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

On multifractal analysis and large deviations
of singular hyperbolic attractors

Yi Shi111Y. Shi was partially supported by National Key R&D Program of China (2021YFA1001900) and NSFC (12071007, 11831001, 12090015).  ,  Xueting Tian222X. Tian was partially supported by NSFC (12071082) and Science and Technology Innovation Action Program of Science and Technology Commission of Shanghai Municipality (STCSM, No. 21JC1400700).   Paulo Varandas333P. Varandas was partially supported by the grant CEECIND/03721/2017 of the Stimulus of Scientific Employment, Individual Support 2017 Call, awarded by FCT-Portugal, and by CMUP, which is financed by national funds through FCT-Portugal, under the project UIDB/00144/2020.   and Xiaodong Wang444X. Wang was partially supported by National Key R&D Program of China (2021YFA1001900), NSFC (12071285, 11701366), Science and Technology Innovation Action Program of Science and Technology Commission of Shanghai Municipality (STCSM, No. 20JC1413200) and Innovation Program of Shanghai Municipal Education Commission (No. 2021-01-07-00-02-E00087).
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 C1C^{1}-generic singular hyperbolic attractors and CrC^{r}-generic geometric Lorenz attractors (r2)(r\geq 2). 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 𝒳r(M)\mathscr{X}^{r}(M), r1r\geq 1, denote the space of CrC^{r}-vector fields on a closed smooth Riemannian manifold MM endowed with the CrC^{r}-topology. Given a vector field X𝒳1(M)X\in\mathcal{X}^{1}(M) and a compact invariant subset Λ\Lambda of the C1C^{1}-flow (ϕt)t(\phi_{t})_{t\in\mathbb{R}} generated by XX, we denote by C(Λ,)C(\Lambda,\mathbb{R}) the space of continuous functions on Λ\Lambda. For any gC(Λ,)g\in C(\Lambda,\mathbb{R}), Birkhoff’s ergodic theorem ensures that for any μinv(Λ)\mu\in\mathcal{M}_{inv}(\Lambda), the time average limT1T0Tg(ϕt(x))dt\displaystyle\lim_{T\rightarrow\infty}\frac{1}{T}\int_{0}^{T}g(\phi_{t}(x)){\rm d}t exists for μ\mu-almost all point xΛx\in\Lambda. Defining, for each aa\in\mathbb{R}, the gg-level set

Rg(a):={xΛ:limT1T0Tg(ϕt(x))dt=a},R_{g}(a)\colon=\left\{x\in\Lambda\colon\lim_{T\rightarrow\infty}\frac{1}{T}\int_{0}^{T}g(\phi_{t}(x)){\rm d}t=a\right\},

and the gg-irregular set by

Ig:={xΛ:limT1T0Tg(ϕt(x))dt does not exist},I_{g}\colon=\left\{x\in\Lambda\colon\lim_{T\rightarrow\infty}\frac{1}{T}\int_{0}^{T}g(\phi_{t}(x))\,{\rm d}t\text{ does not exist}\right\},

one obtains the multifractal decomposition

Λ=IgaRg(a)\Lambda=I_{g}\;\cup\;\bigcup_{a\in\mathbb{R}}R_{g}(a)

of the flow with respect to the observable gg. 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 ν\nu on Λ\Lambda (possibly non-invariant) one would like to provide sharp estimates for the ν\nu-measure the deviation sets

{xΛ:1T0Tg(ϕt(x))dt>c}\left\{x\in\Lambda:\frac{1}{T}\int_{0}^{T}g(\phi_{t}(x)){\rm d}t>c\right\}

for all gC(Λ,)g\in C(\Lambda,\mathbb{R}) and all real numbers cc. 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. Afraı˘\breve{\rm\i}movicˇ\check{\rm c}-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 CrC^{r} vector fields (r2r\in\mathbb{N}_{\geq 2}) exhibiting a geometric Lorenz attractor is an open subset in 𝒳r(M3)\mathscr{X}^{r}(M^{3}) (cf. [52]). In the study of C1C^{1} 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 r2r\geq 2, there exist a Baire residual subset r\mathcal{R}^{r} and a dense subset 𝒟r\mathcal{D}^{r} in 𝒳r(M)\mathscr{X}^{r}(M) such that for a geometric Lorenz attractor Λ\Lambda of XrX\in\mathcal{R}^{r}, the space of ergodic probability measures supported on Λ\Lambda is connected, while for a geometric Lorenz attractor Λ\Lambda^{\prime} of X𝒟rX\in\mathcal{D}^{r}, the space of ergodic probability measures supported on Λ\Lambda^{\prime} is not connected; and a similar statement holds for C1C^{1}-singular hyperbolic attractors in higher dimension [52, Theorem A&B]. In rough terms, the underlying idea is that while the set

1(Λ)={μinv(Λ):μ(Sing(Λ))=0}\mathcal{M}_{1}(\Lambda)=\Big{\{}\mu\in\mathcal{M}_{inv}(\Lambda)\colon\mu(\operatorname{Sing}(\Lambda))=0\Big{\}}

of probability measures which give zero weight to the singularity set Sing(Λ)\operatorname{Sing}(\Lambda) of a singular hyperbolic attractor Λ\Lambda 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 Rg(a)R_{g}(a) involves the uniqueness of equilibrium states for Hölder continuous observables gg. However this is still an open problem for singular hyperbolic attractors. On the other hand, if the observable gg 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 IgI_{g}. 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 CrC^{r}-generic geometric Lorenz attractors and C1C^{1}-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 Λ\Lambda. 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 𝒳r(M)\mathscr{X}^{r}(M), r1r\geq 1, denote the space of CrC^{r}-vector fields on a closed smooth Riemannian manifold MM. For X𝒳r(M)X\in\mathscr{X}^{r}(M), denote by ϕtX\phi_{t}^{X} or ϕt\phi_{t} for simplicity the CrC^{r}-flow generated by XX and denote by Dϕt{\rm D}\phi_{t} the tangent map of ϕt\phi_{t}. Moreover, given any ϕt\phi_{t}-invariant set Λ\Lambda we denote by Sing(Λ)\operatorname{Sing}(\Lambda) the set of singularities for the vector field XX in Λ\Lambda. Let (Λ)\mathcal{M}(\Lambda) be the space of all probability measures supported on Λ\Lambda endowed with the weak*-topology. Let dd^{*} be a metric on the space (Λ)\mathcal{M}(\Lambda) 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 {φi}i=1\{\varphi_{i}\}_{i=1}^{\infty} of C(Λ,)C(\Lambda,\mathbb{R}) where φi\varphi_{i} is not the zero function for every i1i\geq 1, and for any μ,ν(Λ)\mu,\nu\in\mathcal{M}(\Lambda) set

d(μ,ν)=i=1|φidμφidν|2iφi,\displaystyle d^{*}(\mu,\nu)=\sum_{i=1}^{\infty}\frac{|\int\varphi_{i}{\rm d}\mu-\int\varphi_{i}{\rm d}\nu|}{2^{i}\|\varphi_{i}\|},

where \|\cdot\| denotes the CC^{\infty}-norm on C(Λ,)C(\Lambda,\mathbb{R}). The set of invariant (resp. ergodic) probability measures of XX supported on Λ\Lambda is denoted by inv(Λ)\mathcal{M}_{inv}(\Lambda) (resp. erg(Λ)\mathcal{M}_{erg}(\Lambda)). We denote by hμ(X)h_{\mu}(X) the metric entropy of the invariant probability measure μinv(Λ)\mu\in\mathcal{M}_{inv}(\Lambda), defined as the metric entropy of μ\mu with respect to the time-1 map ϕ1X\phi^{X}_{1} 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 X𝒳1(M)X\in\mathscr{X}^{1}(M), a compact ϕt\phi_{t}-invariant set Λ\Lambda is hyperbolic if Λ\Lambda admits a continuous Dϕt{\rm D}\phi_{t}-invariant splitting TΛM=EsXEuT_{\Lambda}M=E^{s}\oplus\langle X\rangle\oplus E^{u}, where X\langle X\rangle denotes the one-dimensional linear space generated by the vector field, and EsE^{s} (resp. EuE^{u}) is uniformly contracted (resp. expanded) by Dϕt{\rm D}\phi_{t}, that is to say, there exist constants C>0C>0 and η>0\eta>0 such that for any xΛx\in\Lambda and any t0t\geq 0,

  • Dϕt(v)Ceηtv\|{\rm D}\phi_{t}(v)\|\leq Ce^{-\eta t}\|v\|, for any vEs(x)v\in E^{s}(x); and

  • Dϕt(v)Ceηtv\|{\rm D}\phi_{-t}(v)\|\leq Ce^{-\eta t}\|v\|, for any vEu(x)v\in E^{u}(x).

for any xΛx\in\Lambda and t0t\geq 0. If Λ\Lambda is transitive, ie it admits a dense orbit, then the dimension dim(Es)\dim(E^{s}) 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 X𝒳1(M)X\in\mathscr{X}^{1}(M), a compact and invariant set Λ\Lambda is singular hyperbolic if it admits a continuous Dϕt{\rm D}\phi_{t}-invariant splitting TΛM=EssEcuT_{\Lambda}M=E^{ss}\oplus E^{cu} and constants C,η>0C,\eta>0 such that, for any xΛx\in\Lambda and any t0t\geq 0,

  • EssEcuE^{ss}\oplus E^{cu} is a dominated splitting : Dϕt|Ess(x)Dϕt|Ecu(ϕt(x))<Ceηt\|{\rm D}\phi_{t}|_{E^{ss}(x)}\|\cdot\|{\rm D}\phi_{-t}|_{E^{cu}(\phi_{t}(x))}\|<Ce^{-\eta t}, and

  • EssE^{ss} is uniformly contracted by Dϕt{\rm D}\phi_{t} : Dϕt(v)<Ceηtv\|{\rm D}\phi_{t}(v)\|<Ce^{-\eta t}\|v\| for any vEss(x){0}v\in E^{ss}(x)\setminus\{0\};

  • EcuE^{cu} is sectionally expanded by Dϕt{\rm D}\phi_{t} : |detDϕt|Vx|>Ceηt|\det{\rm D}\phi_{t}|_{V_{x}}|>Ce^{\eta t} for any 2-dimensional subspace VxExcuV_{x}\subset E^{cu}_{x}.

Remark 2.3.

The following properties hold:

  1. 1.

    Given X𝒳1(M)X\in\mathscr{X}^{1}(M), it follows from the definition of singular-hyperbolicity that all hyperbolic periodic orbits of a singular-hyperbolic set Λ\Lambda have the same index.

  2. 2.

    Given X𝒳1(M)X\in\mathscr{X}^{1}(M) and an invariant compact set Λ\Lambda which contains regular points (i.e. points that are not singularities), if Λ\Lambda is hyperbolic, then it must contain no singularities. On the other hand, if Sing(Λ)=\operatorname{Sing}(\Lambda)=\emptyset, then Λ\Lambda is hyperbolic if and only if Λ\Lambda is singular hyperbolic for XX or for X-X (cf.  [38]).

  3. 3.

    Singular hyperbolicity is a C1C^{1}-robust property. More precisely, if Λ\Lambda is a singular hyperbolic invariant compact set of X𝒳1(M)X\in\mathscr{X}^{1}(M) associated with splitting TΛM=EssEcuT_{\Lambda}M=E^{ss}\oplus E^{cu} and constants (C,η)(C,\eta), then there exists an open neighborhood UU of Λ\Lambda and a neighborhood 𝒰𝒳1(M)\mathcal{U}\subset\mathscr{X}^{1}(M) of XX such that for any Y𝒰Y\in\mathcal{U}, the maximal invariant set of ϕtY\phi_{t}^{Y} in UU is a singular hyperbolic set for YY with the same stable dimension and constants (C,η)(C,\eta) (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 pp of XX, the stable/unstable manifolds associated to its orbit Orb(p)\operatorname{Orb}(p) is defined as follows:

Ws(Orb(p))={x:limt+d(ϕt(x),Orb(p))=0},W^{s}(\operatorname{Orb}(p))=\left\{x\colon\lim_{t\rightarrow+\infty}d\big{(}\phi_{t}(x),\operatorname{Orb}(p)\big{)}=0\right\},
Wu(Orb(p))={x:limt+d(ϕt(x),Orb(p))=0}.W^{u}(\operatorname{Orb}(p))=\left\{x\colon\lim_{t\rightarrow+\infty}d\big{(}\phi_{-t}(x),\operatorname{Orb}(p)\big{)}=0\right\}.

The classical theory on invariant manifolds by [28] ensures that Ws(Orb(p))W^{s}(\operatorname{Orb}(p)) and Wu(Orb(p))W^{u}(\operatorname{Orb}(p)) are submanifolds of MM.

Definition 2.4.

Let X𝒳1(M)X\in\mathscr{X}^{1}(M) and assume p,qp,q are two hyperbolic periodic points of XX. The homoclinic class of pp is defined as

H(p)=Ws(Orb(p))Wu(Orb(p))¯,H(p)=\overline{W^{s}(\operatorname{Orb}(p))\pitchfork W^{u}(\operatorname{Orb}(p))},

that is, the closure of the set of transversal intersections between the stable and unstable manifolds of the periodic orbit Orb(p)\operatorname{Orb}(p) of pp. The two hyperbolic periodic orbits Orb(p)\operatorname{Orb}(p) and Orb(q)\operatorname{Orb}(q) are homoclinically related if the stable manifold of Orb(p)\operatorname{Orb}(p) has non-empty transverse intersections with the unstable manifold of Orb(q)\operatorname{Orb}(q) 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 M3M^{3}. Recall that for a hyperbolic singularity σ\sigma of a vector field X𝒳1(M)X\in\mathscr{X}^{1}(M), its local stable manifold of size δ\delta is

Wδs(σ)={x:limt+ϕt(x)=σ and d(ϕt(x),σ)<δ,t0}.W_{\delta}^{s}(\sigma)=\left\{x\colon\lim_{t\rightarrow+\infty}\phi_{t}(x)=\sigma\text{~{}and~{}}d(\phi_{t}(x),\sigma)<\delta,\forall t\geq 0\right\}.

When we do not specify the size δ\delta of the local stable manifold, we just write as Wlocs(σ)W^{s}_{loc}(\sigma).

Definition 2.5.

We say X𝒳r(M3)X\in\mathscr{X}^{r}(M^{3}) (r1r\geq 1) exhibits a geometric Lorenz attractor Λ\Lambda, if XX has an attracting region UM3U\subset M^{3} such that Λ=t>0ϕtX(U)\Lambda=\bigcap_{t>0}\phi^{X}_{t}(U) is a singular hyperbolic attractor and satisfies:

  • There exists a unique singularity σΛ\sigma\in\Lambda with three exponents λ1<λ2<0<λ3\lambda_{1}<\lambda_{2}<0<\lambda_{3}, which satisfy λ1+λ3<0\lambda_{1}+\lambda_{3}<0 and λ2+λ3>0\lambda_{2}+\lambda_{3}>0.

  • Λ\Lambda admits a CrC^{r}-smooth cross section Σ\Sigma which is C1C^{1}-diffeomorphic to [1,1]×[1,1][-1,1]\times[-1,1], such that l={0}×[1,1]=W𝑙𝑜𝑐s(σ)Σl=\{0\}\times[-1,1]=W^{s}_{\it loc}(\sigma)\cap\Sigma, and for every zUW𝑙𝑜𝑐s(σ)z\in U\setminus W^{s}_{\it loc}(\sigma), there exists t>0t>0 such that ϕtX(z)Σ\phi_{t}^{X}(z)\in\Sigma.

  • Up to the previous identification, the Poincaré map P:ΣlΣP:\Sigma\setminus l\rightarrow\Sigma is a skew-product map

    P(x,y)=(f(x),H(x,y)),(x,y)[1,1]2l.P(x,y)=\big{(}f(x)~{},~{}H(x,y)\big{)},\qquad\forall(x,y)\in[-1,1]^{2}\setminus l.

    Moreover, it satisfies

    • H(x,y)<0H(x,y)<0 for x>0x>0, and H(x,y)>0H(x,y)>0 for x<0x<0;

    • sup(x,y)Σl|H(x,y)/y|<1\sup_{(x,y)\in\Sigma\setminus l}\big{|}\partial H(x,y)/\partial y\big{|}<1, and sup(x,y)Σl|H(x,y)/x|<1\sup_{(x,y)\in\Sigma\setminus l}\big{|}\partial H(x,y)/\partial x\big{|}<1;

    • the one-dimensional quotient map f:[1,1]{0}[1,1]f:[-1,1]\setminus\{0\}\rightarrow[-1,1] is C1C^{1}-smooth and satisfies limx0f(x)=1\lim_{x\rightarrow 0^{-}}f(x)=1, limx0+f(x)=1\lim_{x\rightarrow 0^{+}}f(x)=-1, 1<f(x)<1-1<f(x)<1 and f(x)>2f^{\prime}(x)>\sqrt{2} for every x[1,1]{0}x\in[-1,1]\setminus\{0\}.

