover, the map
is a bijection. Abramov’s theorem [2, 44] states that and hence, the topological entropy of the flow satisfies
Throughout we will use the notation for a flow on a compact metric space and for a suspension flow. Suspension flows are endowed with a natural metric, known as the Bowen-Walters metric (see e.g. [9]).
2.5 Specification properties
For completeness of the paper, we recall the notions of specification property introduced in [53] and the gluing orbit property.
Definition 2.10.
Let be a compact metric space and be a continuous map. We say satisfies the specification property if for any , there exists an integer such that the following holds: for any finite collection of intervals with and and for any points , there exists a point such that
The gluing orbit property was introduced in [11, 12, 59] (which was called transitive specification property in [59]) and is a weak version of specification property (see explanations for instance in [59, Page 553]).
Definition 2.11.
Let be a compact metric space and be a continuous map. We say satisfies the gluing orbit property if for any , there exists an integer such that the following holds: for any finite collections of points and integers , there exist integers and a point such that
where we set .
3 Statements of main theorems
3.1 Multifractal analysis
Recall that for a -vector field , we denote by the -flow generated by and by the tangent map of . We also use and for simplicity if there is no confusion. In the following, by we mean a -dimensional closed smooth Riemannian manifold. We use the symbol “” to denote the topological entropy of a set, see detailed definitions in Section 2.2. Recall that a subset of a topological space is residual if it contains a dense subset of . Our first main result ensures that the Birkhoff irregular points form a residual subset of geometric Lorenz attractors, and that level sets are typically dense and satisfy a relative variational principle. More precisely:
Theorem A.
There exists a Baire residual subset so that, if is a geometric Lorenz attractor of and , then either:
-
1.
is empty and for all , or
-
2.
is a residual subset of and .
Moreover, if is non-empty then for any satisfying
the level set is dense in and
Remark 3.1.
We consider the original geometric model introduced by Guckenheimer and Williams [26, 27] where it is required that the induced one-dimensional Lorenz map satisfies that , see Definition 2.5. This is because we have to use the eventually onto property of to obtain that all periodic orbits in the geometric Lorenz attractor are homoclinically related (Definition 2.4), see Proposition 2.6 and Corollary 4.14.
For singular hyperbolic attractors, one obtains the following result for -generic vector fields.
Theorem B.
There exists a Baire residual subset so that if is a singular hyperbolic attractor of and then either
-
1.
is empty and and for all , or
-
2.
is a residual subset of and .
Moreover, if is non-empty then, for any satisfying
the level set is dense in and
The proofs of Theorem A and B will be given in Section 5.2 after the proof of a technical theorem–Theorem 5.1.
Remark 3.2.
We give two remarks here.
-
1.
It is clear from Theorem A that, for typical geometric Lorenz attractors, is non-empty if and only if there exist so that , a property satisfied by a -open and dense set of continuous observables. In particular, if is the Baire residual subset given by Theorem A and is a geometric Lorenz attractor of then
is open and dense in . A similar conclusion holds in the context of Theorem B.
- 2.
3.2 Large deviations
Recall that denotes the space of all probability measures supported on endowed with the weak*-topology. Our next result is a level-2 large deviations principle, which detects exponential convergence to equilibrium on the space for a singular hyperbolic attractor or a Lorenz attractor . We need to set further notations. Given , let be the empirical measure function at time t, defined by
In other words, is the empirical probability on determined by the point at time . We say that an invariant probability measure satisfies a level-2 large deviations principle if there exists a lower-semicontinuous rate function so that
for any closed subset , and
for any open subset . If the latter holds then the contraction principle (see e.g. [22, Theorem 4.2.1]) ensures the level-1 large deviations principle
and
for each continuous observable .
Our next result establishes large deviations bounds for weak Gibbs measures, in which the rate function is expressed in terms of thermodynamic quantities. For a continuous potential , define by
When is singular hyperbolic, the entropy map is upper semi-continuous [42], therefore is lower semi-continuous.
Theorem C.
(Level-2 large deviations)
There exist a Baire residual subset and a Baire residual subset such that if is a Lorenz attractor of or is a singular hyperbolic attractor of , then the following properties are satisfied.
Assume is a weak Gibbs measure with respect to a Hölder continuous potential with being the -full measure set such that (2.1) satisfies.
Then one has:
-
1.
(upper bound) There exists so that
for any closed subset .
-
2.
(lower bound) If is an open set, then
-
3.
(lower bound for Gibbs measure) If is a Gibbs measure with respect to , then
for any open subset .
Some comments are in order. Theorem C can be compared to the local large deviations principle established by Rey-Bellet and Young [49]. The constant , defined by (6.6), measures tails of constants appearing in the concept of weak Gibbs measure. If is a (strong) Gibbs measure then: (i) and the upper bound in the theorem reduces to , and (ii) there is no need to restrict to probability measures supported on and the lower bound reduces to (item 3). The function satisfying Theorem C is called the large deviations rate function. Moreover, while large deviations for flows usually involve the use of local cross-sections and Poincaré maps, creating dynamical systems with non-compact phase spaces and discontinuities (see e.g.[4, 6]), our approach deals with the flows (and time-t maps) directly, hence it avoids creating such technical obstructions.
Remark 3.3.