Refer to caption
Figure 1: Geometric Lorenz attractor and return map

It has been proved that geometric Lorenz attractor is a homoclinic class [5, Theorem 6.8]. Moreover, for r2{}r\in\mathbb{N}_{\geq 2}\cup\{\infty\}, vector fields exhibiting a geometric Lorenz attractor forms a CrC^{r}-open set in 𝒳r(M)\mathscr{X}^{r}(M) [52, Proposition 4.7] .

Proposition 2.6.

Let r2{}r\in\mathbb{N}_{\geq 2}\cup\{\infty\} and X𝒳r(M3)X\in\mathscr{X}^{r}(M^{3}). If XX exhibits a geometric Lorenz attractor Λ\Lambda with attracting region UU, then Λ\Lambda is a singular hyperbolic homoclinic class, and every pair of periodic orbits are homoclinic related. Moreover, there exists a CrC^{r}-neighborhood 𝒰\mathcal{U} of XX in 𝒳r(M3)\mathscr{X}^{r}(M^{3}), such that for every Y𝒰Y\in\mathcal{U}, UU is an attracting region of YY, and the maximal invariant set ΛY=t>0ϕtY(U)\Lambda_{Y}=\bigcap_{t>0}\phi_{t}^{Y}(U) is a geometric Lorenz attractor.

2.2 Topological entropy on general subsets

Let (K,d)(K,d) be a compact metric space. The topological entropy of a continuous flow Φ=(ϕt)t\Phi=(\phi_{t})_{t} on a compact invariant set ZKZ\subset K can be defined as the topological entropy of its time-1 map ϕ1\phi_{1} on ZZ. When ZZ 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 ZKZ\subset K. For any xKx\in K, for any ε>0\varepsilon>0 and any t0t\geq 0, consider the (t,ε)(t,\varepsilon)-Bowen ball

B(x,t,ε,Φ):={yK:d(ϕτ(x),ϕτ(y))<ε,τ[0,t)}.B(x,t,\varepsilon,\Phi)\colon=\big{\{}y\in K:d(\phi_{\tau}(x),\phi_{\tau}(y))<\varepsilon,\tau\in[0,t)\big{\}}.

Now, given a finite or countable collections γ={B(xi,ti,ε,Φ)}i\gamma=\{B(x_{i},t_{i},\varepsilon,\Phi)\}_{i} of Bowen balls which covers ZZ and a fixed ss\in\mathbb{R}, we define:

Q(Z,s,γ,ε)\displaystyle Q(Z,s,\gamma,\varepsilon) =\displaystyle= B(xi,ti,ε,Φ)γexp(sti).\displaystyle\sum_{B(x_{i},t_{i},\varepsilon,\Phi)\in\gamma}\exp\left(-st_{i}\right).

For NN\in{\mathbb{N}}, define

M(Z,s,ε,N)\displaystyle M(Z,s,\varepsilon,N) =\displaystyle= infγQ(Z,s,γ,ε),\displaystyle\inf_{\gamma}\;Q(Z,s,\gamma,\varepsilon),

where the infimum is taken over all finite or countable collections γ={B(xi,ti,ε,Φ)}i\gamma=\{B(x_{i},t_{i},\varepsilon,\Phi)\}_{i} of Bowen balls which covers ZZ such that xiZx_{i}\in Z and tiNt_{i}\geq N, for all i1i\geq 1. Since M(Z,s,ε,N)M(Z,s,\varepsilon,N) is non-decreasing with respect to NN, the limit

M(Z,s,ε):=limNM(Z,s,ε,N)M(Z,s,\varepsilon)\colon=\lim_{N\rightarrow\infty}M(Z,s,\varepsilon,N)

does exist. It is not hard to check that

htop(Φ,Z,ε):=inf{s:M(Z,s,ε)=0}=sup{s:M(Z,s,ε)=}h_{\text{top}}(\Phi,Z,\varepsilon)\colon=\inf\big{\{}s\colon M(Z,s,\varepsilon)=0\big{\}}=\sup\big{\{}s\colon M(Z,s,\varepsilon)=\infty\big{\}}

is well defined. Then the topological entropy of ZZ with respect to the flow Φ=(ϕt)\Phi=(\phi_{t}) is

htop(Φ,Z):=limε0htop(Φ,Z,ε).h_{\text{top}}(\Phi,Z)\colon=\lim_{\varepsilon\to 0}h_{\text{top}}(\Phi,Z,\varepsilon).

Note that similarly with the definition of the metric entropy with respect to a flow, the topological entropy of ZZ with respect to the flow Φ=(ϕt)t\Phi=(\phi_{t})_{t} equals to that with respect to its time-one map ϕ1\phi_{1}.

2.3 Topological pressure and weak Gibbs measure

In this subsection, we always assume X𝒳1(M)X\in\mathcal{X}^{1}(M) and Λ\Lambda is an invariant compact set. Recall that we denote by Φ=(ϕt)t\Phi=(\phi_{t})_{t} the flow generated by XX. One first recalls the notion of topological pressure associated to a potential and equilibrium state.

Definition 2.7.

Given a continuous function ψ:Λ\psi\colon\Lambda\rightarrow\mathbb{R}, here called a potential, the topological pressure Ptop(Λ,ψ)P_{\rm top}(\Lambda,\psi) is defined as

Ptop(Λ,ψ)=supμinv(Λ){hμ(X)+ψdμ}.P_{\rm top}(\Lambda,\psi)=\sup_{\mu\in\mathcal{M}_{inv}(\Lambda)}\left\{h_{\mu}(X)+\int\psi{\rm d}\mu\right\}.

An invariant probability measure attaining the previous supremum is called an equilibrium state of (ϕtX)t(\phi^{X}_{t})_{t} with respect to ψ\psi.

Now we define the notion of weak Gibbs measure associated to which we will study the large deviations bounds.

Definition 2.8.

Let X𝒳1(M)X\in\mathscr{X}^{1}(M) and Λ\Lambda be an invariant compact set. Given a Hölder continuous potential ψ:Λ\psi\colon\Lambda\to\mathbb{R}, an invariant measure μψinv(Λ)\mu_{\psi}\in\mathcal{M}_{inv}(\Lambda) is called a weak Gibbs measure with respect to ψ\psi if there exists a full μψ\mu_{\psi}-measure subset ΛHΛ\Lambda_{H}\subset\Lambda and ε0>0\varepsilon_{0}>0 so that the following holds: for any xΛHx\in\Lambda_{H}, t>0t>0 and ε(0,ε0)\varepsilon\in(0,\varepsilon_{0}), there exists a constant Ct(x,ε)>1C_{t}(x,\varepsilon)>1 such that limt1tlogCt(x,ε)=0\lim\limits_{t\to\infty}\frac{1}{t}\log C_{t}(x,\varepsilon)=0 and

1Ct(x,ε)μψ(B(y,t,ε,Φ))etPtop(X,ψ)+0tψ(ϕs(x))dsCt(x,ε).\frac{1}{C_{t}(x,\varepsilon)}\leq\frac{\mu_{\psi}\Big{(}B\big{(}y,t,\varepsilon,\Phi\big{)}\Big{)}}{e^{-t\,P_{\rm top}(X,\psi)+\int_{0}^{t}\psi(\phi_{s}(x))\,{\rm d}s}}\leq C_{t}(x,\varepsilon). (2.1)

for any dynamic Bowen ball B(y,t,ε,Φ)B(x,t,ε0,Φ)B\big{(}y,t,\varepsilon,\Phi\big{)}\subset B\big{(}x,t,\varepsilon_{0},\Phi\big{)}.

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 ΛH\Lambda_{H} 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 Ct:Λ[1,+]C_{t}:\Lambda\to[1,+\infty] defined by Ct(x,ε)=+C_{t}(x,\varepsilon)=+\infty for every xΛΛHx\in\Lambda\setminus\Lambda_{H}. It is a standing assumption in the sense that, for each t>0t>0, the map Ct:Λ[1,+)C_{t}:\Lambda\to[1,+\infty) is lower semi-continuous, i.e. for each a1a\geq 1 the set

{xΛ:Ct(x,ε)[1,a]}is closed.\Big{\{}x\in\Lambda\colon C_{t}(x,\varepsilon)\in[1,a]\Big{\}}\quad\text{is closed}.

A special case of weak Gibbs property occurs when Ct(x,ε)C_{t}(x,\varepsilon) can be chosen as a constant Ct(ε)C_{t}(\varepsilon) independent of xx. In case the constants Ct(x,ε)C_{t}(x,\varepsilon) can be chosen independently of both tt and xx and ΛH=Λ\Lambda_{H}=\Lambda the probability measure μ\mu 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 f:KKf\colon K\to K be a homeomorphism on a compact metric space (K,d)(K,d) and consider a continuous roof function ρ:K(0,+)\rho\colon K\to(0,+\infty). We define the suspension space to be

Kρ={(x,s)K×[0,+):0sρ(x)}/,K_{\rho}=\{(x,s)\in K\times[0,+\infty)\colon 0\leq s\leq\rho(x)\}/\sim,

where the equivalence relation \sim identifies (x,ρ(x))(x,\rho(x)) with (f(x),0)(f(x),0), for all xKx\in K. Let π\pi denote the quotient map from K×[0,+)K\times[0,+\infty) to KρK_{\rho}. We define the flow 𝔉={ft}\mathfrak{F}=\{f_{t}\} on the quotient space KρK_{\rho} by

ft(x,s)=π(x,s+t).f_{t}(x,s)=\pi(x,s+t).

For any function g:Kρg\colon K_{\rho}\to\mathbb{R}, we associate the function φg:K\varphi_{g}\colon K\to\mathbb{R} by φg(x)=0ρ(x)g(x,t)dt\displaystyle\varphi_{g}(x)=\int_{0}^{\rho(x)}g(x,t){\rm d}t. Since the roof function ρ\rho is continuous, φg\varphi_{g} is continuous as long as gg is. Moreover, to each invariant probability μ(f,X)\mu\in\mathcal{M}(f,X) we associate the measure μρ\mu_{\rho} given by

Kρgdμρ=KφgdμKρdμgC(Kρ,).\displaystyle\int_{K_{\rho}}g{\rm d}\mu_{\rho}=\frac{\int_{K}\varphi_{g}{\rm d}\mu}{\int_{K}\rho{\rm d}\mu}\qquad\forall g\in C(K_{\rho},\mathbb{R}).

Observe that since ρ\rho is bounded away from zero, the measure μρ\mu_{\rho} is well-defined and 𝔉\mathfrak{F}-invariant, i.e. μρ(ft1A)=μρ(A)\mu_{\rho}(f_{t}^{-1}A)=\mu_{\rho}(A) for all t0t\geq 0 and measurable sets AA. Moreover, the map

:(f,K)(𝔉,Kρ)given byμμρ\mathcal{R}\colon\mathcal{M}(f,K)\to\mathcal{M}(\mathfrak{F},K_{\rho})\quad\text{given by}\quad\mu\mapsto\mu_{\rho}

is a bijection. Abramov’s theorem [2, 44] states that hμρ(𝔉)=hμ(f)/ρdμ\displaystyle h_{\mu_{\rho}}(\mathfrak{F})=h_{\mu}(f)/\int\rho{\rm d}\mu and hence, the topological entropy htop(𝔉)h_{\text{top}}(\mathfrak{F}) of the flow satisfies

htop(𝔉)=sup{hμρ(𝔉):μρM(𝔉,Kρ)}=sup{hμ(f)ρdμ:μM(f,K)}.h_{\text{top}}(\mathfrak{F})=\sup\Big{\{}h_{\mu_{\rho}}(\mathfrak{F})\colon\mu_{\rho}\in M(\mathfrak{F},K_{\rho})\Big{\}}=\sup\left\{\frac{h_{\mu}(f)}{\int\rho{\rm d}\mu}\colon\mu\in M(f,K)\right\}.

Throughout we will use the notation Φ=(ϕt)t\Phi=(\phi_{t})_{t} for a flow on a compact metric space and 𝔉=(ft)t\mathfrak{F}=(f_{t})_{t} 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 KK be a compact metric space and f:KKf\colon K\rightarrow K be a continuous map. We say ff satisfies the specification property if for any ε>0\varepsilon>0, there exists an integer m(ε)m(\varepsilon) such that the following holds: for any finite collection {[ai,bi]}i=1k\{[a_{i},b_{i}]\}_{i=1}^{k} of intervals with ai,bia_{i},b_{i}\in{\mathbb{N}} and ai+1bim(ε),i=1,2,,k1a_{i+1}-b_{i}\geq m(\varepsilon),i=1,2,\cdots,k-1 and for any points x1,x2,,xkKx_{1},x_{2},\cdots,x_{k}\in K, there exists a point xKx\in K such that

d(fai+t(x),ft(xi))<ε,for all t=1,2,,biai and all i=1,2,k.d(f^{a_{i}+t}(x),f^{t}(x_{i}))<\varepsilon,~{}~{}\text{for all $t=1,2,\cdots,b_{i}-a_{i}$ and all $i=1,2,\cdots k$}.

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 KK be a compact metric space and f:KKf\colon K\rightarrow K be a continuous map. We say ff satisfies the gluing orbit property if for any ε>0\varepsilon>0, there exists an integer M(ε)M(\varepsilon) such that the following holds: for any finite collections of points {xi}i=1k\{x_{i}\}_{i=1}^{k} and integers {ni}i=1k\{n_{i}\}_{i=1}^{k}, there exist integers m1,m2,mk1M(ε)m_{1},m_{2}\cdots,m_{k-1}\leq M(\varepsilon) and a point xKx\in K such that

d(ft+j=0i1(mj+nj)(x),ft(xi))<ε,for all t=1,2,,ni1 and all i=1,2,k,d\left(f^{t+\sum_{j=0}^{i-1}(m_{j}+n_{j})}(x),f^{t}(x_{i})\right)<\varepsilon,~{}~{}\text{for all $t=1,2,\cdots,n_{i}-1$ and all $i=1,2,\cdots k$},

where we set m0=n0=0m_{0}=n_{0}=0.

We refer the readers to [11, 12, 59, 58] for more studies of systems admitting specification property or the gluing orbit property.

3 Statements of main theorems

3.1 Multifractal analysis

Recall that for a Cr(r1)C^{r}(r\geq 1)-vector field X𝒳r(M)X\in\mathscr{X}^{r}(M), we denote by ϕtX:MM\phi_{t}^{X}\colon M\rightarrow M the CrC^{r}-flow generated by XX and by DϕtX{\rm D}\phi_{t}^{X} the tangent map of ϕtX\phi_{t}^{X}. We also use Φ=(ϕt)t\Phi=(\phi_{t})_{t} and Dϕt{\rm D}\phi_{t} for simplicity if there is no confusion. In the following, by M3M^{3} we mean a 33-dimensional closed smooth Riemannian manifold. We use the symbol “htop()h_{\rm top}(\cdot)” to denote the topological entropy of a set, see detailed definitions in Section 2.2. Recall that a subset \mathcal{R} of a topological space XX is residual if it contains a dense GδG_{\delta} subset of XX. 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 r𝒳r(M3),(r2)\mathcal{R}^{r}\subset\mathscr{X}^{r}(M^{3}),~{}(r\in\mathbb{N}_{\geq 2}) so that, if Λ\Lambda is a geometric Lorenz attractor of XrX\in\mathcal{R}^{r} and gC(Λ,)g\in C(\Lambda,\mathbb{R}), then either:

  1. 1.

    IgI_{g} is empty and gdμ=gdν\displaystyle\int g{\rm d}\mu=\int g{\rm d}\nu for all μ,νinv(Λ)\mu,\nu\in\mathcal{M}_{inv}(\Lambda), or

  2. 2.

    IgI_{g} is a residual subset of Λ\Lambda and htop(Ig)=htop(Λ)h_{\rm top}(I_{g})=h_{\rm top}(\Lambda).

Moreover, if IgI_{g} is non-empty then for any aa\in\mathbb{R} satisfying

infμinv(Λ)gdμ<a<supμinv(Λ)gdμ,\displaystyle\inf_{\mu\in\mathcal{M}_{inv}(\Lambda)}\int g{\rm d}\mu<a<\sup_{\mu\in\mathcal{M}_{inv}(\Lambda)}\int g{\rm d}\mu,

the level set Rg(a)R_{g}(a) is dense in Λ\Lambda and

htop(Rg(a))=sup{hμ(X):gdμ=a,μinv(Λ)}.h_{\rm top}(R_{g}(a))=\sup\left\{h_{\mu}(X)\colon\int g{\rm d}\mu=a,\mu\in\mathcal{M}_{inv}(\Lambda)\right\}.
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 ff satisfies that f>2f^{\prime}>\sqrt{2}, see Definition 2.5. This is because we have to use the eventually onto property of ff 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 C1C^{1}-generic vector fields.

Theorem B.

There exists a Baire residual subset 𝒳1(M)\mathcal{R}\subset\mathscr{X}^{1}(M) so that if Λ\Lambda is a singular hyperbolic attractor of XX\in\mathcal{R} and gC(Λ,)g\in C(\Lambda,\mathbb{R}) then either

  1. 1.

    IgI_{g} is empty and and gdμ=gdν\displaystyle\int g{\rm d}\mu=\int g{\rm d}\nu for all μ,νinv(Λ)\mu,\nu\in\mathcal{M}_{inv}(\Lambda), or

  2. 2.

    IgI_{g} is a residual subset of Λ\Lambda and htop(Ig)=htop(Λ)h_{\rm top}(I_{g})=h_{\rm top}(\Lambda).

Moreover, if IgI_{g} is non-empty then, for any aa\in\mathbb{R} satisfying

infμinv(Λ)gdμ<a<supμinv(Λ)gdμ,\displaystyle\inf_{\mu\in\mathcal{M}_{inv}(\Lambda)}\int g{\rm d}\mu<a<\sup_{\mu\in\mathcal{M}_{inv}(\Lambda)}\int g{\rm d}\mu,

the level set Rg(a)R_{g}(a) is dense in Λ\Lambda and

htop(Rg(a))=sup{hμ(X):gdμ=a,μinv(Λ)}.h_{\rm top}(R_{g}(a))=\sup\left\{h_{\mu}(X)\colon\int g{\rm d}\mu=a,\mu\in\mathcal{M}_{inv}(\Lambda)\right\}.

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. 1.

    It is clear from Theorem A that, for typical geometric Lorenz attractors, IgI_{g} is non-empty if and only if there exist μ,νinv(Λ)\mu,\nu\in\mathcal{M}_{inv}(\Lambda) so that gdμgdν\int g{\rm d}\mu\neq\int g{\rm d}\nu, a property satisfied by a C0C^{0}-open and dense set of continuous observables. In particular, if r𝒳r(M3),(r2)\mathcal{R}^{r}\subset\mathscr{X}^{r}(M^{3}),~{}(r\in\mathbb{N}_{\geq 2}) is the Baire residual subset given by Theorem A and Λ\Lambda is a geometric Lorenz attractor of XrX\in\mathcal{R}^{r} then

    {gC(Λ,):Ig is residual in Λ and htop(Ig)=htop(Λ)}\Big{\{}g\in C(\Lambda,\mathbb{R})\colon\text{$I_{g}$ is residual in $\Lambda$ and $h_{\rm top}(I_{g})=h_{\rm top}(\Lambda)$}\Big{\}}

    is open and dense in C(Λ,)C(\Lambda,\mathbb{R}). A similar conclusion holds in the context of Theorem B.

  2. 2.

    For the characterizations of level sets in Theorem B & Theorem A, when

    a=infμinv(Λ)gdμ or a=supμinv(Λ)gdμ,a=\displaystyle\inf_{\mu\in\mathcal{M}_{inv}(\Lambda)}\int g{\rm d}\mu\text{~{}~{}~{}or~{}~{}~{}}a=\displaystyle\sup_{\mu\in\mathcal{M}_{inv}(\Lambda)}\int g{\rm d}\mu,

    by [25, Theorem 1.1], one also has the variational principle that

    htop(Rg(a))=sup{hμ(X):gdμ=a,μinv(Λ)}.h_{\rm top}(R_{g}(a))=\sup\left\{h_{\mu}(X)\colon\int g{\rm d}\mu=a,\mu\in\mathcal{M}_{inv}(\Lambda)\right\}.

3.2 Large deviations

Recall that (Λ)\mathcal{M}(\Lambda) denotes the space of all probability measures supported on Λ\Lambda endowed with the weak*-topology. Our next result is a level-2 large deviations principle, which detects exponential convergence to equilibrium on the space (Λ)\mathcal{M}(\Lambda) for a singular hyperbolic attractor or a Lorenz attractor Λ\Lambda. We need to set further notations. Given t>0t>0, let t:Λ(Λ)\mathcal{E}_{t}\colon\Lambda\to\mathcal{M}(\Lambda) be the empirical measure function at time t, defined by

t(x):=1t0tδϕsX(x)ds.\mathcal{E}_{t}(x)\colon=\frac{1}{t}\int_{0}^{t}\delta_{\phi^{X}_{s}(x)}\,{\rm d}s.

In other words, t(x)\mathcal{E}_{t}(x) is the empirical probability on Λ\Lambda determined by the point xΛx\in\Lambda at time tt. We say that an invariant probability measure μ\mu satisfies a level-2 large deviations principle if there exists a lower-semicontinuous rate function :(Λ)[0,+]\mathfrak{I}\colon\mathcal{M}(\Lambda)\to[0,+\infty] so that

lim supt+1tlogμ({xΛ:t(x)𝒦})infν𝒦(ν)\displaystyle\limsup_{t\to+\infty}\frac{1}{t}\log\mu\Big{(}\Big{\{}x\in\Lambda\colon\mathcal{E}_{t}(x)\in\mathcal{K}\Big{\}}\Big{)}\leq-\inf_{\nu\in\mathcal{K}}\mathfrak{I}(\nu)

for any closed subset 𝒦(Λ)\mathcal{K}\subset\mathcal{M}(\Lambda), and

lim inft+1tlogμ({xΛ:t(x)𝒪})infν𝒪(ν)\displaystyle\liminf_{t\to+\infty}\frac{1}{t}\log\mu\Big{(}\Big{\{}x\in\Lambda\colon\mathcal{E}_{t}(x)\in\mathcal{O}\Big{\}}\Big{)}\geq-\inf_{\nu\in\mathcal{O}}\mathfrak{I}(\nu)

for any open subset 𝒪(Λ)\mathcal{O}\subset\mathcal{M}(\Lambda). If the latter holds then the contraction principle (see e.g. [22, Theorem 4.2.1]) ensures the level-1 large deviations principle

lim supt+1tlogμ({xΛ:1t0tg(ϕsX(x))ds[a,b]})inf{(ν):ν(Λ)&g𝑑ν[a,b]}\displaystyle\limsup_{t\to+\infty}\frac{1}{t}\log\mu\Big{(}\Big{\{}x\in\Lambda\colon\frac{1}{t}\int_{0}^{t}g(\phi^{X}_{s}(x))\,{\rm d}s\in[a,b]\Big{\}}\Big{)}\leq-\inf\Big{\{}\mathfrak{I}(\nu)\colon\nu\in\mathcal{M}(\Lambda)\,\&\,\int g\,d\nu\in[a,b]\Big{\}}

and

lim inft+1tlogμ({xΛ:1t0tg(ϕsX(x))ds(a,b)})inf{(ν):ν(Λ)&g𝑑ν(a,b)}\displaystyle\liminf_{t\to+\infty}\frac{1}{t}\log\mu\Big{(}\Big{\{}x\in\Lambda\colon\frac{1}{t}\int_{0}^{t}g(\phi^{X}_{s}(x))\,{\rm d}s\in(a,b)\Big{\}}\Big{)}\geq-\inf\Big{\{}\mathfrak{I}(\nu)\colon\nu\in\mathcal{M}(\Lambda)\,\&\,\int g\,d\nu\in(a,b)\Big{\}}

for each continuous observable g:Λg:\Lambda\to\mathbb{R}.

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 ψ:Λ\psi\colon\Lambda\to\mathbb{R}, define ψ:(Λ)[0,+]\mathfrak{I}_{\psi}\colon\mathcal{M}(\Lambda)\to[0,+\infty] by

ψ(μ)={Ptop(Λ,ψ)hμ(X)ψdμ,if μinv(Λ);+,otherwise. \mathfrak{I}_{\psi}(\mu)=\begin{cases}\begin{array}[]{cl}\displaystyle P_{\rm top}(\Lambda,\psi)-h_{\mu}(X)-\int{\psi}\,{\rm d}\mu&,\text{if }\mu\in\mathcal{M}_{inv}(\Lambda);\\ +\infty&,\text{otherwise. }\end{array}\end{cases}

When Λ\Lambda is singular hyperbolic, the entropy map h:inv(Λ),μhμ(X)h\colon\mathcal{M}_{inv}(\Lambda)\rightarrow\mathbb{R},~{}~{}~{}\mu\mapsto h_{\mu}(X) is upper semi-continuous [42], therefore ψ\mathfrak{I}_{\psi} is lower semi-continuous.

Theorem C.

(Level-2 large deviations) There exist a Baire residual subset r𝒳r(M3),(r2)\mathcal{R}^{r}\subset\mathcal{X}^{r}(M^{3}),~{}(r\in\mathbb{N}_{\geq 2}) and a Baire residual subset 𝒳1(M)\mathcal{R}\subset\mathcal{X}^{1}(M) such that if Λ\Lambda is a Lorenz attractor of XrX\in\mathcal{R}^{r} or Λ\Lambda is a singular hyperbolic attractor of XX\in\mathcal{R}, then the following properties are satisfied.
Assume μψ\mu_{\psi} is a weak Gibbs measure with respect to a Hölder continuous potential ψ:Λ\psi\colon\Lambda\to\mathbb{R} with ΛH\Lambda_{H} being the μψ\mu_{\psi}-full measure set such that (2.1) satisfies. Then one has:

  1. 1.

    (upper bound) There exists c0c_{\infty}\leq 0 so that

    lim supt\displaystyle\limsup_{t\to\infty} 1tlogμψ({xΛ:t(x)𝒦})max{infμ𝒦ψ(μ),c}\displaystyle\frac{1}{t}\log\mu_{\psi}\big{(}\{x\in\Lambda\colon\mathcal{E}_{t}(x)\in\mathcal{K}\}\big{)}\leq\max\Big{\{}-\inf_{\mu\in\mathcal{K}}\mathfrak{I}_{\psi}(\mu)\;,\;c_{\infty}\Big{\}}

    for any closed subset 𝒦(Λ)\mathcal{K}\subset\mathcal{M}(\Lambda).

  2. 2.

    (lower bound) If 𝒪(Λ)\mathcal{O}\subset\mathcal{M}(\Lambda) is an open set, then

    lim inft+1tlogμψ({xΛ:t(x)𝒪})inf{ψ(ν):ν𝒪erg(Λ),ν(ΛH)=1}.\displaystyle\liminf_{t\to+\infty}\frac{1}{t}\log\mu_{\psi}\Big{(}\Big{\{}x\in\Lambda\colon\mathcal{E}_{t}(x)\in\mathcal{O}\Big{\}}\Big{)}\geq\displaystyle-\inf~{}\big{\{}\mathfrak{I}_{\psi}(\nu)\colon\nu\in\mathcal{O}\cap\mathcal{M}_{erg}(\Lambda),\nu(\Lambda_{H})=1\big{\}}.
  3. 3.

    (lower bound for Gibbs measure) If μψ\mu_{\psi} is a Gibbs measure with respect to ψ\psi, then

    lim inft+1tlogμψ({xΛ:t(x)𝒪})infμ𝒪ψ(μ)\displaystyle\liminf_{t\to+\infty}\frac{1}{t}\log\mu_{\psi}\Big{(}\Big{\{}x\in\Lambda\colon\mathcal{E}_{t}(x)\in\mathcal{O}\Big{\}}\Big{)}\geq-\inf_{\mu\in\mathcal{O}}\mathfrak{I}_{\psi}(\mu)

    for any open subset 𝒪(Λ)\mathcal{O}\subset\mathcal{M}(\Lambda).

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 cc_{\infty}, defined by (6.6), measures tails of constants appearing in the concept of weak Gibbs measure. If μψ\mu_{\psi} is a (strong) Gibbs measure then: (i) c=c_{\infty}=-\infty and the upper bound in the theorem reduces to infμ𝒦ψ(μ)-\inf_{\mu\in\mathcal{K}}\mathfrak{I}_{\psi}(\mu), and (ii) there is no need to restrict to probability measures supported on ΛH\Lambda_{H} and the lower bound reduces to infμ𝒪ψ(μ)-\inf_{\mu\in\mathcal{O}}\mathfrak{I}_{\psi}(\mu) (item 3). The function ψ(μ)\mathfrak{I}_{\psi}(\mu) 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 Λ\Lambda is a singular hyperbolic homoclinic class such that each pair of periodic orbits are homoclinically related and 1(Λ)¯=inv(Λ)\overline{\mathcal{M}_{1}(\Lambda)}=\mathcal{M}_{inv}(\Lambda), then the conclusions of Theorem AB 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 Λ\Lambda is a Lorenz attractor of vector fields in a Baire residual subset r𝒳r(M3),(r2)\mathcal{R}^{r}\subset\mathcal{X}^{r}(M^{3}),(r\in\mathbb{N}_{\geq 2}) or Λ\Lambda is a singular hyperbolic attractor Λ\Lambda of vector fields in a Baire residual set 𝒳1(M)\mathcal{R}\subset\mathcal{X}^{1}(M), then Λ\Lambda 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 1(Λ)¯=inv(Λ)\overline{\mathcal{M}_{1}(\Lambda)}=\mathcal{M}_{inv}(\Lambda) (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 Λ\Lambda of a vector field X𝒳1(M)X\in\mathscr{X}^{1}(M), we denote by inv(Λ)\mathcal{M}_{inv}(\Lambda) and erg(Λ)\mathcal{M}_{erg}(\Lambda) the space of invariant and ergodic probability measures supported on Λ\Lambda, respectively, and that dd^{*} is a translation invariant metric on (Λ)\mathcal{M}(\Lambda) compatible with the weak topology.

Definition 4.1.

Given X𝒳1(M)X\in\mathscr{X}^{1}(M) and an invariant compact set Λ\Lambda. We say a convex subset inv(Λ)\mathcal{M}\subseteq\mathcal{M}_{inv}(\Lambda) is entropy-dense if for any ε>0\varepsilon>0 and any μ\mu\in\mathcal{M}, there exists νerg(Λ)\nu\in\mathcal{M}_{erg}(\Lambda) satisfying

d(μ,ν)<εandhν(X)>hμ(X)ε.d^{*}(\mu,\nu)<\varepsilon~{}~{}~{}\text{and}~{}~{}~{}h_{\nu}(X)>h_{\mu}(X)-\varepsilon.

It follows from the definition that the entropy denseness property is hereditary, i.e. if 12inv(Λ)\mathcal{M}_{1}\subset\mathcal{M}_{2}\subset\mathcal{M}_{inv}(\Lambda) are convex and 2\mathcal{M}_{2} is entropy dense, so is 1\mathcal{M}_{1}. 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 inv(Λ)\mathcal{M}_{inv}(\Lambda). We now recall the definition of horseshoe.

Definition 4.2.

Given X𝒳1(M)X\in\mathscr{X}^{1}(M), an invariant compact set Λ\Lambda is called a basic set if it is, transitive, hyperbolic and locally maximal, i. e. there exists an open neighborhood UU of Λ\Lambda, such that Λ=tϕt(U)\Lambda=\cap_{t\in{\mathbb{R}}}\phi_{t}(U). A basic set Λ\Lambda is called a horseshoe if Λ\Lambda is a proper subset of MM, 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 Λ\Lambda is a horseshoe of X𝒳1(M)X\in\mathcal{X}^{1}(M). Then there exists a transitive subshift of finite type (Σ,σ)(\Sigma,\sigma) and a continuous roof function ρ:Σ+\rho\colon\Sigma\rightarrow\mathbb{R}^{+}, such that (Λ,Φ)(\Lambda,\Phi) is semi-conjugate to the suspension (Σρ,𝔉)(\Sigma_{\rho},\mathfrak{F}) through a continuous surjective map π:ΣρΛ\pi\colon\Sigma_{\rho}\rightarrow\Lambda, where Φ=(ϕt)t\Phi=(\phi_{t})_{t} is the flow generated by XX and 𝔉=(σt)t\mathfrak{F}=(\sigma_{t})_{t} is the suspension flow. Moreover, the semi-conjugacy π\pi is finite-to-one and hence preserves entropy:

  1. 1.

    for any μinv(Σρ,𝔉)\mu\in\mathcal{M}_{inv}(\Sigma_{\rho},\mathfrak{F}), there exists a unique νinv(Λ)\nu\in\mathcal{M}_{inv}(\Lambda) satisfying μ=π(ν)\mu=\pi^{*}(\nu) and their metric entropies coincide hμ(𝔉)=hν(X)h_{\mu}(\mathfrak{F})=h_{\nu}(X);

  2. 2.

    for any invariant set ZZ of (Σρ,𝔉)(\Sigma_{\rho},\mathfrak{F}), it satisfies htop(Z,𝔉)=htop(π(Z),X)h_{\rm top}(Z,\mathfrak{F})=h_{\rm top}(\pi(Z),X);

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 Λ\Lambda is a horseshoe of X𝒳1(M)X\in\mathcal{X}^{1}(M). The following properties are satisfied.