We will prove two general results Theorem 5.1 and Theorem 6.1 stating that when is a singular hyperbolic homoclinic class such that each pair of periodic orbits are homoclinically related and , then the conclusions of Theorem A, B and C holds. Then Theorem A and B are direct consequences of Theorem 5.1 and C is a direct consequence of Theorem 6.1. The reason is that when is a Lorenz attractor of vector fields in a Baire residual subset or is a singular hyperbolic attractor of vector fields in a Baire residual set , then is a homoclinic class such that each pair of periodic orbits are homoclinically related (cf. [5, Theorem 6.8] for Lorenz attractors and [20, Theorem B] for singular hyperbolic attractors) and (cf. [52, Theorem A & B]). See also Proposition 2.6 and Corollary 4.14 in this paper.
4 Entropy denseness and the horseshoe approximation property
We give a criterion to study multifractal analysis and large deviations for flows beyond uniform hyperbolicity.
4.1 Entropy denseness of horseshoes
Recall that for each invariant compact set of a vector field , we denote by and the space of invariant and ergodic probability measures supported on , respectively, and that is a translation invariant metric on compatible with the weak∗ topology.
Definition 4.1.
Given and an invariant compact set . We say a convex subset is entropy-dense if for any and any , there exists satisfying
It follows from the definition that the entropy denseness property is hereditary, i.e. if are convex and is entropy dense, so is . In what follows we discuss some consequences of approximating invariant sets by horseshoes. In particular the strongest conclusion is entropy denseness of the entire space . We now recall the definition of horseshoe.
Definition 4.2.
Given , an invariant compact set is called a basic set if it is, transitive, hyperbolic and locally maximal, i. e. there exists an open neighborhood of , such that . A basic set is called a horseshoe if is a proper subset of , is not reduced to a single orbit of a hyperbolic critical element and its intersection with any local cross-section to the flow is totally disconnected.
Following the classical arguments of Bowen [13, 15] on Axiom A vector fields, every horseshoe is semi-conjugate to the suspension of a transitive subshift of finite type (SFT) with a continuous roof function through a finite-to-one continuous map. See also a detailed explanation in [9, Section 2.4]. We formulate this as the following theorem.
Theorem 4.3 (Bowen).
Assume is a horseshoe of . Then there exists a transitive subshift of finite type and a continuous roof function , such that is semi-conjugate to the suspension through a continuous surjective map , where is the flow generated by and is the suspension flow. Moreover, the semi-conjugacy is finite-to-one and hence preserves entropy:
-
1.
for any , there exists a unique satisfying and their metric entropies coincide ;
-
2.
for any invariant set of , it satisfies ;
In the 1970’s, Sigmund [53, 54] studied the space of invariant measures of basic sets for Axiom A systems. The following theorem is from [54, Theorems 2&3].
Theorem 4.4 (Sigmund).
Assume is a horseshoe of . The following properties are satisfied.
-
1.
For any , the set of -generic points is non-empty;
-
2.
There exists a residual subset such that each is ergodic and .
Remark.
Although Sigmund’s original statements were for basic sets, his arguments could be easily applied to horseshoes (isolated hyperbolic non-trivial transitive sets). See also remarks after [1, Theorem 3.5]
The following lemma states that two horseshoes are contained in a larger one once they are homoclinically related. The proof applies the -lemma [43, Lemma 7.1] (also see [64, Theorem 5.1]) and is classical, thus we omit it.
Lemma 4.5.
Let and be two horseshoes of . Assume there exists hyperbolic periodic points and such that and are homoclinically related. Then there exists a larger horseshoe that contains both and .
The next proposition ensures that horseshoes are entropy-dense.
Proposition 4.6.
Assume that . If is a horseshoe then is entropy-dense.
Proof.
Although the result is probably known we shall include a proof as we could not find a reference. Assume that is a horseshoe, by Theorem 4.3, it is semiconjugate to a suspension flow over a transitive subshift of finite type with a continuous roof function . As the semi-conjugacy preserves entropy, we may deal directly with the case of the suspension flow . Moreover, since the entropy map is upper-semicontinuous and satisfies the gluing orbit property (also known as transitive specification), Theorem B in [24] guarantees that is entropy-dense.
Fix and an arbitrary -invariant probability measure , determined by a -invariant probability measure (recall Subsection 2.4). Take small, to be determined later. Since the subshift of finite type is entropy-dense, one picks a -invariant and ergodic probability so that
-
•
is small enough such that
-
•
Then the Abramov formula for the -invariant probabilities and (recall Subsection 2.4) ensures that
provided that is small enough. Diminishing if necessary we may also get that and are also -close in the -metric. This completes the proof of the proposition. ∎
4.2 Horseshoe approximation property for singular hyperbolic homoclinic classes
We introduce a notion of horseshoe approximation property, a condition stronger than the entropy-denseness condition in Definition 4.1.
Definition 4.7.
Given and an invariant compact set , we say a convex subset has the horseshoe approximation property if for each and any , there exist a horseshoe and
(4.1) |
Remark 4.8.
By Proposition 4.6, if is a horseshoe, then admits the horseshoe approximation property naturally. Moreover, it is clear from (4.1) in the previous definition, that the horseshoe satisfies . Finally, in some specific contexts the approximating horseshoe can be constructed using an analogue of Katok’s arguments [30] and all measures supported on are within an -neighborhood (in the weak∗ topology) of the original probability (see e.g. Lemma 4.10).
The horseshoe approximation property will be essential in the technique to deal with multifractal analysis. We proceed to analyse the horseshoe approximation in the context of singular hyperbolic sets. Given a compact and invariant set of , denote by the set of periodic measures supported on and set
We first prove the following auxiliary lemma.
Lemma 4.9.
Let and be a singular hyperbolic homoclinic class of . If each pair of periodic orbits of are homoclinic related, then
where is the convex hull of .