  1. 1.

    For any μinv(Λ)\mu\in\mathcal{M}_{inv}(\Lambda), the set Gμ:={xM:limt1t0tδϕs(x)ds=μ}\displaystyle G_{\mu}:=\left\{x\in M\colon\lim_{t\rightarrow\infty}\frac{1}{t}\int_{0}^{t}\delta_{\phi_{s}(x)}{\rm d}s=\mu\right\} of μ\mu-generic points is non-empty;

  2. 2.

    There exists a residual subset 𝒢inv(Λ)\mathcal{G}\subset\mathcal{M}_{inv}(\Lambda) such that each μ𝒢\mu\in\mathcal{G} is ergodic and Supp(μ)=Λ\operatorname{\,Supp}(\mu)=\Lambda.

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 λ\lambda-lemma [43, Lemma 7.1] (also see [64, Theorem 5.1]) and is classical, thus we omit it.

Lemma 4.5.

Let Λ1\Lambda_{1} and Λ2\Lambda_{2} be two horseshoes of X𝒳1(M)X\in\mathcal{X}^{1}(M). Assume there exists hyperbolic periodic points p1Λ1p_{1}\in\Lambda_{1} and p2Λ2p_{2}\in\Lambda_{2} such that Orb(p1)\operatorname{Orb}(p_{1}) and Orb(p2)\operatorname{Orb}(p_{2}) are homoclinically related. Then there exists a larger horseshoe Λ\Lambda that contains both Λ1\Lambda_{1} and Λ2\Lambda_{2}.

The next proposition ensures that horseshoes are entropy-dense.

Proposition 4.6.

Assume that X𝒳1(M)X\in\mathscr{X}^{1}(M). If Λ\Lambda is a horseshoe then inv(Λ)\mathcal{M}_{inv}(\Lambda) is entropy-dense.

Proof.

Although the result is probably known we shall include a proof as we could not find a reference. Assume that Λ\Lambda is a horseshoe, by Theorem 4.3, it is semiconjugate to a suspension flow 𝔉\mathfrak{F} over a transitive subshift of finite type (Σ,σ)(\Sigma,\sigma) with a continuous roof function ρ\rho. As the semi-conjugacy preserves entropy, we may deal directly with the case of the suspension flow 𝔉\mathfrak{F}. Moreover, since the entropy map is upper-semicontinuous and σ\sigma satisfies the gluing orbit property (also known as transitive specification), Theorem B in [24] guarantees that σ\sigma is entropy-dense.

Fix ε>0\varepsilon>0 and an arbitrary 𝔉\mathfrak{F}-invariant probability measure μρ\mu_{\rho}, determined by a σ\sigma-invariant probability measure μ\mu (recall Subsection 2.4). Take δ>0\delta>0 small, to be determined later. Since the subshift of finite type σ\sigma is entropy-dense, one picks a σ\sigma-invariant and ergodic probability ν\nu so that

  • d(μ,ν)d^{*}(\mu,\nu) is small enough such that

    1δ<ρdμρdν<1+δ.1-\delta<\frac{\int\rho{\rm d}\mu}{\int\rho{\rm d}\nu}<1+\delta.
  • hν(σ)>hμ(σ)δ.h_{\nu}(\sigma)>h_{\mu}(\sigma)-\delta.

Then the Abramov formula for the 𝔉\mathfrak{F}-invariant probabilities μρ\mu_{\rho} and νρ\nu_{\rho} (recall Subsection 2.4) ensures that

hνρ(𝔉)=hν(σ)ρdν>ρdμρdν(hμρ(𝔉)δρdμ)>(1δ)(hμρ(𝔉)δρdμ)>hμρ(𝔉)ε\displaystyle h_{\nu_{\rho}}(\mathfrak{F})=\frac{h_{\nu}(\sigma)}{\int\rho{\rm d}\nu}>\frac{\int\rho{\rm d}\mu}{\int\rho{\rm d}\nu}\cdot\Big{(}h_{\mu_{\rho}}(\mathfrak{F})-\frac{\delta}{\int\rho{\rm d}\mu}\Big{)}>(1-\delta)\cdot\Big{(}h_{\mu_{\rho}}(\mathfrak{F})-\frac{\delta}{\int\rho{\rm d}\mu}\Big{)}>h_{\mu_{\rho}}(\mathfrak{F})-\varepsilon

provided that δ\delta is small enough. Diminishing δ\delta if necessary we may also get that μρ\mu_{\rho} and νρ\nu_{\rho} are also ε\varepsilon-close in the dd^{*}-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 X𝒳1(M)X\in\mathscr{X}^{1}(M) and an invariant compact set Λ\Lambda, we say a convex subset inv(Λ)\mathcal{M}\subseteq\mathcal{M}_{inv}(\Lambda) has the horseshoe approximation property if for each ε>0\varepsilon>0 and any μ\mu\in\mathcal{M}, there exist a horseshoe ΛΛ\Lambda^{\prime}\subset\Lambda and νerg(Λ)\nu\in\mathcal{M}_{erg}(\Lambda^{\prime})

d(ν,μ)<εandhν(X)>hμ(X)ε.d^{*}(\nu,\mu)<\varepsilon~{}~{}~{}\text{and}~{}~{}~{}h_{\nu}(X)>h_{\mu}(X)-\varepsilon. (4.1)
Remark 4.8.

By Proposition 4.6, if Λ\Lambda is a horseshoe, then inv(Λ)\mathcal{M}_{inv}(\Lambda) admits the horseshoe approximation property naturally. Moreover, it is clear from (4.1) in the previous definition, that the horseshoe Λ\Lambda^{\prime} satisfies htop(Λ)>hμ(X)εh_{\rm top}(\Lambda^{\prime})>h_{\mu}(X)-\varepsilon. Finally, in some specific contexts the approximating horseshoe Λ\Lambda^{\prime} can be constructed using an analogue of Katok’s arguments [30] and all measures supported on Λ\Lambda^{\prime} are within an ε\varepsilon-neighborhood (in the weak topology) of the original probability μ\mu (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 Λ\Lambda of X𝒳1(M)X\in\mathscr{X}^{1}(M), denote by per(Λ)\mathcal{M}_{per}(\Lambda) the set of periodic measures supported on Λ\Lambda and set

1(Λ)={μinv(Λ):μ(Sing(Λ))=0}and0(Λ)=erg(Λ)1(Λ).\mathcal{M}_{1}(\Lambda)=\left\{\mu\in\mathcal{M}_{inv}(\Lambda):\mu(\operatorname{Sing}(\Lambda))=0\right\}~{}~{}\text{and}~{}~{}\mathcal{M}_{0}(\Lambda)=\mathcal{M}_{erg}(\Lambda)\cap\mathcal{M}_{1}(\Lambda).

We first prove the following auxiliary lemma.

Lemma 4.9.

Let X𝒳1(M)X\in\mathscr{X}^{1}(M) and Λ\Lambda be a singular hyperbolic homoclinic class of XX. If each pair of periodic orbits of Λ\Lambda are homoclinic related, then

per(Λ)¯=Convex(0(Λ))¯=1(Λ)¯,\overline{\mathcal{M}_{per}(\Lambda)}=\overline{\emph{Convex}(\mathcal{M}_{0}(\Lambda))}=\overline{\mathcal{M}_{1}(\Lambda)},

where Convex(0(Λ))\emph{Convex}(\mathcal{M}_{0}(\Lambda)) is the convex hull of 0(Λ)\mathcal{M}_{0}(\Lambda).

Proof.

As the second equality above is immediate from the ergodic decomposition theorem we are left to prove the first one. By [52, Proposition 3.1], 0(Λ)per(Λ)¯\mathcal{M}_{0}(\Lambda)\subset\overline{\mathcal{M}_{per}(\Lambda)}. Since Λ\Lambda is a homoclinic class, the set per(Λ)¯\overline{\mathcal{M}_{per}(\Lambda)} is convex by [1, Proposition 4.7& Remark 4.6]. Hence we have that Convex(0(Λ))¯per(Λ)¯\overline{\text{Convex}(\mathcal{M}_{0}(\Lambda))}\subset\overline{\mathcal{M}_{per}(\Lambda)}. The inclusion per(Λ)¯Convex(0)¯\overline{\mathcal{M}_{per}(\Lambda)}\subset\overline{\text{Convex}(\mathcal{M}_{0})} is obvious since per(Λ)0(Λ)\mathcal{M}_{per}(\Lambda)\subset\mathcal{M}_{0}(\Lambda). ∎

Given X𝒳1(M)X\in\mathscr{X}^{1}(M), let Λ\Lambda be an invariant compact set displaying a singular hyperbolic splitting TΛM=EssEcuT_{\Lambda}M=E^{ss}\oplus E^{cu}. Then for any ergodic measure μ0(Λ)\mu\in\mathcal{M}_{0}(\Lambda), the splitting EssEcu|Supp(μ)E^{ss}\oplus E^{cu}|_{\operatorname{\,Supp}(\mu)} is a dominated splitting and the index of μ\mu equals dim(Ess)\dim(E^{ss}) obviously. Using Katok’s arguments in [30, Theorem 4.3] one obtains the following lemma. Given X𝒳1(M)X\in\mathscr{X}^{1}(M) and assume Λ\Lambda is an invariant compact set. We say that XX satisfies the star condition on Λ\Lambda if there exist a neighborhood UU of Λ\Lambda and a C1C^{1}-neighborhood 𝒰\mathcal{U} of XX in 𝒳1(M)\mathscr{X}^{1}(M) such that every periodic orbit contained in UU associated to a vector field Y𝒰Y\in\mathcal{U} is hyperbolic.

Lemma 4.10.

Let X𝒳1(M)X\in\mathscr{X}^{1}(M) and Λ\Lambda be a singular hyperbolic homoclinic class of XX. Assume each pair of periodic orbits of Λ\Lambda are homoclinic related, and μ0(Λ)\mu\in\mathcal{M}_{0}(\Lambda). Then for any ε>0\varepsilon>0 there exists a horseshoe ΛεΛ\Lambda_{\varepsilon}\subset\Lambda contained in the ε\varepsilon-neighborhood of Supp(μ)\operatorname{\,Supp}(\mu) (in the Hausdorff distance), and so that d(μ,ν)<εd^{*}(\mu,\nu)<\varepsilon for any νinv(Λε)\nu\in\mathcal{M}_{inv}(\Lambda_{\varepsilon}), and there exists ν0erg(Λε)\nu_{0}\in\mathcal{M}_{erg}(\Lambda_{\varepsilon}) satisfying hν0(X)>hμ(X)εh_{\nu_{0}}(X)>h_{\mu}(X)-\varepsilon.

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 Λ\Lambda is a singular hyperbolic homoclinic class, the vector field XX satisfies the star condition over Λ\Lambda. The existence of a horseshoe ΛεΛ\Lambda_{\varepsilon}\subset\Lambda satisfying htop(Λε)>hμ(X)εh_{\rm top}(\Lambda_{\varepsilon})>h_{\mu}(X)-\varepsilon follows directly from [33, Proposition 2.9]. Moreover, the horseshoe Λε\Lambda_{\varepsilon} is constructed by shadowing the orbit of a generic point of μ\mu, following the classical arguments of Katok [30, Theorem 4.3], hence Λε\Lambda_{\varepsilon} is contained in the ε\varepsilon-neighborhood of Supp(μ)\operatorname{\,Supp}(\mu) and every invariant measure ν\nu supported on Λε\Lambda_{\varepsilon} is close to μ\mu in the weak-topology. Finally, by the variational principle and the fact that htop(Λε)>hμ(X)εh_{\rm top}(\Lambda_{\varepsilon})>h_{\mu}(X)-\varepsilon, there exists ν0erg(Λε)\nu_{0}\in\mathcal{M}_{erg}(\Lambda_{\varepsilon}) satisfying hν0(X)>hμ(X)εh_{\nu_{0}}(X)>h_{\mu}(X)-\varepsilon. ∎

Remark 4.11.

In fact, the horseshoe Λε\Lambda_{\varepsilon} obtained in Lemma 4.10 is conjugate (not only semi-conjugate) to the suspension flow of a full shift with continuous roof function. We refer the reader to [33, Proposition 2.9] for more details and to [51, Theorem 5.6] for an approach.

The approximation by horseshoes in the conclusion of Lemma 4.10 can actually be extended to arbitrary invariant measures in 1(Λ)\mathcal{M}_{1}(\Lambda). More precisely:

Proposition 4.12.

Let X𝒳1(M)X\in\mathscr{X}^{1}(M) and Λ\Lambda be a singular hyperbolic homoclinic class of XX. If each pair of periodic orbits of Λ\Lambda are homoclinic related, then for every μ1(Λ)\mu\in\mathcal{M}_{1}(\Lambda) and ε>0\varepsilon>0, there exists a horseshoe ΛεΛ\Lambda_{\varepsilon}\subset\Lambda and there exists νerg(Λε)\nu\in\mathcal{M}_{erg}(\Lambda_{\varepsilon}) satisfying

d(μ,ν)<εandhν(X)>hμ(X)ε.d^{*}(\mu,\nu)<\varepsilon~{}~{}\text{and}~{}~{}h_{\nu}(X)>h_{\mu}(X)-\varepsilon.

Thus 1(Λ){\mathcal{M}_{1}(\Lambda)} has the horseshoe approximation property, and so it is entropy-dense.

Proof.

Let μ1(Λ)\mu\in\mathcal{M}_{1}(\Lambda) and ε>0\varepsilon>0 be fixed. By ergodic decomposition and affinity of the metric entropy, there exist ergodic measures ω1,ω2,,ωk0(Λ)\omega_{1},\omega_{2},\cdots,\omega_{k}\in\mathcal{M}_{0}(\Lambda) and real numbers α1,α2,,αk(0,1)\alpha_{1},\alpha_{2},\cdots,\alpha_{k}\in(0,1) with i=1kαi=1\sum\limits_{i=1}^{k}\alpha_{i}=1 so that the probability ω=i=1kαiωi\omega^{\prime}=\sum\limits_{i=1}^{k}\alpha_{i}\omega_{i} satisfies

d(ω,μ)<ε3andhω(X)>hμ(X)ε3.d^{*}(\omega^{\prime},\mu)<\frac{\varepsilon}{3}~{}~{}\text{and}~{}~{}h_{\omega^{\prime}}(X)>h_{\mu}(X)-\frac{\varepsilon}{3}.

By Lemma 4.10, for each i{1,2,,k}i\in\{1,2,\cdots,k\}, there exists a horseshoe ΛiΛ\Lambda_{i}\subset\Lambda such that every ωerg(Λi)\omega\in\mathcal{M}_{erg}(\Lambda_{i}) satisfies d(ωi,ω)<ε6d^{*}(\omega_{i},\omega)<\frac{\varepsilon}{6}, and there exists νierg(Λi)\nu_{i}\in\mathcal{M}_{erg}(\Lambda_{i}) satisfying

d(ωi,νi)<ε3andhνi(X)>hωi(X)ε3.d^{*}(\omega_{i},\nu_{i})<\frac{\varepsilon}{3}~{}~{}\text{and}~{}~{}h_{\nu_{i}}(X)>h_{\omega_{i}}(X)-\frac{\varepsilon}{3}.

Let Λε\Lambda_{\varepsilon} be a large horseshoe that contains every Λi\Lambda_{i} for i{1,2,,k}i\in\{1,2,\cdots,k\}. Such Λε\Lambda_{\varepsilon} does exist since any two periodic orbits in Λ\Lambda are homoclinically related and any horseshoe must contain (countably many) periodic orbits. Then the probability measure ν=i=1kαiνi\nu^{\prime}=\sum\limits_{i=1}^{k}\alpha_{i}\nu_{i} in inv(Λε)\mathcal{M}_{inv}(\Lambda_{\varepsilon}) satisfies d(ν,μ)d(ν,ω)+d(ω,μ)<2ε3,d^{*}(\nu^{\prime},\mu)\leq d^{*}(\nu^{\prime},\omega^{\prime})+d^{*}(\omega^{\prime},\mu)<\frac{2\varepsilon}{3}, and

hν(X)=i=1kαihνi(X)>i=1kαihωi(X)ε3>hμ(X)2ε3.h_{\nu^{\prime}}(X)=\sum\limits_{i=1}^{k}\alpha_{i}\cdot h_{\nu_{i}}(X)>\sum\limits_{i=1}^{k}\alpha_{i}\cdot h_{\omega_{i}}(X)-\frac{\varepsilon}{3}>h_{\mu}(X)-\frac{2\varepsilon}{3}.

By Proposition 4.6, we know that inv(Λε)\mathcal{M}_{inv}(\Lambda_{\varepsilon}) is entropy-dense. Thus, for the invariant measure νinv(Λε)\nu^{\prime}\in\mathcal{M}_{inv}(\Lambda_{\varepsilon}) and ε>0\varepsilon>0 above, there exists νerg(Λε)\nu\in\mathcal{M}_{erg}(\Lambda_{\varepsilon}) satisfying

d(ν,ν)<ε3andhν(X)>hν(X)ε3.d^{*}(\nu^{\prime},\nu)<\frac{\varepsilon}{3}~{}~{}\text{and}~{}~{}h_{\nu}(X)>h_{\nu^{\prime}}(X)-\frac{\varepsilon}{3}.

In consequence, the ergodic probability ν\nu supported on Λε\Lambda_{\varepsilon} satisfies

d(μ,ν)d(μ,ν)+d(ν,ν)<ε,d^{*}(\mu,\nu)\leq d^{*}(\mu,\nu^{\prime})+d^{*}(\nu^{\prime},\nu)<\varepsilon,

and

hν(X)>hν(X)ε3>hμ(X)ε.h_{\nu}(X)>h_{\nu^{\prime}}(X)-\frac{\varepsilon}{3}>h_{\mu}(X)-\varepsilon.

Proposition 4.13.

Let X𝒳1(M)X\in\mathscr{X}^{1}(M) and Λ\Lambda be a singular hyperbolic homoclinic class of XX. Assume each pair of periodic orbits of Λ\Lambda are homoclinic related, and inv(Λ)=1(Λ)¯\mathcal{M}_{inv}(\Lambda)=\overline{\mathcal{M}_{1}(\Lambda)}. Then inv(Λ)\mathcal{M}_{inv}(\Lambda) has the horseshoe approximation property.

Proof.

Fix an arbitrary μinv(Λ)=1(Λ)¯\mu\in\mathcal{M}_{inv}(\Lambda)=\overline{\mathcal{M}_{1}(\Lambda)}. Using Proposition 4.12 it suffices to show that μ\mu is approximated, both in the weak topology and entropy, by measures in 1(Λ){\mathcal{M}_{1}(\Lambda)}.

Fix an arbitrary constant ε>0\varepsilon>0. If μ1(Λ)\mu\in{\mathcal{M}_{1}(\Lambda)} there is nothing to prove. Otherwise, one can write μ=αμ1+(1α)μ2\mu=\alpha\mu_{1}+(1-\alpha)\mu_{2} for some 0<α10<\alpha\leq 1 and probabilities μ1,μ2inv(Λ)\mu_{1},\mu_{2}\in\mathcal{M}_{inv}(\Lambda) so that μ1(Sing(X))=1\mu_{1}(\operatorname{Sing}(X))=1 and μ2(Sing(X))=0\mu_{2}(\operatorname{Sing}(X))=0. In other words, μ1\mu_{1} is supported in the invariant set formed by singularities and μ21(Λ)\mu_{2}\in\mathcal{M}_{1}(\Lambda). By the assumption, there exists a sequence of probabilities νn1(Λ)\nu_{n}\in{\mathcal{M}_{1}(\Lambda)} with d(νn,μ1)0d^{*}(\nu_{n},\mu_{1})\rightarrow 0 as nn\rightarrow\infty. Notice that 1(Λ)\mathcal{M}_{1}(\Lambda) is a convex set. Thus ανn+(1α)μ21(Λ)\alpha\nu_{n}+(1-\alpha)\mu_{2}\in\mathcal{M}_{1}(\Lambda) for each n1n\geq 1. Moreover, using that dd^{*} is translation invariance and affinity of the entropy we get

d(ανn+(1α)μ2,μ)=d(ανn+(1α)μ2,αμ1+(1α)μ2)=d(ανn,αμ1)\displaystyle d^{*}\big{(}\alpha\nu_{n}+(1-\alpha)\mu_{2},\mu\big{)}=d^{*}\big{(}\alpha\nu_{n}+(1-\alpha)\mu_{2},\alpha\mu_{1}+(1-\alpha)\mu_{2}\big{)}=d^{*}\big{(}\alpha\nu_{n},\alpha\mu_{1}\big{)}

and

hανn+(1α)μ2(X)=αhνn(X)+(1α)hμ2(X)(1α)hμ2(X)=hμ(X).h_{\alpha\nu_{n}+(1-\alpha)\mu_{2}}(X)=\alpha h_{\nu_{n}}(X)+(1-\alpha)h_{\mu_{2}}(X)\geq(1-\alpha)h_{\mu_{2}}(X)=h_{\mu}(X).

By taking n1n\geq 1 large, we conclude that the probability μn:=ανn+(1α)μ21(Λ)\mu_{n}\colon=\alpha\nu_{n}+(1-\alpha)\mu_{2}\in\mathcal{M}_{1}(\Lambda) satisfies d(μn,μ)<εandhμn(X)hμ(X).d^{*}(\mu_{n},\mu)<{\varepsilon}~{}~{}\text{and}~{}~{}h_{\mu_{n}}(X)\geq h_{\mu}(X). This completes the proof. ∎

Recently, S. Crovisier and D. Yang [20] proved that for C1C^{1}-open and dense set of vector field X𝒳1(M)X\in\mathcal{X}^{1}(M), any singular hyperbolic attractor Λ\Lambda is a robustly transitive attractor. Moreover, if Λ\Lambda is non-trivial, then it is a homoclinic class and any two periodic orbits contained in Λ\Lambda are homoclinically related. On the other hand, the main theorems (Theorem A, B, B’) in [52] state that if Λ\Lambda is a singular hyperbolic attractor of XX in a residual subset 𝒳1(M)\mathcal{R}\subset\mathscr{X}^{1}(M), or Λ\Lambda is a geometric Lorenz attractor of XX in a residual subset r𝒳r(M3)\mathcal{R}^{r}\subset\mathscr{X}^{r}(M^{3}), then inv(Λ)=per(Λ)¯\mathcal{M}_{inv}(\Lambda)=\overline{\mathcal{M}_{per}(\Lambda)}, which implies naturally that inv(Λ)=1(Λ)¯\mathcal{M}_{inv}(\Lambda)=\overline{\mathcal{M}_{1}(\Lambda)}. Thus one obtains the following consequence from Proposition 4.13.

Corollary 4.14.

The following holds:

  1. 1.

    There exists a residual subset r𝒳r(M3)\mathcal{R}^{r}\subset\mathscr{X}^{r}(M^{3}) where r2{}r\in\mathbb{N}_{\geq 2}\cup\{\infty\} such that if Λ\Lambda is a geometric Lorenz attractor of XrX\in\mathcal{R}^{r}, then inv(Λ)=1(Λ)¯\mathcal{M}_{inv}(\Lambda)=\overline{\mathcal{M}_{1}(\Lambda)} and thus inv(Λ)\mathcal{M}_{inv}(\Lambda) has the horseshoe approximation property and is entropy-dense.

  2. 2.

    There exists a residual subset 𝒳1(M)\mathcal{R}\subset\mathscr{X}^{1}(M) such that if Λ\Lambda is a non-trivial singular hyperbolic attractor of XX\in\mathcal{R}, then inv(Λ)=1(Λ)¯\mathcal{M}_{inv}(\Lambda)=\overline{\mathcal{M}_{1}(\Lambda)} and thus inv(Λ)\mathcal{M}_{inv}(\Lambda) 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 X𝒳1(M)X\in\mathscr{X}^{1}(M) and Λ\Lambda be a singular hyperbolic homoclinic class of XX. If each pair of periodic orbits of Λ\Lambda are homoclinic related, then for every μ1(Λ)¯\mu\in\overline{\mathcal{M}_{1}(\Lambda)} and ε>0\varepsilon>0, there exist a horseshoe ΛεΛ\Lambda_{\varepsilon}\subseteq\Lambda and νerg(Λε)\nu\in\mathcal{M}_{erg}(\Lambda_{\varepsilon}) so that htop(Λε)>htop(Λ)εh_{\rm top}(\Lambda_{\varepsilon})>h_{\rm top}(\Lambda)-\varepsilon and d(ν,μ)<εd^{*}(\nu,\mu)<\varepsilon.

Proof.

In the special case that ΛSing(X)=\Lambda\cap\operatorname{Sing}(X)=\emptyset we have that Λ\Lambda hyperbolic (recall Remark 2.3), hence Λ\Lambda itself is a basic set of XX. Moreover, if this is the case then inv(Λ)=1(Λ)=per(Λ)¯\mathcal{M}_{inv}(\Lambda)=\mathcal{M}_{1}(\Lambda)=\overline{\mathcal{M}_{per}(\Lambda)} by [54, Theorem 1]. Thus 1(Λ)¯\overline{\mathcal{M}_{1}(\Lambda)} has the strong horseshoe approximation property by Proposition 4.13 and one concludes.

It remains to consider the case where ΛSing(X)\Lambda\cap\operatorname{Sing}(X)\neq\emptyset. We first construct a nested sequence of horseshoes whose entropies approximate to htop(Λ)h_{\rm top}(\Lambda). Notice first that every periodic orbit in Λ\Lambda is hyperbolic, hence the non-trivial homoclinic class Λ\Lambda contains countably many periodic orbits which we list as {γn}n1\{\gamma_{n}\}_{n\geq 1}. Moreover, Λ\Lambda being a non-trivial homoclinic class ensures that htop(Λ)>0h_{\rm top}(\Lambda)>0. By the variational principle and Lemma 4.10, there is a sequence of horseshoes {Δn}n1\{\Delta_{n}\}_{n\geq 1} contained in Λ\Lambda such that htop(Δn)>htop(Λ)1/nh_{\rm top}(\Delta_{n})>h_{\rm top}(\Lambda)-1/n. Notice that any two periodic orbits contained in Λ\Lambda have the same stable index and are homoclinically related with each other, and each horseshoe Δn\Delta_{n} must contain infinitely many periodic orbits for each n1n\geq 1. Thus, inductively, we can construct a nested sequence of transitive horseshoes {Λn}n1\{\Lambda_{n}\}_{n\geq 1} contained in Λ\Lambda as follows:

  • Let Λ1\Lambda_{1} be a horseshoe that contains γ1\gamma_{1} and Δ1\Delta_{1}. Such Λ1\Lambda_{1} does exist since γ1\gamma_{1} is homoclinically related with all periodic orbits contained in Δ1\Delta_{1}.

  • For n2n\geq 2, let Λn\Lambda_{n} be a horseshoe that contains γn\gamma_{n} and also contains the two horseshoes Λn1\Lambda_{n-1} and Δn\Delta_{n}. Such a horseshoe exists by Lemma 4.5.

Then we have that htop(Λn)htop(Δn)>htop(Λ)1/nh_{\rm top}(\Lambda_{n})\geq h_{\rm top}(\Delta_{n})>h_{\rm top}(\Lambda)-1/n for every n1n\geq 1.

We claim that that the previous sequence of horseshoes {Λn}n\{\Lambda_{n}\}_{n\in{\mathbb{N}}} satisfies the conclusion. Indeed, for any ε>0\varepsilon>0 and μ1(Λ)¯\mu\in\overline{\mathcal{M}_{1}(\Lambda)}, by Lemma 4.9 there exists n1n_{1}\in\mathbb{N} such that d(μ,νn1)<εd^{*}(\mu,\nu_{n_{1}})<\varepsilon where νn1\nu_{n_{1}} is the periodic measure associated to γn1\gamma_{n_{1}}. Take n2n_{2}\in\mathbb{N} such that 1/n2<ε1/n_{2}<\varepsilon and n=max{n1,n2}n=\max\{n_{1},n_{2}\}. Therefore νn1erg(Λn)\nu_{n_{1}}\in\mathcal{M}_{erg}(\Lambda_{n}) and

htop(Λn)>htop(Λ)1/nhμ(X)1/n>hμ(X)ε.h_{\rm top}(\Lambda_{n})>h_{\rm top}(\Lambda)-1/n\geq h_{\mu}(X)-1/n>h_{\mu}(X)-\varepsilon.

This completes the proof of Proposition 4.15. ∎

5 Multifractal analysis

In this section, we aim to study the multifractal analysis of singular hyperbolic attractors of C1C^{1}-generic vector fields and geometric Lorenz attractors of CrC^{r}-generic vector fields (r2r\geq 2). We prove the following theorem in general case. With the arguments in Section 4.2, one will see that Theorems AB 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 X𝒳1(M)X\in\mathscr{X}^{1}(M) and Λ\Lambda be a singular hyperbolic homoclinic class of XX. Assume each pair of periodic orbits contained in Λ\Lambda are homoclinically related and inv(Λ)=1(Λ)¯\mathcal{M}_{inv}(\Lambda)=\overline{\mathcal{M}_{1}(\Lambda)}. Given gC(Λ,)g\in C(\Lambda,\mathbb{R}), then either

  1. 1.

    IgI_{g} is empty and gdμ=gdν\displaystyle\int g{\rm d}\mu=\displaystyle\int g{\rm d}\nu for all μ,νinv(Λ)\mu,\nu\in\mathcal{M}_{inv}(\Lambda); or

  2. 2.

    IgI_{g} is residual in Λ\Lambda and htop(Ig)=htop(Λ)h_{\rm top}(I_{g})=h_{\rm top}(\Lambda).

Moreover, for each aa\in{\mathbb{R}} satisfying infμ1(Λ)gdμ<a<supμ1(Λ)gdμ,\displaystyle\inf_{\mu\in\mathcal{M}_{1}(\Lambda)}\int g{\rm d}\mu<a<\sup_{\mu\in\mathcal{M}_{1}(\Lambda)}\int g{\rm d}\mu, the level set Rg(a)R_{g}(a) is dense in Λ\Lambda and

htop(Rg(a))=sup{hμ(f):gdμ=a,μinv(Λ)}.h_{\rm top}(R_{g}(a))=\sup\left\{h_{\mu}(f)\colon\int g{\rm d}\mu=a,\mu\in\mathcal{M}_{inv}(\Lambda)\right\}.

Furthermore the set of functions satisfying the second item form an open and dense subset in C(Λ,)C(\Lambda,\mathbb{R}).

Remark 5.2.
  1. 1.

    Similarly to Remark 3.2, for the characterizations of level sets in Theorem 5.1, when

    a=infμinv(Λ)gdμ or a=supμinv(Λ)gdμ,a=\displaystyle\inf_{\mu\in\mathcal{M}_{inv}(\Lambda)}\int g{\rm d}\mu\text{~{}~{}~{}or~{}~{}~{}}a=\displaystyle\sup_{\mu\in\mathcal{M}_{inv}(\Lambda)}\int g{\rm d}\mu,

    by [25], one also has the variational principle that

    htop(Rg(a))=sup{hμ(X):gdμ=a,μinv(Λ)}.h_{\rm top}(R_{g}(a))=\sup\left\{h_{\mu}(X)\colon\int g{\rm d}\mu=a,\mu\in\mathcal{M}_{inv}(\Lambda)\right\}.
  2. 2.

    In [63, Section 9], the author also constructed an increasing sequence of basic sets to study the multifractal spectra for Katok maps.

5.1 Entropy estimates for irregular sets and level sets

Given a compact metric space (K,d)(K,d), for a homeomorphism f:KKf\colon K\rightarrow K and a continuous roof function ρ:K(0,+)\rho\colon K\rightarrow(0,+\infty), we consider the suspension flow (Kρ,𝔉)(K_{\rho},\mathfrak{F}) where 𝔉=(ft)t\mathfrak{F}=(f_{t})_{t} as defined in Section 2.4. Analogous to the discrete case, for a continuous function g:Kρg\colon K_{\rho}\rightarrow\mathbb{R}, we define

g¯(x,s)=lim infT1T0Tg(ft(x,s))dtandg¯(x,s)=lim supT1T0Tg(ft(x,s))dt.\underline{g}(x,s)=\liminf_{T\to\infty}\frac{1}{T}\int_{0}^{T}g(f_{t}(x,s)){\rm d}t\qquad\textrm{and}\qquad\overline{g}(x,s)=\limsup_{T\to\infty}\frac{1}{T}\int_{0}^{T}g(f_{t}(x,s)){\rm d}t.

Define the irregular set

Igρ(𝔉):={(x,s)Kρ:g¯(x,s)<g¯(x,s)},I^{\rho}_{g}(\mathfrak{F}):=\big{\{}(x,s)\in K_{\rho}\colon\underline{g}(x,s)<\overline{g}(x,s)\big{\}},

and for aa\in\mathbb{R}, define the level set

Rgρ(𝔉,a)={(x,s)Kρ:g¯(x,s)=g¯(x,s)=a}.R^{\rho}_{g}(\mathfrak{F},a)=\big{\{}(x,s)\in K_{\rho}\colon\underline{g}(x,s)=\overline{g}(x,s)=a\big{\}}.