Proof.
Given , let be an invariant compact set displaying a singular hyperbolic splitting . Then for any ergodic measure , the splitting is a dominated splitting and the index of equals obviously. Using Katok’s arguments in [30, Theorem 4.3] one obtains the following lemma. Given and assume is an invariant compact set. We say that satisfies the star condition on if there exist a neighborhood of and a -neighborhood of in such that every periodic orbit contained in associated to a vector field is hyperbolic.
Lemma 4.10.
Let and be a singular hyperbolic homoclinic class of . Assume each pair of periodic orbits of are homoclinic related, and . Then for any there exists a horseshoe contained in the -neighborhood of (in the Hausdorff distance), and so that for any , and there exists satisfying .
Proof.
We only give a sketch here since the proof is essentially contained in [33, Proposition 2.9] (see also [42, Theorem 4.1]). Note that since is a singular hyperbolic homoclinic class, the vector field satisfies the star condition over . The existence of a horseshoe satisfying follows directly from [33, Proposition 2.9]. Moreover, the horseshoe is constructed by shadowing the orbit of a generic point of , following the classical arguments of Katok [30, Theorem 4.3], hence is contained in the -neighborhood of and every invariant measure supported on is close to in the weak∗-topology. Finally, by the variational principle and the fact that , there exists satisfying . ∎
Remark 4.11.
The approximation by horseshoes in the conclusion of Lemma 4.10 can actually be extended to arbitrary invariant measures in . More precisely:
Proposition 4.12.
Let and be a singular hyperbolic homoclinic class of . If each pair of periodic orbits of are homoclinic related, then for every and , there exists a horseshoe and there exists satisfying
Thus has the horseshoe approximation property, and so it is entropy-dense.
Proof.
Let and be fixed. By ergodic decomposition and affinity of the metric entropy, there exist ergodic measures and real numbers with so that the probability satisfies
By Lemma 4.10, for each , there exists a horseshoe such that every satisfies , and there exists satisfying
Let be a large horseshoe that contains every for . Such does exist since any two periodic orbits in are homoclinically related and any horseshoe must contain (countably many) periodic orbits. Then the probability measure in satisfies and
By Proposition 4.6, we know that is entropy-dense. Thus, for the invariant measure and above, there exists satisfying
In consequence, the ergodic probability supported on satisfies
and
∎
Proposition 4.13.
Let and be a singular hyperbolic homoclinic class of . Assume each pair of periodic orbits of are homoclinic related, and . Then has the horseshoe approximation property.
Proof.
Fix an arbitrary . Using Proposition 4.12 it suffices to show that is approximated, both in the weak∗ topology and entropy, by measures in .
Fix an arbitrary constant . If there is nothing to prove. Otherwise, one can write for some and probabilities so that and . In other words, is supported in the invariant set formed by singularities and . By the assumption, there exists a sequence of probabilities with as . Notice that is a convex set. Thus for each . Moreover, using that is translation invariance and affinity of the entropy we get
and
By taking large, we conclude that the probability satisfies This completes the proof. ∎
Recently, S. Crovisier and D. Yang [20] proved that for -open and dense set of vector field , any singular hyperbolic attractor is a robustly transitive attractor. Moreover, if is non-trivial, then it is a homoclinic class and any two periodic orbits contained in are homoclinically related. On the other hand, the main theorems (Theorem A, B, B’) in [52] state that if is a singular hyperbolic attractor of in a residual subset , or is a geometric Lorenz attractor of in a residual subset , then , which implies naturally that . Thus one obtains the following consequence from Proposition 4.13.
Corollary 4.14.
The following holds:
-
1.
There exists a residual subset where such that if is a geometric Lorenz attractor of , then and thus has the horseshoe approximation property and is entropy-dense.
-
2.
There exists a residual subset such that if is a non-trivial singular hyperbolic attractor of , then and thus has the horseshoe approximation property and is entropy-dense.
The following proposition, whose strong conclusion will not be used in full strength in this paper guarantees that the entropy of a singular hyperbolic homoclinic class can be approximated by a horseshoe supporting ergodic measures which are dense enough. More precisely:
Proposition 4.15.
Let and be a singular hyperbolic homoclinic class of . If each pair of periodic orbits of are homoclinic related, then for every and , there exist a horseshoe and so that and .
Proof.
In the special case that we have that hyperbolic (recall Remark 2.3), hence itself is a basic set of . Moreover, if this is the case then by [54, Theorem 1]. Thus has the strong horseshoe approximation property by Proposition 4.13 and one concludes.
It remains to consider the case where . We first construct a nested sequence of horseshoes whose entropies approximate to . Notice first that every periodic orbit in is hyperbolic, hence the non-trivial homoclinic class contains countably many periodic orbits which we list as . Moreover, being a non-trivial homoclinic class ensures that . By the variational principle and Lemma 4.10, there is a sequence of horseshoes contained in such that . Notice that any two periodic orbits contained in have the same stable index and are homoclinically related with each other, and each horseshoe must contain infinitely many periodic orbits for each . Thus, inductively, we can construct a nested sequence of transitive horseshoes contained in as follows:
-
•
Let be a horseshoe that contains and . Such does exist since is homoclinically related with all periodic orbits contained in .
-
•
For , let be a horseshoe that contains and also contains the two horseshoes and . Such a horseshoe exists by Lemma 4.5.
Then we have that for every .
5 Multifractal analysis
In this section, we aim to study the multifractal analysis of singular hyperbolic attractors of -generic vector fields and geometric Lorenz attractors of -generic vector fields (). We prove the following theorem in general case. With the arguments in Section 4.2, one will see that Theorems A & B are consequences of Theorem 5.1 below (cf. Subsection 5.2). The strategy is to use the horseshoe approximation property to transfer the difficulty to the description of suitably chosen horseshoes. Once this is accomplished, then one can use Thompson’s results to get full topological entropy of irregular sets [58] and variational principle of level sets [57].
Theorem 5.1.
Let and be a singular hyperbolic homoclinic class of . Assume each pair of periodic orbits contained in are homoclinically related and . Given , then either
-
1.
is empty and for all ; or
-
2.
is residual in and .
Moreover, for each satisfying the level set is dense in and
Furthermore the set of functions satisfying the second item form an open and dense subset in .
Remark 5.2.
5.1 Entropy estimates for irregular sets and level sets
Given a compact metric space , for a homeomorphism and a continuous roof function , we consider the suspension flow where as defined in Section 2.4. Analogous to the discrete case, for a continuous function , we define
Define the irregular set
and for , define the level set
For a dynamical system satisfying the specification property, Thompson proved the following variational principle of level sets [57, Theorem 4.2] and full topological entropy of irregular sets [58, Theorem 5.1] for the suspension flow of .
Theorem 5.3 (Thompson [58], [57]).
Let be a compact metric space, be a homeomorphism satisfying the specification property and be a continuous function. Let denote the suspension flow over with roof function and let be a continuous function. Then:
-
1.
For any ,
-
2.
If then
Remark 5.4.
Note that a suspension flow over a transitive subshift of finite type (SFT) is topologically conjugate to a suspension flow over a topologically mixing SFT, and a topologically mixing SFT satisfies the specification property, thus the conclusions of Theorem 5.3 hold for suspension flows over a transitive SFT.
As a corollary, we have the following conclusion for horseshoes of vector fields, which complement previous results on the multifractal analysis of hyperbolic flows [9, 16, 46]. Recall that for an invariant compact set of and for a continuous function , the -irregular set is
and, for each , the -level set is
Corollary 5.5.
Let be a horseshoe of and be a continuous function. The following properties hold:
-
1.
For any ,
-
2.
If , then
Proof.
By Theorem 4.3, there exists a suspension flow over a transitive SFT such that is semi-conjugate to , where is a continuous function, is the flow generated by and is the suspension flow. That is to say, there exists a finite-to-one continuous map satisfying that for any . In particular preserves entropy. As , then . Moreover, it is easy to check that for each , and . Thus Corollary 5.5 is a consequence of Theorems 4.3 and Remark 5.4. ∎
Now we apply Theorem 5.3 and Corollary 5.5 to singular hyperbolic homoclinic classes of vector fields.
Proposition 5.6.
Let be a singular hyperbolic homoclinic class of such that each pair of periodic orbits are homoclinically related and be a continuous function. Assume that is a convex subset satisfying the horseshoe approximation property and
Then the following properties are satisfied.
-
1.
The topological entropy of the -irregular set satisfies
-
2.
For any , the topological entropy of the level set satisfies
Remark.
It will be clear from the proof that for establishing item (2) above one uses exclusively the horseshoe approximation property assumption.
Proof.
Denote by for simplicity. Fix . Since is convex, one has that For each , take so that
-
•
-
•
We give the following claim first.
Claim 5.7.
There exists a horseshoe such that
-
(a)
;
-
(b)
there exist such that
Proof of Claim 5.7.
The proof is by applying the fact that satisfies the horseshoe approximation property twice and by Lemma 4.5. We give a short explanation. First, by the definition of and , there exist such that and . Since satisfies the horseshoe approximation property, there exist two horseshoes such that
On the other hand, by the horseshoe approximation property again, there exist two horseshoes and such that Items and are satisfied for and . Then by Lemma 4.5, one can take a larger horseshoe that contains all the horseshoes , thus the above statements hold for . ∎
In the following, we take the horseshoe and measures from Claim 5.7.
Entropy of the -irregular set .
Entropy of the level set .
We claim there exists an invariant measure such that and tends to as . By item (b.1) above and the fact , one has that
Without loss of generality, we may assume that (the other case is analogous). By item (a) above, there exists such that
Thus one has
Then, by the affinity of the integral and the entropy function, one has that the probability measure satisfies and
By Item 1 in Corollary 5.5, one concludes that
As the right hand-side above tends to as tends to infinity, this proves item 2 of Proposition 5.6. ∎
5.2 Proofs of Theorems A & B
We first prove Theorem 5.1.
Proof of Theorem 5.1.
Assume and is a non-trivial singular hyperbolic homoclinic class of such that each pair of periodic orbits contained in are homoclinically related and . Proposition 4.13 implies has the horseshoe approximation property.
Let . If there exists such that for every , then and hence .
Entropy of .
By Proposition 4.12, one has that satisfies the horseshoe approximation property, thus by Item 1 of Proposition 5.6, one has
Recall that and if . Thus by the variational principle, one has
As a consequence,
Denseness and entropy of .
Take such that
As above, this ensures that there exist so that .
To obtain the denseness of in , one first constructs a nested sequence of horseshoes approximating . Note that all periodic orbits contained in are hyperbolic. Since is non-trivial, there are countably infinitely many periodic orbits contained in and one lists them as . Moreover, each pair of periodic orbits in are homoclinically related. One constructs inductively as follows:
-
•
Let be a horseshoe that contains and . Such a horseshoe exists because and are homoclinically related.