For a dynamical system (K,f)(K,f) 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 (Kρ,𝔉)(K_{\rho},\mathfrak{F}) of (K,f)(K,f).

Theorem 5.3 (Thompson [58][57]).

Let (K,d)(K,d) be a compact metric space, f:KKf\colon K\rightarrow K be a homeomorphism satisfying the specification property and ρ:K(0,+)\rho\colon K\to(0,+\infty) be a continuous function. Let (Kρ,𝔉)(K_{\rho},\mathfrak{F}) denote the suspension flow over (K,f)(K,f) with roof function ρ\rho and let g:Kρg\colon K_{\rho}\rightarrow\mathbb{R} be a continuous function. Then:

  1. 1.

    For any aa\in\mathbb{R},

    htop(𝔉,Rgρ(𝔉,a))=sup{hμ(𝔉):μinv(𝔉,Kρ)andgdμ=a}.h_{\rm top}(\mathfrak{F},R^{\rho}_{g}(\mathfrak{F},a))=\sup\left\{h_{\mu}(\mathfrak{F})\colon\mu\in\mathcal{M}_{inv}(\mathfrak{F},K_{\rho})~{}\textrm{and}~{}\int g{\rm d}\mu=a\right\}.
  2. 2.

    If infμinv(𝔉,Kρ)gdμ<supμinv(𝔉,Xρ)gdμ\displaystyle\inf_{\mu\in\mathcal{M}_{inv}(\mathfrak{F},K_{\rho})}\int g{\rm d}\mu<\sup_{\mu\in\mathcal{M}_{inv}(\mathfrak{F},X_{\rho})}\int g{\rm d}\mu then

    htop(𝔉,Igρ(𝔉))=htop(𝔉).h_{\rm top}(\mathfrak{F},I^{\rho}_{g}(\mathfrak{F}))=h_{\rm top}(\mathfrak{F}).
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 C1C^{1} vector fields, which complement previous results on the multifractal analysis of hyperbolic flows [9, 16, 46]. Recall that for an invariant compact set Λ\Lambda of X𝒳1(M)X\in\mathcal{X}^{1}(M) and for a continuous function g:Λg\colon\Lambda\rightarrow\mathbb{R}, the gg-irregular set is

Ig:={xΛ:limT1T0Tg(ϕt(x))dt does not exist}I_{g}:=\left\{x\in\Lambda\colon\lim_{T\rightarrow\infty}\frac{1}{T}\int_{0}^{T}g(\phi_{t}(x))\,{\rm d}t\text{ does not exist}\right\}

and, for each aa\in\mathbb{R}, the gg-level set is

Rg(a):={xΛ:limT1T0Tg(ϕt(x))dt=a}.R_{g}(a):=\left\{x\in\Lambda\colon\lim_{T\rightarrow\infty}\frac{1}{T}\int_{0}^{T}g(\phi_{t}(x)){\rm d}t=a\right\}.
Corollary 5.5.

Let Λ\Lambda be a horseshoe of X𝒳1(M)X\in\mathcal{X}^{1}(M) and g:Λg\colon\Lambda\rightarrow\mathbb{R} be a continuous function. The following properties hold:

  1. 1.

    For any aa\in\mathbb{R},

    htop(Rg(a))=sup{hμ(X):μinv(Λ)andgdμ=a}.h_{\rm top}(R_{g}(a))=\sup\left\{h_{\mu}(X)\colon\mu\in\mathcal{M}_{inv}(\Lambda)~{}\textrm{and}~{}\int g{\rm d}\mu=a\right\}.
  2. 2.

    If infμinv(Λ)gdμ<supμinv(Λ)gdμ\displaystyle\inf_{\mu\in\mathcal{M}_{inv}(\Lambda)}\int g{\rm d}\mu<\sup_{\mu\in\mathcal{M}_{inv}(\Lambda)}\int g{\rm d}\mu, then htop(Ig)=htop(Λ).h_{\rm top}(I_{g})=h_{\rm top}(\Lambda).

Proof.

By Theorem 4.3, there exists a suspension flow (Σρ,𝔉)(\Sigma_{\rho},\mathfrak{F}) over a transitive SFT (Σ,σ)(\Sigma,\sigma) such that (Λ,Φ)(\Lambda,\Phi) is semi-conjugate to (Σρ,𝔉)(\Sigma_{\rho},\mathfrak{F}), where ρ:Σ+\rho\colon\Sigma\rightarrow\mathbb{R}^{+} is a continuous function, Φ=(ϕt)t\Phi=(\phi_{t})_{t} is the flow generated by XX and 𝔉=(σt)t\mathfrak{F}=(\sigma_{t})_{t} is the suspension flow. That is to say, there exists a finite-to-one continuous map π:ΣρΛ\pi\colon\Sigma_{\rho}\rightarrow\Lambda satisfying that πσt=ϕtπ\pi\circ\sigma_{t}=\phi_{t}\circ\pi for any tt\in\mathbb{R}. In particular π\pi preserves entropy. As gC(Λ,)g\in C(\Lambda,\mathbb{R}), then g^=gπC(Σρ,)\hat{g}=g\circ\pi\in C(\Sigma_{\rho},\mathbb{R}). Moreover, it is easy to check that π(Rg^ρ(𝔉,a))=Rg(a)\pi(R_{\hat{g}}^{\rho}(\mathfrak{F},a))=R_{g}(a) for each aa\in\mathbb{R}, and π(Ig^ρ(𝔉))=Ig\pi(I_{\hat{g}}^{\rho}(\mathfrak{F}))=I_{g}. 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 C1C^{1} vector fields.

Proposition 5.6.

Let Λ\Lambda be a singular hyperbolic homoclinic class of X𝒳1(M)X\in\mathcal{X}^{1}(M) such that each pair of periodic orbits are homoclinically related and g:Λg\colon\Lambda\rightarrow\mathbb{R} be a continuous function. Assume that inv(Λ)\mathcal{M}\subseteq\mathcal{M}_{inv}(\Lambda) is a convex subset satisfying the horseshoe approximation property and

infμgdμ<supμgdμ.\displaystyle\inf_{\mu\in\mathcal{M}}\int g{\rm d}\mu<\sup_{\mu\in\mathcal{M}}\int g{\rm d}\mu.

Then the following properties are satisfied.

  1. 1.

    The topological entropy of the gg-irregular set IgI_{g} satisfies

    htop(Ig)h(X):=sup{hμ(X):μ}.h_{\rm top}(I_{g})\geq h_{\mathcal{M}}(X)\colon=\sup\left\{h_{\mu}(X)\colon\mu\in\mathcal{M}\right\}.
  2. 2.

    For any a(infμgdμ,supμgdμ)\displaystyle a\in\left(\inf_{\mu\in\mathcal{M}}\int g{\rm d}\mu,\sup_{\mu\in\mathcal{M}}\int g{\rm d}\mu\right), the topological entropy of the level set Rg(a)R_{g}(a) satisfies

    htop(Rg(a))h(a):=sup{hμ(X):μ and gdμ=a}.h_{\rm top}(R_{g}(a))\geq h_{\mathcal{M}}(a)\colon=\sup\left\{h_{\mu}(X)\colon\mu\in\mathcal{M}\text{~{}and ~{}}\int g{\rm d}\mu=a\right\}.
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 a¯=infμgdμ and a¯=supμgdμ\displaystyle\underline{a}=\inf_{\mu\in\mathcal{M}}\int g{\rm d}\mu\text{~{}and~{}}\overline{a}=\sup_{\mu\in\mathcal{M}}\int g{\rm d}\mu for simplicity. Fix a(a¯,a¯)a\in(\underline{a},\overline{a}). Since \mathcal{M} is convex, one has that {μ:gdμ=a}.\left\{\mu\in\mathcal{M}\colon\int g{\rm d}\mu=a\right\}\neq\emptyset. For each nn\in\mathbb{N}, take μn,νn\mu_{n},\nu_{n}\in\mathcal{M} so that

  • gdμn=a and hμn(X)>h(a)1n;\displaystyle\int g{\rm d}\mu_{n}=a\text{~{}and~{}}h_{\mu_{n}}(X)>h_{\mathcal{M}}(a)-\frac{1}{n};

  • hνn(X)>h(X)1n.h_{\nu_{n}}(X)>h_{\mathcal{M}}(X)-\frac{1}{n}.

We give the following claim first.

Claim 5.7.

There exists a horseshoe ΛnΛ\Lambda_{n}\subseteq\Lambda such that

  • (a)

    infμinv(Λn)gdμ<a¯+a2<a<a+a¯2<supμinv(Λn)gdμ\displaystyle\inf_{\mu\in\mathcal{M}_{inv}(\Lambda_{n})}\int g{\rm d}\mu<\frac{\underline{a}+a}{2}<a<\frac{a+\overline{a}}{2}<\sup_{\mu\in\mathcal{M}_{inv}(\Lambda_{n})}\int g{\rm d}\mu;

  • (b)

    there exist μn,νninv(Λn)\mu_{n}^{\prime},\nu_{n}^{\prime}\in\mathcal{M}_{inv}(\Lambda_{n}) such that
    (b.1)|gdμngdμn|<1n and hμn(X)>hμn(X)1n>h(a)2n,\displaystyle(b.1)~{}\left|\int g{\rm d}\mu_{n}-\int g{\rm d}\mu_{n}^{\prime}\right|<\frac{1}{n}\text{~{}and~{}}h_{\mu_{n}^{\prime}}(X)>h_{\mu_{n}}(X)-\frac{1}{n}>h_{\mathcal{M}}(a)-\frac{2}{n},
    (b.2)hνn(X)>hνn(X)1n>h(X)2n.\displaystyle(b.2)~{}h_{\nu_{n}^{\prime}}(X)>h_{\nu_{n}}(X)-\frac{1}{n}>h_{\mathcal{M}}(X)-\frac{2}{n}.

Proof of Claim 5.7.

The proof is by applying the fact that \mathcal{M} satisfies the horseshoe approximation property twice and by Lemma 4.5. We give a short explanation. First, by the definition of a¯\underline{a} and a¯\overline{a}, there exist η1,η2\eta_{1},\eta_{2}\in\mathcal{M} such that gdη1<a¯+a2\displaystyle\int g{\rm d}\eta_{1}<\frac{\underline{a}+a}{2} and gdη2>a¯+a2\displaystyle\int g{\rm d}\eta_{2}>\frac{\overline{a}+a}{2}. Since \mathcal{M} satisfies the horseshoe approximation property, there exist two horseshoes Δ1,Δ2\Delta_{1},\Delta_{2} such that

infμinv(Δ1)gdμ<a¯+a2<a<a+a¯2<supμinv(Δ2)gdμ.\displaystyle\inf_{\mu\in\mathcal{M}_{inv}(\Delta_{1})}\int g{\rm d}\mu<\frac{\underline{a}+a}{2}<a<\frac{a+\overline{a}}{2}<\sup_{\mu\in\mathcal{M}_{inv}(\Delta_{2})}\int g{\rm d}\mu.

On the other hand, by the horseshoe approximation property again, there exist two horseshoes Δn1,Δn2\Delta_{n}^{1},\Delta_{n}^{2} and μninv(Δn1),νninv(Δn2)\mu_{n}^{\prime}\in\mathcal{M}_{inv}(\Delta^{1}_{n}),\nu_{n}^{\prime}\in\mathcal{M}_{inv}(\Delta^{2}_{n}) such that Items (b.1)(b.1) and (b.2)(b.2) are satisfied for μn\mu_{n}^{\prime} and νn\nu_{n}^{\prime}. Then by Lemma 4.5, one can take a larger horseshoe Λn\Lambda_{n} that contains all the horseshoes Δ1,Δ2,Δn1,Δn2\Delta_{1},\Delta_{2},\Delta^{1}_{n},\Delta^{2}_{n}, thus the above statements hold for Λn\Lambda_{n}. ∎

In the following, we take the horseshoe ΛnΛ\Lambda_{n}\subseteq\Lambda and measures μn,νninv(Λn)\mu_{n}^{\prime},\nu_{n}^{\prime}\in\mathcal{M}_{inv}(\Lambda_{n}) from Claim 5.7.

Entropy of the gg-irregular set IgI_{g}.

By item 2 of Corollary 5.5, item (a) above implies that

htop(IgΛn)=htop(Λn).h_{\rm top}(I_{g}\cap\Lambda_{n})=h_{\rm top}(\Lambda_{n}).

Thus by item (b.2) above, one has

htop(Ig)htop(IgΛn)=htop(Λn)hνn(X)>h(X)2n.h_{\rm top}(I_{g})\geq h_{\rm top}(I_{g}\cap\Lambda_{n})=h_{\rm top}(\Lambda_{n})\geq h_{\nu_{n}^{\prime}}(X)>h_{\mathcal{M}}(X)-\frac{2}{n}.

Let n+n\to+\infty we conclude that Item 1 of Proposition 5.6 holds.

Entropy of the level set Rg(a)R_{g}(a).

We claim there exists an invariant measure ωninv(Λn)\omega_{n}\in\mathcal{M}_{inv}(\Lambda_{n}) such that gdωn=a\displaystyle\int g{\rm d}\omega_{n}=a and hωn(X)h_{\omega_{n}}(X) tends to h(a)h_{\mathcal{M}}(a) as nn\to\infty. By item (b.1) above and the fact gdμn=a\displaystyle\int g{\rm d}\mu_{n}=a, one has that

a1n<gdμn<a+1n.\displaystyle a-\frac{1}{n}<\int g{\rm d}\mu_{n}^{\prime}<a+\frac{1}{n}.

Without loss of generality, we may assume that agdμn<a+1n\displaystyle a\leq\int g{\rm d}\mu_{n}^{\prime}<a+\frac{1}{n} (the other case is analogous). By item (a) above, there exists θninv(Λn)\theta_{n}\in\mathcal{M}_{inv}(\Lambda_{n}) such that

a¯gdθn<a¯+a2(<a).\displaystyle\underline{a}\leq\int g{\rm d}\theta_{n}<\frac{\underline{a}+a}{2}\quad\big{(}<a\big{)}.

Thus one has

0<aa¯2<gdμngdθn<aa¯+1n.0<\frac{a-\underline{a}}{2}<\int g{\rm d}\mu_{n}^{\prime}-\int g{\rm d}\theta_{n}<a-\underline{a}+\frac{1}{n}.

Then, by the affinity of the integral and the entropy function, one has that the probability measure ωn=gdμnagdμngdθnθn+agdθngdμngdθnμn\displaystyle\omega_{n}=\frac{\int g{\rm d}\mu_{n}^{\prime}-a}{\int g{\rm d}\mu_{n}^{\prime}-\int g{\rm d}\theta_{n}}\theta_{n}+\frac{a-\int g{\rm d}\theta_{n}}{\int g{\rm d}\mu_{n}^{\prime}-\int g{\rm d}\theta_{n}}\mu_{n}^{\prime} satisfies gdωn=a,\displaystyle\int g{\rm d}\omega_{n}=a, and

hωn(X)\displaystyle h_{\omega_{n}}(X) agdθngdμngdθnhμn(X)\displaystyle\geq\frac{a-\int g{\rm d}\theta_{n}}{\int g{\rm d}\mu_{n}^{\prime}-\int g{\rm d}\theta_{n}}h_{\mu_{n}^{\prime}}(X)
=(1+agdμngdμngdθn)hμn(X)\displaystyle=\left(1+\frac{a-\int g{\rm d}\mu^{\prime}_{n}}{\int g{\rm d}\mu_{n}^{\prime}-\int g{\rm d}\theta_{n}}\right)h_{\mu_{n}^{\prime}}(X)
(11n2aa¯)(h(a)2n).\displaystyle\geq\left(1-\frac{1}{n}\frac{2}{a-\underline{a}}\right)\left(h_{\mathcal{M}}(a)-\frac{2}{n}\right).

By Item 1 in Corollary 5.5, one concludes that

htop(Rg(a))htop(Rg(a)Λn)hωn(X)(11n2aa¯)(h(a)2n).h_{\rm top}(R_{g}(a))\geq h_{\rm top}(R_{g}(a)\cap\Lambda_{n})\geq h_{\omega_{n}}(X)\geq\left(1-\frac{1}{n}\frac{2}{a-\underline{a}}\right)\left(h_{\mathcal{M}}(a)-\frac{2}{n}\right).

As the right hand-side above tends to h(a)h_{\mathcal{M}}(a) as nn tends to infinity, this proves item 2 of Proposition 5.6. ∎

5.2 Proofs of Theorems AB

We first prove Theorem 5.1.

Proof of Theorem 5.1.