-
•
For , let be a horseshoe that contains and . Such a horseshoe exists because is homoclinically related with in the sense that is homoclinically with all periodic orbits contained in . Note that if is contained in , then .
By construction, one has that for all . Moreover, by denseness of the periodic orbits in one has that tends to (in the Hausdorff distance) as . Recall that one has assumed to be two periodic measures, thus there exists such that . As a consequence, for all . By Item 2 of Theorem 4.4, for each , there exists such that
Take a suitable for each , such that the linear combination
Note that . By Item 1 of Theorem 4.4, the set of -generic points is non-empty. Take , then where is the positive limit set of . This implies that . Note that , hence is dense in . As a consequence, one has is dense in . By the fact that , one has is dense in .
The estimation of the entropy follows similar arguments as for . Since satisfies the horseshoe approximation property, using Item 2 of Proposition 5.6, one has
The inverse inequality is obtained as an adaptation of Bowen’s result [14, Theorem 2] to the flow case, as we now explain. For a point , denote by and the limit sets of empirical measures of under the action of the flow and its time-one map , respectively. In other words,
and
To simplify notations, let . We need the following:
Claim.
For any and any , one has .
Proof of the claim.
Take . Each satisfies that and thus . On the other hand, for any , the measure is invariant by the flow and belongs to . As a consequence, one has since the metric entropy is affine on the space of invariant probability measures. ∎
Finally, it remains to show that the set
is open and dense in . The openness is by continuity of the integrals in the weak∗ topology. To prove denseness, take two different periodic orbits in , take such that
Such exists since and are two distinct periodic orbits. Let be the two periodic measures associated to respectively. Note that . For any ,
-
•
if , then ;
-
•
if otherwise , let , then and .
This shows is dense in and completes the proof of Theorem 5.1. ∎
Proofs of Theorem A & B.
By Item 1 of Corollary 4.14, there exists a residual subset where such that for any , if is a geometric Lorenz attractor for , then . Moreover, by Proposition 2.6, every pair of periodic orbits are homoclinically related.
6 Large deviations
Here we will focus on large deviations for singular hyperbolic attractors, including the geometric Lorenz attractor. The theory of large deviations for singular hyperbolic attractors is still not completely understood, appart from the level-1 large deviations upper bounds associated to its SRB measure in [4, 6, 21]. This section is devoted to the proof of Theorem C, which generalizes the large deviation results by L. S. Young [67] for flows with singularities. Our approach is inspired by [47, Section 3], which establishes criteria for level-2 large deviations principles for discrete-time dynamical systems. We overcome this fact dealing simultaneously with the flow dynamics (using the horseshoe approximation property and corresponding entropy denseness results) and induced discrete-time dynamics (taking suitable time- maps). Due to the presence of singularities, we still have to consider the following special subset
of . Recall that is the space of all probability measures supported on . We prove the following theorem in this section.
Theorem 6.1.
(Level-2 large deviations) Let and be a singular hyperbolic homoclinic class such that each pair of periodic orbits in are homoclinically related and . Assume is a weak Gibbs measure with respect to a Hölder continuous potential with being the -full measure set such that (2.1) satisfies. Then one has:
-
1.
(upper bound) There exists so that
for any closed subset .
-
2.
(lower bound) If is an open set and is ergodic satisfying , then
-
3.
(lower bound for Gibbs measure) If is a Gibbs measure with respect to , then
for any open subset .
For completeness of the paper, we give a short explanation of the proof of Theorem C.
Proof of Theorem C.
One takes the residual subset where and the residual subset in the proofs of Theorem A & B. In both cases when is a geometric Lorenz attractor for or is a singular hyperbolic attractor for , one has that and is a singular hyperbolic homoclinic class in which every pair of periodic orbits are homoclinically related. Thus Theorem C is a consequence of Theorem 6.1 ∎
In what follows, unless emphasized, we assume that is a singular hyperbolic homoclinic class of such that each pair of periodic orbits in are homoclinically related and we also assume that . Assume is a weak Gibbs measure with respect to a Hölder continuous potential and be the -full measure set satisfied for (2.1). To be precise, there exists such that for any and , there exist constants satisfying:
(6.1) |
for any dynamic Bowen ball .
6.1 Lower bound
We give the lower bounds of large deviations in this section, that is to prove item 2 & 3 of Theorem 6.1. The following instrumental result proves item 2.
Proposition 6.2.
Let and be as in the assumption above. If is an open set, is ergodic and then
Proof.
The argument is inspired by [47, Proposition 3.1], with the necessary adaptations to the context of weak Gibbs measures. Recall that one takes a dense subset of where is not the zero function for every and, for any ,
In consequence:
-
(i)
for every and ,
-
(ii)
for every .
Since is open in the weak∗-topology and , by the definition of the weak∗ topology one may choose and finitely many functions as above so that the open neighborhood
is contained in and
As is a weak Gibbs measure, recall that for any , and there exist constants satisfying (6.1). Let be small and fixed (to be chosen below). Firstly, by [48], the set of real numbers so that is ergodic for the time- map is Baire generic. One chooses such an small enough such that
and let denote the time- map. Then Proposition 2.1 in [47] applied to the open neighborhood of ensures that there is and for every there exists a set 666Equation (6.2) is a modified version of the statement of Proposition 2.1. Nevertheless, in the notation of [47], the argument carries out identically, if one replaces the sets at equation (2.25) in [47] by for some family of sets such that tends to 1 as . This is because of the fact .
(6.2) |
such that is -separated and has cardinality larger than or equal to . Using that it is clear that the set is -separated with respect to the flow . Taking , properties (i) and (ii) above imply
Analogously, if and then
Therefore one may reduce , if necessary, to guarantee that the summands in the right hand side above are arbitrarily small and consequently
Therefore, using the definition of weak Gibbs measure in (6.1), that and that the dynamic balls are pairwise disjoint, one concludes that
Taking , which makes as well, and by the choice of in (6.2), we conclude that
Since is small and arbitrary, this proves the proposition. ∎
As a consequence of Proposition 6.2, one can now prove the lower bound estimate for Gibbs measures in item 3 of Theorem 6.1. More precisely:
Corollary 6.3.
Let and be from the assumption of Proposition 6.2. Assume further that is a Gibbs measure with respect to . Given an open subset one has that
Proof.
Take an open subset , it is sufficient to show that, for each one has
(6.3) |
Note that since is a Gibbs measure, the set can be chosen as , thus one has for any .
If , then and there is nothing to prove. Hence one just needs to consider where . Since by assumption , Proposition 4.13 guarantees that admits the horseshoe approximation property. Thus, for any there exists so that is a horseshoe, and the followings are satisfied:
where is the supremum norm of . In particular one has . The following estimation holds
6.2 Upper bound
The large deviations upper bounds for the flow are inspired by [61, Theorem A] and [47, Theorem 3.2]. A first fundamental step is the following instrumental result, which explores ideas from Misiurewicz’s proof of the variational principle.
Lemma 6.4.
Let be an invariant compact set of a vector field and let be a non-empty set. If denotes the maximal cardinality of -separated sets in then,
where denotes the closed convex hull of .
If, in addition, the entropy function
is upper semicontinuous, then
where denotes the closure of .
Proof.
The proof is inspired by [47, Lemma 3.1], in the discrete time context. For completeness, we shall provide a sketch of the proof. Let be a non-empty set and let . For each let be a -separated set (with respect to the flow ) with cardinality . Choose a sequence so that
(6.4) |
and consider the probability measures
Up to consider some convergent subsequence, we may assume without loss of generality that is weak∗ convergent to . It is clear that . Moreover, as the sequence is a convex combination of probability measures in then . Therefore, using (6.4), in order to prove the first statement in the lemma it is enough to show that
(6.5) |
By Gronwall’s inequality, there exists so that
for every and . Since is a -separated set with respect to the flow , for any , there exists such that . Thus by the fact that , one has
where . Note that , so is a a -separated set with respect to the time-one map . Choosing a partition of with and , one concludes that each element of the partition contains at most one element of , and so
It is not hard to check that the probability measures converge to as well. Moreover, the argument used in the proof of the variational principle (see e.g. [47, Lemma 3.1] or [62, Theorem 8.6]) ensures that
This proves (6.5), and the first statement in the lemma.
Now, assume that is a non-empty set and that is upper semicontinuous. As is compact, for each there exists a finite open cover of by balls of radius . In particular there exists so that
Using the first statement of the lemma, there exists so that
In particular, any weak∗ accumulation point of belongs to and, by upper semicontinuity of the entropy, This completes the proof of the lemma. ∎
The previous result will allow to obtain the desired large deviations upper bounds. We observe that where denotes the time-1 of the flow . Moreover, notice that the entropy function associated to singular-hyperbolic attractors is upper-semicontinuous (cf. [42]). Thus, the large deviations upper bound in item 1 of Theorem 6.1 is now a direct consequence of the following proposition.
Proposition 6.5.
Let be an invariant compact set of a vector field and let be a closed and convex subset so that . Assume is a weak Gibbs measure with respect to a Hölder continuous potential and be the -full measure set satisfied for (2.1). Consider the non-positive real number
(6.6) |
Then
(6.7) |
Furthermore, if the entropy function is upper-semicontinuous then (6.7) holds even if is not convex.
Proof.
Assume first that is a closed and convex subset. As is Hölder continuous, then it is bounded and, given , one can write where
. Note that some of the sets , which are closed and convex, may be empty. For each non-empty , either there exists so that for every or . For that reason we will assume, without loss of generality, that all intersect the space of invariant probability measures. Now, for each so that , one has
(6.8) |
The maximal cardinality of a -separated set in the first set in the right hand-side above is bounded above by which, by Lemma 6.4, satisfies
Given small and fixed, pick a -maximal separated set . If then, using the weak Gibbs property, one ensures that
and, consequently,
This, combined with (6.8), ensures that is bounded above by
for each small . Taking the as in each of the terms in the previous expression we conclude that
thus proving the first statement in the proposition. Finally, if is upper semicontinuous then the large deviations upper bound holds for arbitrary closed sets as a consequence of the previous argument and the corresponding statement in Lemma 6.4 for such class of sets. This finishes the proof of the proposition. ∎
6.3 A local level-1 large deviations principle
Finally we note that Theorem 6.1 implies large deviations principle for singular hyperbolic sets and averages of continuous observables. Indeed, Theorem 6.1 together with Corollary 6.3 and the contraction principle (see [19]) implies on the following:
Corollary 6.6.
(Level-1 large deviations) Let and be a singular hyperbolic homoclinic class such that each pair of periodic orbits in are homoclinically related and . Assume is a Gibbs measure with respect to a Hölder continuous potential . For any continuous observable it holds that
and
where the lower-semicontinuous rate function is given by
Moreover, if there exist so that and then the infima in the right-hand side of the previous inequalities are strictly negative.
Acknowledgments
References
- [1] F. Abdenur, C. Bonatti and S. Crovisier, Nonuniform hyperbolicity for -generic diffeomorphisms. Israel J. Math. 183 (2011), 1–60.