Assume X𝒳1(M)X\in\mathscr{X}^{1}(M) and Λ\Lambda is a non-trivial singular hyperbolic homoclinic class of XX such that each pair of periodic orbits contained in Λ\Lambda are homoclinically related and inv(Λ)=1(Λ)¯\mathcal{M}_{inv}(\Lambda)=\overline{\mathcal{M}_{1}(\Lambda)}. Proposition 4.13 implies inv(Λ)\mathcal{M}_{inv}(\Lambda) has the horseshoe approximation property.

Let gC(Λ,)g\in C(\Lambda,\mathbb{R}). If there exists a0a_{0}\in\mathbb{R} such that gdμ=a0\displaystyle\int g{\rm d}\mu=a_{0} for every μinv(Λ)\mu\in\mathcal{M}_{inv}(\Lambda), then Λ=Rg(a0)\Lambda=R_{g}(a_{0}) and hence Ig=I_{g}=\emptyset.

Now assume there are ω1,ω2inv(Λ)\omega_{1},\omega_{2}\in\mathcal{M}_{inv}(\Lambda) such that gdω1gdω2\displaystyle\int g{\rm d}\omega_{1}\neq\int g{\rm d}\omega_{2}. By Lemma 4.9, there exist ν1,ν2per(Λ)\nu_{1},\nu_{2}\in\mathcal{M}_{per}(\Lambda) such that

gdν1gdν2.\displaystyle\int g{\rm d}\nu_{1}\neq\int g{\rm d}\nu_{2}.

Since each pair of periodic orbits contained in Λ\Lambda are homoclinic related, the stable manifold of any periodic orbit is dense in Λ\Lambda. Then [16, Theorem A] implies that IgI_{g} is residual in Λ\Lambda (see alternatively the proof of [16, Corollary VI]).

Entropy of IgI_{g}.

By Proposition 4.12, one has that 1(Λ)\mathcal{M}_{1}(\Lambda) satisfies the horseshoe approximation property, thus by Item 1 of Proposition 5.6, one has

htop(Ig)h1(Λ)(X):=sup{hμ(X):μ1(Λ)}.h_{\rm top}(I_{g})\geq h_{\mathcal{M}_{1}(\Lambda)}(X)\colon=\sup\{h_{\mu}(X)\colon\mu\in\mathcal{M}_{1}(\Lambda)\}.

Recall that 1(Λ)={μinv(Λ):μ(Sing(Λ))=0}\mathcal{M}_{1}(\Lambda)=\Big{\{}\mu\in\mathcal{M}_{inv}(\Lambda)\colon\mu(\operatorname{Sing}(\Lambda))=0\Big{\}} and hμ(X)=0h_{\mu}(X)=0 if μ(Sing(Λ))=1\mu(\operatorname{Sing}(\Lambda))=1. Thus by the variational principle, one has

h1(Λ)(X)=sup{hμ(X):μinv(Λ)}=htop(Λ).h_{\mathcal{M}_{1}(\Lambda)}(X)=\sup\{h_{\mu}(X)\colon\mu\in\mathcal{M}_{inv}(\Lambda)\}=h_{\rm top}(\Lambda).

As a consequence,

htop(Ig)=htop(Λ).h_{\rm top}(I_{g})=h_{\rm top}(\Lambda).

Denseness and entropy of Rg(a)R_{g}(a).

Take aa\in\mathbb{R} such that

infμ1(Λ)gdμ<a<supμ1(Λ)gdμ.\displaystyle\inf_{\mu\in\mathcal{M}_{1}(\Lambda)}\int g{\rm d}\mu<a<\sup_{\mu\in\mathcal{M}_{1}(\Lambda)}\int g{\rm d}\mu.

As above, this ensures that there exist μ1,μ2per(Λ)\mu_{1},\mu_{2}\in\mathcal{M}_{per}(\Lambda) so that gdμ1<a<gdμ2\displaystyle\int g{\rm d}\mu_{1}<a<\int g{\rm d}\mu_{2}.

To obtain the denseness of Rg(a)R_{g}(a) in Λ\Lambda, one first constructs a nested sequence of horseshoes {Λn}n\{\Lambda_{n}\}_{n\in\mathbb{N}} approximating Λ\Lambda. Note that all periodic orbits contained in Λ\Lambda are hyperbolic. Since Λ\Lambda is non-trivial, there are countably infinitely many periodic orbits contained in Λ\Lambda and one lists them as {γn}n\{\gamma_{n}\}_{n\in\mathbb{N}}. Moreover, each pair of periodic orbits in Λ\Lambda are homoclinically related. One constructs {Λn}n\{\Lambda_{n}\}_{n\in\mathbb{N}} inductively as follows:

  • Let Λ0\Lambda_{0} be a horseshoe that contains γ0\gamma_{0} and γ1\gamma_{1}. Such a horseshoe exists because γ0\gamma_{0} and γ1\gamma_{1} are homoclinically related.

  • For n1n\geq 1, let Λn\Lambda_{n} be a horseshoe that contains Λn1\Lambda_{n-1} and γn\gamma_{n}. Such a horseshoe exists because γn\gamma_{n} is homoclinically related with Λn1\Lambda_{n-1} in the sense that γn\gamma_{n} is homoclinically with all periodic orbits contained in Λn1\Lambda_{n-1}. Note that if γn\gamma_{n} is contained in Λn1\Lambda_{n-1}, then Λn=Λn1\Lambda_{n}=\Lambda_{n-1}.

By construction, one has that ΛnΛn+1\Lambda_{n}\subset\Lambda_{n+1} for all nn\in\mathbb{N}. Moreover, by denseness of the periodic orbits in Λ\Lambda one has that Λn\Lambda_{n} tends to Λ\Lambda (in the Hausdorff distance) as nn\rightarrow\infty. Recall that one has assumed μ1,μ2\mu_{1},\mu_{2} to be two periodic measures, thus there exists n0n_{0}\in\mathbb{N} such that μ1,μ2inv(Λn0)\mu_{1},\mu_{2}\in\mathcal{M}_{inv}(\Lambda_{n_{0}}). As a consequence, μ1,μ2inv(Λn)\mu_{1},\mu_{2}\in\mathcal{M}_{inv}(\Lambda_{n}) for all nn0n\geq n_{0}. By Item 2 of Theorem 4.4, for each nn0n\geq n_{0}, there exists μ1n,μ2ninv(Λn)\mu_{1}^{n},\mu_{2}^{n}\in\mathcal{M}_{inv}(\Lambda_{n}) such that

gdμ1n<a<gdμ2nandSupp(μ1n)=Supp(μ2n)=Λn.\displaystyle\int g{\rm d}\mu_{1}^{n}<a<\int g{\rm d}\mu_{2}^{n}\qquad{\rm and}\qquad\operatorname{\,Supp}(\mu_{1}^{n})=\operatorname{\,Supp}(\mu_{2}^{n})=\Lambda_{n}.

Take a suitable θn(0,1)\theta_{n}\in(0,1) for each nn0n\geq n_{0}, such that the linear combination

νn=θnμ1n+(1θn)μ2nsatisfiesgdνn=a.\nu_{n}=\theta_{n}\mu_{1}^{n}+(1-\theta_{n})\mu_{2}^{n}\qquad{\rm satisfies}\qquad\displaystyle\int g{\rm d}\nu_{n}=a.

Note that Supp(νn)=Λn\operatorname{\,Supp}(\nu_{n})=\Lambda_{n}. By Item 1 of Theorem 4.4, the set GνnG_{\nu_{n}} of νn\nu_{n}-generic points is non-empty. Take xnGνnΛnx_{n}\in G_{\nu_{n}}\cap\Lambda_{n}, then Supp(νn)ω(xn,Φ)\operatorname{\,Supp}(\nu_{n})\subset\omega(x_{n},\Phi) where ω(xn,Φ)\omega(x_{n},\Phi) is the positive limit set of xnx_{n}. This implies that ω(xn,Φ)=Λn\omega(x_{n},\Phi)=\Lambda_{n}. Note that Orb(xn)Gνn\operatorname{Orb}(x_{n})\subset G_{\nu_{n}}, hence GνnG_{\nu_{n}} is dense in Λn\Lambda_{n}. As a consequence, one has nn0Gνn\bigcup_{n\geq n_{0}}G_{\nu_{n}} is dense in Λ\Lambda. By the fact that nn0GνnRg(a)\bigcup_{n\geq n_{0}}G_{\nu_{n}}\subset R_{g}(a), one has Rg(a)R_{g}(a) is dense in Λ\Lambda.

The estimation of the entropy htop(Rg(a))h_{\rm top}(R_{g}(a)) follows similar arguments as for htop(Ig)h_{\rm top}(I_{g}). Since inv(Λ)\mathcal{M}_{inv}(\Lambda) satisfies the horseshoe approximation property, using Item 2 of Proposition 5.6, one has

htop(Rg(a))hinv(Λ)(a):=sup{hμ(X):μinv(Λ) and gdμ=a}.\displaystyle h_{\rm top}(R_{g}(a))\geq h_{\mathcal{M}_{inv}(\Lambda)}(a)\colon=\sup\left\{h_{\mu}(X)\colon\mu\in\mathcal{M}_{inv}(\Lambda)\text{~{}and ~{}}\int g{\rm d}\mu=a\right\}.

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 xΛx\in\Lambda, denote by V(x)V(x) and V(x,ϕ1)V(x,\phi_{1}) the limit sets of empirical measures of xx under the action of the flow (ϕt)t(\phi_{t})_{t} and its time-one map ϕ1\phi_{1}, respectively. In other words,

V(x)={μinv(Λ):ti+ s.t. μ=limi1ti0tiδϕs(x)ds}V(x)=\left\{\mu\in\mathcal{M}_{inv}(\Lambda)\colon\exists t_{i}\rightarrow+\infty\text{~{}s.t.~{}}\mu=\lim_{i\rightarrow\infty}\frac{1}{t_{i}}\int_{0}^{t_{i}}\delta_{\phi_{s}(x)}\,{\rm d}s\right\}

and

V(x,ϕ1)={μinv(Λ,ϕ1):ni+ s.t. μ=limi1nik=0niδϕk(x)}.V(x,\phi_{1})=\left\{\mu\in\mathcal{M}_{inv}(\Lambda,\phi_{1})\colon\exists n_{i}\rightarrow+\infty\text{~{}s.t.~{}}\mu=\lim_{i\rightarrow\infty}\frac{1}{n_{i}}\sum_{k=0}^{n_{i}}\delta_{\phi_{k}(x)}\,\right\}.

To simplify notations, let b=sup{hμ(X):μinv(Λ) and gdμ=a}\displaystyle b=\sup\Big{\{}h_{\mu}(X)\colon\mu\in\mathcal{M}_{inv}(\Lambda)\text{~{}and ~{}}\int g{\rm d}\mu=a\Big{\}}. We need the following:

Claim.

For any xRg(a)x\in R_{g}(a) and any νV(x,ϕ1)\nu\in V(x,\phi_{1}), one has hν(ϕ1)bh_{\nu}(\phi_{1})\leq b.

Proof of the claim.

Take xRg(a)x\in R_{g}(a). Each μV(x)\mu\in V(x) satisfies that gdμ=a\int g{\rm d}\mu=a and thus hμ(X)bh_{\mu}(X)\leq b. On the other hand, for any νV(x,ϕ1)\nu\in V(x,\phi_{1}), the measure μ=01(ϕs)νds\mu=\int_{0}^{1}(\phi_{s})_{*}\nu\,{\rm d}s is invariant by the flow and belongs to V(x)V(x). As a consequence, one has hν(ϕ1)=hμ(ϕ1)=hμ(X)bh_{\nu}(\phi_{1})=h_{\mu}(\phi_{1})=h_{\mu}(X)\leq b since the metric entropy is affine on the space of invariant probability measures. ∎

The above claim implies that

Rg(a)QR(b):={xΛ:νV(x,ϕ1) with hν(ϕ1)b}.R_{g}(a)\subset QR(b)\colon=\{x\in\Lambda\colon\exists\nu\in V(x,\phi_{1})\text{~{}\rm with~{}}h_{\nu}(\phi_{1})\leq b\}.

By [14, Theorem 2], one concludes that

htop(Rg(a))htop(QR(b),ϕ1)b.h_{\rm top}(R_{g}(a))\leq h_{\rm top}(QR(b),\phi_{1})\leq b.

Finally, it remains to show that the set

C^={gC(Λ,):gdμ1gdμ2 for some μ1,μ21(Λ)}\hat{C}=\left\{g\in C(\Lambda,\mathbb{R})\colon\int g{\rm d}\mu_{1}\neq\int g{\rm d}\mu_{2}\text{~{}for some~{}}\mu_{1},\mu_{2}\in\mathcal{M}_{1}(\Lambda)\right\}

is open and dense in C(Λ,)C(\Lambda,\mathbb{R}). The openness is by continuity of the integrals in the weak topology. To prove denseness, take two different periodic orbits Orb(p),Orb(q)\operatorname{Orb}(p),\operatorname{Orb}(q) in Λ\Lambda, take g^C(Λ,)\hat{g}\in C(\Lambda,\mathbb{R}) such that

g^|Orb(p)=0,g^|Orb(q)=1 and 0g^(x)1xΛ.\hat{g}|_{\operatorname{Orb}(p)}=0,~{}\hat{g}|_{\operatorname{Orb}(q)}=1~{}\text{~{}and~{}}0\leq\hat{g}(x)\leq 1~{}\forall x\in\Lambda.

Such g^\hat{g} exists since Orb(p)\operatorname{Orb}(p) and Orb(q)\operatorname{Orb}(q) are two distinct periodic orbits. Let ν1,ν2\nu_{1},\nu_{2} be the two periodic measures associated to Orb(p),Orb(q)\operatorname{Orb}(p),\operatorname{Orb}(q) respectively. Note that ν1,ν21(Λ)\nu_{1},\nu_{2}\in\mathcal{M}_{1}(\Lambda). For any gC(Λ,)g\in C(\Lambda,\mathbb{R}),

  • if gdν1gdν2\displaystyle\int g{\rm d}\nu_{1}\neq\int g{\rm d}\nu_{2}, then gC^g\in\hat{C};

  • if otherwise gdν1=gdν2\displaystyle\int g{\rm d}\nu_{1}=\int g{\rm d}\nu_{2}, let gn=g+1ng^g_{n}=g+\frac{1}{n}\hat{g}, then gnC^g_{n}\in\hat{C} and g=limngng=\lim\limits_{n\rightarrow\infty}g_{n}.

This shows C^\hat{C} is dense in C(Λ,)C(\Lambda,\mathbb{R}) and completes the proof of Theorem 5.1. ∎

Now we are ready to prove Theorem AB.

Proofs of Theorem AB.

By Item 1 of Corollary 4.14, there exists a residual subset r𝒳r(M3)\mathcal{R}^{r}\subset\mathcal{X}^{r}(M^{3}) where r2r\in\mathbb{N}_{\geq 2} such that for any XrX\in\mathcal{R}^{r}, if Λ\Lambda is a geometric Lorenz attractor for XX, then inv(Λ)=1(Λ)¯\mathcal{M}_{inv}(\Lambda)=\overline{\mathcal{M}_{1}(\Lambda)}. Moreover, by Proposition 2.6, every pair of periodic orbits are homoclinically related.

By Item 2 of Corollary 4.14, there exists a residual subset 𝒳1(M)\mathcal{R}\subset\mathcal{X}^{1}(M) such that for any XX\in\mathcal{R}, if Λ\Lambda is a singular hyperbolic attractor for XX, then inv(Λ)=1(Λ)¯\mathcal{M}_{inv}(\Lambda)=\overline{\mathcal{M}_{1}(\Lambda)} and every pair of periodic orbits are homoclinically related by [20, Theorem B].

Then Theorem AB are direct consequences of Theorem 5.1. ∎

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-s0s_{0} maps). Due to the presence of singularities, we still have to consider the following special subset

1(Λ)={μinv(Λ):μ(Sing(Λ))=0}\mathcal{M}_{1}(\Lambda)=\Big{\{}\mu\in\mathcal{M}_{inv}(\Lambda)\colon\mu(\operatorname{Sing}(\Lambda))=0\Big{\}}

of inv(Λ)\mathcal{M}_{inv}(\Lambda). Recall that (Λ)\mathcal{M}(\Lambda) is the space of all probability measures supported on Λ\Lambda. We prove the following theorem in this section.

Theorem 6.1.

(Level-2 large deviations) Let X𝒳1(M)X\in\mathscr{X}^{1}(M) and Λ\Lambda be a singular hyperbolic homoclinic class such that each pair of periodic orbits in Λ\Lambda are homoclinically related and 1(Λ)¯=inv(Λ)\overline{\mathcal{M}_{1}(\Lambda)}=\mathcal{M}_{inv}(\Lambda). Assume μψ\mu_{\psi} is a weak Gibbs measure with respect to a Hölder continuous potential ψ:Λ\psi\colon\Lambda\to\mathbb{R} with ΛH\Lambda_{H} being the μψ\mu_{\psi}-full measure set such that (2.1) satisfies. Then one has:

  1. 1.