- [2] L. M. Abramov, On the entropy of a flow. Doklady Akademii Nauk SSSR 128 (1959), 873–875.
- [3] V. S. Aframovi, V. V. Bykov and L. P. Sil’nikov, The origin and structure of the Lorenz attractor. (Russian) Dokl. Akad. Nauk SSSR 234 (1977), 336–339.
- [4] V. Araújo, Large deviations bound for semiflows over a non-uniformly expanding base. Bull. Braz. Math. Soc. (N.S.) 38 (2007), 335–376.
- [5] V. Araújo and M. J. Pacifico, Three-dimensional flows, A Series of Modern Surveys in Mathematics, Springer-Verlag, Berlin Heidelberg 2010.
- [6] V. Araújo, A. Souza and E. Trindade, Upper large deviations bound for singular-hyperbolic attracting sets. J. Dynam. Differential Equations 31 (2019), 601–652.
- [7] L. Barreira and G. Iommi, Suspension flows over countable Markov shifts. J. Stat. Phys. 124 (2006), 207–230.
- [8] L. Barreira, Y. Pesin and J. Schmeling, Multifractal spectra and multifractal rigidity for horseshoes. J. Dynam. Control Systems 3 (1997), 33–49.
- [9] L. Barreira and B. Saussol, Multifractal analysis of hyperbolic flows. Comm. Math. Phys. 214 (2000), 339–371.
- [10] L. Barreira and J. Schmeling, Sets of “non-typical” points have full topological entropy and full Hausdorff dimension. Israel J. Math. 116 (2000), 29–70.
- [11] T. Bomfim, M. J. Torres and P. Varandas, The gluing orbit property and partial hyperbolicity. J. Diff. Equations 272 (2021), 203–221.
- [12] T. Bomfim and P. Varandas, The gluing orbit property, uniform hyperbolicity and large deviations principles for semiflows. J. Diff. Equations 267 (2019), 228–266.
- [13] R. Bowen, Periodic orbits for hyperbolic flows. Amer. J. Math. 94 (1972), 1–30.
- [14] R. Bowen, Topological entropy for noncompact sets. Trans. Amer. Math. Soc. 184 (1973), 125–136.
- [15] R. Bowen, Symbolic dynamics for hyperbolic flows. Amer. J. Math. 95 (1973), 429–460.
- [16] M. Carvalho and P. Varandas, Genericity of historic behavior for maps and flows. Nonlinearity 34 (2021) 7030–7044.
- [17] V. Climenhaga, Multifractal formalism derived from thermodynamics for general dynamical systems. Electr. Research Announcements, 17 2010 1–11.
- [18] V. Climenhaga, D. Thompson and K. Yamamoto. Large deviations for systems with non-uniform structure. Trans. Amer. Math. Soc. 369 (2017), 4167–4192.
- [19] H. Comman and J. Rivera-Letelier, Large deviation principles for non- uniformly hyperbolic rational maps. Ergodic Theory Dynam. Systems 31 (2011), 321–349.
- [20] S. Crovisier and D. Yang, Robust transitivity of singular hyperbolic attractors. Math. Z. 298 (2021), 469–488.
- [21] S. Crovisier, D. Yang and J. Zhang, Empirical measures of partially hyperbolic attractors. Comm. Math. Phys. 375 (2020), 725–764.
- [22] A. Dembo and O. Zeitouni. Large deviations techniques and applications. Second Edition, Springer, New-York, 1998.
- [23] M. Denker and M. Kesseböhmer, Thermodynamical formalism, large deviation and multifractals. Stochastic Climate Models, Progress in Probability 49, 159–169, 2001.
- [24] A. Eizenberg, Y. Kifer and B. Weiss, Large Deviations for -actions. Comm. Math. Phys. 164 (1994), 433–454.
- [25] D. Feng and W. Huang, Lyapunov spectrum of asymptotically sub-additive potentials. Comm. Math. Phys. 297 (2010), 1–43.
- [26] J. Guckenheimer, A strange, strange attractor, in The Hopf Bifurcation Theorems and its Applications. Applied Mathematical Series 19, Springer-Verlag, 1976, 368–381.
- [27] J. Guckenheimer and R. F. Williams, Structural stability of Lorenz attractors. Publ. Math. Inst. Hautes Études Sci., 50 (1979), 59–72.
- [28] M. Hirsch, C. Pugh and M. Shub, Invariant Manifolds. Lecture Notes in Mathematics 583, Springer Verlag, Berlin (1977).
- [29] F. Hofbauer, Examples for the nonuniqueness of the equilibrium state. Trans. Amer. Math. Soc. 228 (1977), 223–241.
- [30] A. Katok, Lyapunov exponents, entropy and periodic points of diffeomorphisms. Publ. Math. Inst. Hautes Études Sci. 51 (1980), 137–173.
- [31] Y. Kifer, Large deviations in dynamical systems and stochastic processes. Trans. Amer. Math. Soc. 321 (1990), 505–524.
- [32] M. Li, S. Gan and L. Wen, Robustly transitive singular sets via approach of extended linear Poincaré flow. Discrete Contin. Dyn. Syst. 13 (2005), 239–269.
- [33] M. Li, Y. Shi, S. Wang and X. Wang, Measures of intermediate entropies for star vector fields. Israel J. Math. 240 (2020), 791–819.
- [34] E. N. Lorenz, Deterministic nonperiodic flow. J. Atmospheric Sci. 20 (1963), 130–141.
- [35] R. Metzger and C. A. Morales, On sectional-hyperbolic systems. Ergodic Theory Dynam. Systems 28 (2008), 1587–1597.
- [36] I. Melbourne and M. Nicol, Large deviations for nonuniformly hyperbolic systems. Trans. Amer. Math. Soc. 360 (2008), 6661–6676
- [37] C. A. Morales, M. J. Pacifico and E. Pujals, Singular hyperbolic systems. Proc. Amer. Math. Soc. 127 (1999), 3393–3401.
- [38] C. A. Morales, M. J. Pacifico and E. Pujals, Robust transitive singular sets for 3-flows are partially hyperbolic attractors or repellers. Ann. of Math. 160 (2004), 1–58.
- [39] L. Olsen, A multifractal formalism. Adv. Math. 116 (1995), 82–196.
- [40] L. Olsen and S. Winter, Normal and non-normal points of self-similar sets and divergence points of self-similar measures. J. Lond. Math. Soc. 67 (2003), 103–122.
- [41] M. J. Pacifico and M. Todd, Thermodynamic formalism for contracting Lorenz flows. J. Stat. Phys. 139 (2010), 159–176.
- [42] M. J. Pacifico, F. Yang and J. Yang, Entropy theory for sectional hyperbolic flows. Ann. Inst. H. Poincaré Anal. Non Linéaire 38 (2021), 1001–1030.
- [43] J. Palis and W. de Melo, Geometric theory of dynamical systems. An introduction. Springer-Verlag, New York-Berlin, 1982.
- [44] W. Parry and M. Pollicott, Zeta functions and the periodic orbit structure of hyperbolic dynamics. Asteŕisque 187–188, Soc. Math, France, 1990.
- [45] Y. Pesin, Dimension Theory in Dynamical Systems, Contemporary views and applications. Chicago Lectures in Mathematics. University of Chicago Press, Chicago, IL, 1997. xii+304 pp. ISBN: 0-226-66221-7; 0-226-66222-5.
- [46] Y. Pesin and V. Sadovskaya, Multifractal analysis of conformal Axiom A flows. Commun. Math. Phys. 216 (2001) 277–312
- [47] C. Pfister and W. Sullivan, Large deviations estimates for dynamical systems without the specification property. Applications to the -shifts. Nonlinearity 18 (2005), 237–261.
- [48] C. Pugh and M. Shub, Ergodic elements of ergodic actions. Compositio Math. 23 (1971), 115–122.
- [49] L. Rey-Bellet and L.-S Young, Large deviations in non-uniformly hyperbolic dynamical systems. Ergodic Theory Dynam. Systems 28 (2008), 578–612.
- [50] E. A. Sataev, Invariant measures for singularly hyperbolic attractors. (Russian. Russian summary) Mat. Sb. 201 (2010), 107–160; translation in Sb. Math. 201 (2010), 419–470.
- [51] Y. Shi, S. Gan and L. Wen, On the singular hyperbolicity of star flows. J. Mod. Dyn. 8 (2014), 191–219.
- [52] Y. Shi, X. Tian and X. Wang, The space of ergodic measures for Lorenz attractors. ArXiv:2006.08193.
- [53] K. Sigmund, Generic properties of invariant measures for Axiom A diffeomorphisms. Invent. Math. 11 (1970), 99–109.
- [54] K. Sigmund, On the Space of Invariant Measures for Hyperbolic Flows. American Journal of Mathematics 94 (1972), 31–37.
- [55] N. Sumi, P. Varandas and K. Yamamoto, Specification and partial hyperbolicity for flows. Dynam. Sys. 30 (2015), 501–524.
- [56] F. Takens and E. Verbitski, Multifractal analysis of local entropies for expansive homeomorphisms with specification. Comm. Math. Phys. 203 (1999), 593–612.
- [57] D. J. Thompson, A variational principle for topological pressure for certain non-compact sets. J. London Math. Soc. 80 (2009), 585–602.
- [58] D. J. Thompson, The irregular set for maps with the specification property has full topological pressure. Dynam. Syst. 25 (2010), 25–51.
- [59] X. Tian and W. Sun, Diffeomorphisms with various C1 stable properties. Acta Math. Sci. Ser. B (Engl. Ed.) 32 (2012), 552–558.
- [60] X. Tian and P. Varandas, Topological entropy of level sets of empirical measures for non-uniformly expanding maps. Discrete Cont. Dynam. Sys. 37 10 (2017), 5407–5431.
- [61] P. Varandas, Non-uniform specification and large deviations for weak Gibbs measures. Journal of Stat. Phys., 146 (2012), 330–358.
- [62] P. Walters, An Introduction to Ergodic Theory. Graduate Texts in Mathematics, 79, Springer, Berlin (2000).
- [63] T. Wang, Unique equilibrium states, large deviations and Lyapunov spectra for the Katok map. Ergodic Theory Dynam. Systems 41 (2021), 2182–2219.
- [64] L. Wen, Differentiable Dynamical Systems. Graduate Studies in Mathematics, 173, American Mathematical Society, Providence, RI, (2016).
- [65] X. Wen and L. Wen, No-shadowing for singular hyperbolic sets with a singularity. Discrete Contin. Dyn. Syst. 40 (2020), 6043–6059.
- [66] R. F. Williams, The structure of Lorenz attractors. Publ. Math. Inst. Hautes Études Sci., 50 (1979), 73–99.
- [67] L. Young, Some large deviation results for dynamical systems. Trans. Amer.Math. Soc. 318 (1990), 525–543.