    (upper bound) There exists c0c_{\infty}\leq 0 so that

    lim supt\displaystyle\limsup_{t\to\infty} 1tlogμψ({xΛ:t(x)𝒦})max{infμ𝒦ψ(μ),c}\displaystyle\frac{1}{t}\log\mu_{\psi}\big{(}\{x\in\Lambda\colon\mathcal{E}_{t}(x)\in\mathcal{K}\}\big{)}\leq\max\Big{\{}-\inf_{\mu\in\mathcal{K}}\mathfrak{I}_{\psi}(\mu)\;,\;c_{\infty}\Big{\}}

    for any closed subset 𝒦(Λ)\mathcal{K}\subset\mathcal{M}(\Lambda).

  2. 2.

    (lower bound) If 𝒪(Λ)\mathcal{O}\subset\mathcal{M}(\Lambda) is an open set and ν𝒪\nu\in\mathcal{O} is ergodic satisfying ν(ΛH)=1\nu(\Lambda_{H})=1, then

    lim inft+1tlogμψ({xΛ:t(x)𝒪})Ptop(Λ,ψ)+hν(X)+ψdν.\displaystyle\liminf_{t\to+\infty}\frac{1}{t}\log\mu_{\psi}\Big{(}\Big{\{}x\in\Lambda\colon\mathcal{E}_{t}(x)\in\mathcal{O}\Big{\}}\Big{)}\geq\displaystyle-P_{\rm top}(\Lambda,\psi)+h_{\nu}(X)+\int{\psi}\,{\rm d}\nu.
  3. 3.

    (lower bound for Gibbs measure) If μψ\mu_{\psi} is a Gibbs measure with respect to ψ\psi, then

    lim inft+1tlogμψ({xΛ:t(x)𝒪})infμ𝒪ψ(μ)\displaystyle\liminf_{t\to+\infty}\frac{1}{t}\log\mu_{\psi}\Big{(}\Big{\{}x\in\Lambda\colon\mathcal{E}_{t}(x)\in\mathcal{O}\Big{\}}\Big{)}\geq-\inf_{\mu\in\mathcal{O}}\mathfrak{I}_{\psi}(\mu)

    for any open subset 𝒪(Λ)\mathcal{O}\subset\mathcal{M}(\Lambda).

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 r𝒳r(M3)\mathcal{R}^{r}\subset\mathcal{X}^{r}(M^{3}) where r2r\in\mathbb{N}_{\geq 2} and the residual subset 𝒳1(M)\mathcal{R}\subset\mathcal{X}^{1}(M) in the proofs of Theorem AB. In both cases when Λ\Lambda is a geometric Lorenz attractor for XrX\in\mathcal{R}^{r} or Λ\Lambda is a singular hyperbolic attractor for XX\in\mathbb{R}, one has that inv(Λ)=1(Λ)¯\mathcal{M}_{inv}(\Lambda)=\overline{\mathcal{M}_{1}(\Lambda)} and Λ\Lambda 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 Λ\Lambda is a singular hyperbolic homoclinic class of X𝒳1(M)X\in\mathscr{X}^{1}(M) such that each pair of periodic orbits in Λ\Lambda are homoclinically related and we also assume that 1(Λ)¯=inv(Λ)\overline{\mathcal{M}_{1}(\Lambda)}=\mathcal{M}_{inv}(\Lambda). Assume μψinv(Λ)\mu_{\psi}\in\mathcal{M}_{inv}(\Lambda) is a weak Gibbs measure with respect to a Hölder continuous potential ψ:Λ\psi\colon\Lambda\rightarrow\mathbb{R} and ΛHΛ\Lambda_{H}\subset\Lambda be the μψ\mu_{\psi}-full measure set satisfied for (2.1). To be precise, there exists ε0>0\varepsilon_{0}>0 such that for any xΛH,t>0x\in\Lambda_{H},t>0 and ε(0,ε0)\varepsilon\in(0,\varepsilon_{0}), there exist constants Ct(x,ε)>0C_{t}(x,\varepsilon)>0 satisfying:

1Ct(x,ε)μψ(B(y,t,ε,Φ))etPtop(X,ψ)+0tψ(ϕs(x))dsCt(x,ε).\frac{1}{C_{t}(x,\varepsilon)}\leq\frac{\mu_{\psi}\Big{(}B\big{(}y,t,\varepsilon,\Phi\big{)}\Big{)}}{e^{-t\,P_{\rm top}(X,\psi)+\int_{0}^{t}\psi(\phi_{s}(x))\,{\rm d}s}}\leq C_{t}(x,\varepsilon). (6.1)

for any dynamic Bowen ball B(y,t,ε,Φ)B(x,t,ε0,Φ)B(y,t,\varepsilon,\Phi)\subset B(x,t,\varepsilon_{0},\Phi).

6.1 Lower bound

We give the lower bounds of large deviations in this section, that is to prove item 23 of Theorem 6.1. The following instrumental result proves item 2.

Proposition 6.2.

Let Λ,ψ,μψ\Lambda,\psi,\mu_{\psi} and ΛH\Lambda_{H} be as in the assumption above. If 𝒪(Λ)\mathcal{O}\subset\mathcal{M}(\Lambda) is an open set, ν𝒪\nu\in\mathcal{O} is ergodic and ν(ΛH)=1\nu(\Lambda_{H})=1 then

lim inft+1tlogμψ({xΛ:t(x)𝒪})Ptop(Λ,ψ)+hν(X)+ψdν.\displaystyle\liminf_{t\to+\infty}\frac{1}{t}\log\mu_{\psi}\Big{(}\Big{\{}x\in\Lambda\colon\mathcal{E}_{t}(x)\in\mathcal{O}\Big{\}}\Big{)}\geq\displaystyle-P_{\rm top}(\Lambda,\psi)+h_{\nu}(X)+\int{\psi}\,{\rm d}\nu.
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 {φi}i=1\big{\{}\varphi_{i}\big{\}}_{i=1}^{\infty} of C(Λ,)C(\Lambda,\mathbb{R}) where φi\varphi_{i} is not the zero function for every i1i\geq 1 and, for any μ,ν(Λ)\mu,\nu\in\mathcal{M}(\Lambda),

d(μ,ν)=i=1|φidμφidν|2iφi.\displaystyle d^{*}(\mu,\nu)=\sum_{i=1}^{\infty}\frac{|\int\varphi_{i}{\rm d}\mu-\int\varphi_{i}{\rm d}\nu|}{2^{i}\|\varphi_{i}\|}.

In consequence:

  1. (i)

    d(βμ,βν)=βd(μ,ν)\displaystyle d^{*}(\beta\mu,\beta\nu)=\beta\,\displaystyle d^{*}(\mu,\nu) for every μ,ν(Λ)\mu,\nu\in\mathcal{M}(\Lambda) and β>0\beta>0,

  2. (ii)

    d(μ1+μ2,ν1+ν2)d(μ1,ν1)+d(μ2,ν2)\displaystyle d^{*}(\mu_{1}+\mu_{2},\nu_{1}+\nu_{2})\leq\displaystyle d^{*}(\mu_{1},\nu_{1})+\displaystyle d^{*}(\mu_{2},\nu_{2}) for every μ1,μ2,ν1,ν2(Λ)\mu_{1},\mu_{2},\nu_{1},\nu_{2}\in\mathcal{M}(\Lambda).

Since 𝒪\mathcal{O} is open in the weak-topology and ν𝒪\nu\in\mathcal{O}, by the definition of the weak topology one may choose δ>0\delta>0 and finitely many functions φ1,,φi0C(Λ,)\varphi_{1},\dots,\varphi_{i_{0}}\in C(\Lambda,\mathbb{R}) as above so that the open neighborhood

𝒪3δ:={η(Λ):|φi𝑑νφi𝑑η|<3δ, 1ii0}\mathcal{O}^{3\delta}:=\Big{\{}\eta\in\mathcal{M}(\Lambda)\colon\Big{|}\int\varphi_{i}\,d\nu-\int\varphi_{i}\,d\eta\Big{|}<3\delta,\;\forall\,1\leq i\leq i_{0}\Big{\}}

is contained in 𝒪\mathcal{O} and

|ψ𝑑νψ𝑑η|<4δfor anyη𝒪3δ.\displaystyle\left|\int\psi\,d\nu-\int\psi\,d\eta\right|<4\delta\quad\text{for any}\quad\eta\in\mathcal{O}^{3\delta}.

As μψ\mu_{\psi} is a weak Gibbs measure, recall that for any xΛHx\in\Lambda_{H}, ε>0\varepsilon>0 and t>0t>0 there exist constants Ct(x,ε)>0C_{t}(x,\varepsilon)>0 satisfying  (6.1). Let ε>0\varepsilon>0 be small and fixed (to be chosen below). Firstly, by [48], the set of real numbers s0>0s_{0}>0 so that ν\nu is ergodic for the time-s0s_{0} map f=ϕs0f=\phi_{s_{0}} is Baire generic. One chooses such an s0>0s_{0}>0 small enough such that

supxΛd(1s00s0δϕs(x)𝑑s,δx)<δ\sup_{x\in\Lambda}d^{*}\Big{(}\frac{1}{s_{0}}\int_{0}^{s_{0}}\delta_{\phi_{s}(x)}\,ds,\delta_{x}\Big{)}<\delta

and let f=ϕs0f=\phi_{s_{0}} denote the time-s0s_{0} map. Then Proposition 2.1 in [47] applied to the open neighborhood 𝒪δ\mathcal{O}^{\delta} of ν\nu ensures that there is N1N\geq 1 and for every nNn\geq N 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 Xn,FBδX_{n,F}^{B^{\delta}} at equation (2.25) in [47] by Xn,FBδAnX_{n,F}^{B^{\delta}}\cap A_{n} for some family of sets (An)n1(A_{n})_{n\geq 1} such that ν(An)\nu(A_{n}) tends to 1 as nn~{}\to\infty. This is because of the fact ν(ΛH)=1\nu(\Lambda_{H})=1.

D{xΛH:nf(x)𝒪δ&Cs0n(x,ε/2)eδs0n}wherenf(x)=1ni=0n1δfi(x)D\subset\left\{x\in\Lambda_{H}\colon\mathcal{E}^{f}_{n}(x)\in\mathcal{O}^{\delta}\;\&\;C_{s_{0}n}(x,\varepsilon/2)\leq e^{\delta s_{0}n}\right\}~{}~{}~{}\text{where}~{}~{}\mathcal{E}^{f}_{n}(x)=\frac{1}{n}\sum_{i=0}^{n-1}\delta_{f^{i}(x)} (6.2)

such that DD is (n,ε)(n,\varepsilon)-separated and has cardinality larger than or equal to en(hν(f)δ)e^{n(h_{\nu}(f)-\delta)}. Using that f=ϕs0f=\phi_{s_{0}} it is clear that the set DD is (s0n,ε)(s_{0}n,\varepsilon)-separated with respect to the flow Φ=(ϕt)t\Phi=(\phi_{t})_{t}. Taking t=s0nt=s_{0}n, properties (i) and (ii) above imply

d(t(x),nf(x))\displaystyle d^{*}(\mathcal{E}_{t}(x),\mathcal{E}^{f}_{n}(x)) =1nd(j=0n11s00s0δϕs(fj(x))𝑑s,j=0n1δfj(x))\displaystyle=\frac{1}{n}d^{*}\Big{(}\sum_{j=0}^{n-1}\frac{1}{s_{0}}\int_{0}^{s_{0}}\delta_{\phi_{s}(f^{j}(x))}\,ds,\sum_{j=0}^{n-1}\delta_{f^{j}(x)}\Big{)}
1nj=0n1d(1s00s0δϕs(fj(x))𝑑s,δfj(x))<δ.\displaystyle\leq\frac{1}{n}\sum_{j=0}^{n-1}d^{*}\Big{(}\frac{1}{s_{0}}\int_{0}^{s_{0}}\delta_{\phi_{s}(f^{j}(x))}\,ds,\delta_{f^{j}(x)}\Big{)}<\delta.

Analogously, if xΛx\in\Lambda and yB(x,t,ε,Φ)y\in B(x,t,\varepsilon,\Phi) then

d(t(y),t(x))\displaystyle d^{*}(\mathcal{E}_{t}(y),\mathcal{E}_{t}(x)) =1nd(j=0n11s00s0δϕs(fj(y))𝑑s,j=0n11s00s0δϕs(fj(x))𝑑s)\displaystyle=\frac{1}{n}d^{*}\Big{(}\sum_{j=0}^{n-1}\frac{1}{s_{0}}\int_{0}^{s_{0}}\delta_{\phi_{s}(f^{j}(y))}\,ds,\sum_{j=0}^{n-1}\frac{1}{s_{0}}\int_{0}^{s_{0}}\delta_{\phi_{s}(f^{j}(x))}\,ds\Big{)}
1nj=0n1d(1s00s0δϕs(fj(x))𝑑s,1s00s0δϕs(fj(y))𝑑s)<δ.\displaystyle\leq\frac{1}{n}\sum_{j=0}^{n-1}d^{*}\Big{(}\frac{1}{s_{0}}\int_{0}^{s_{0}}\delta_{\phi_{s}(f^{j}(x))}\,ds,\frac{1}{s_{0}}\int_{0}^{s_{0}}\delta_{\phi_{s}(f^{j}(y))}\,ds\Big{)}<\delta.

Therefore one may reduce ε\varepsilon, if necessary, to guarantee that the summands in the right hand side above are arbitrarily small and consequently

xDB(x,t,ε,Φ){xΛ:t(x)𝒪3δ}.\bigcup_{x\in D}B(x,t,\varepsilon,\Phi)\subset\left\{x\in\Lambda\colon\mathcal{E}_{t}(x)\in\mathcal{O}^{3\delta}\right\}.

Therefore, using the definition of weak Gibbs measure in  (6.1), that hν(f)=s0hν(X)h_{\nu}(f)=s_{0}h_{\nu}(X) and that the dynamic balls B(x,t,ε/2,Φ)B(x,t,\varepsilon/2,\Phi) are pairwise disjoint, one concludes that

1tlogμψ({xΛ:t(x)𝒪})\displaystyle\frac{1}{t}\log\mu_{\psi}\Big{(}\Big{\{}x\in\Lambda\colon\mathcal{E}_{t}(x)\in\mathcal{O}\Big{\}}\Big{)} 1tlogxDμψ(B(x,t,ε/2,Φ))\displaystyle\geq\frac{1}{t}\log\sum_{x\in D}\mu_{\psi}\big{(}\,B(x,t,\varepsilon/2,\Phi)\,\big{)}
1tlogxD[Ct(x,ε)1etPtop(X,ψ)+0tψ(ϕs(x))ds]\displaystyle\geq\frac{1}{t}\log\sum_{x\in D}\big{[}C_{t}(x,\varepsilon)^{-1}{e^{-t\,P_{\rm top}(X,\psi)+\int_{0}^{t}\psi(\phi_{s}(x))\,{\rm d}s}}\big{]}

Taking nn\to\infty, which makes tt\to\infty as well, and by the choice of DD in  (6.2), we conclude that

lim inft+1tlogμψ({xΛ:t(x)𝒪})Ptop(X,ψ)+hν(X)+ψdν(1+1s0)δ.\liminf_{t\to+\infty}\frac{1}{t}\log\mu_{\psi}\Big{(}\Big{\{}x\in\Lambda\colon\mathcal{E}_{t}(x)\in\mathcal{O}\Big{\}}\Big{)}\geq\displaystyle-P_{\rm top}(X,\psi)+h_{\nu}(X)+\int{\psi}\,{\rm d}\nu-\big{(}1+\frac{1}{s_{0}}\big{)}\delta.

Since δ>0\delta>0 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 Λ\Lambda and ψ\psi be from the assumption of Proposition 6.2. Assume further that μψ\mu_{\psi} is a Gibbs measure with respect to ψ\psi. Given an open subset 𝒪(Λ)\mathcal{O}\subset\mathcal{M}(\Lambda) one has that

lim inft+1tlogμψ({xΛ:t(x)𝒪})infμ𝒪ψ(μ).\displaystyle\liminf_{t\to+\infty}\frac{1}{t}\log\mu_{\psi}\Big{(}\Big{\{}x\in\Lambda\colon\mathcal{E}_{t}(x)\in\mathcal{O}\Big{\}}\Big{)}\geq-\inf_{\mu\in\mathcal{O}}\mathfrak{I}_{\psi}(\mu).
Proof.

Take an open subset 𝒪(Λ)\mathcal{O}\subset\mathcal{M}(\Lambda), it is sufficient to show that, for each μ𝒪\mu\in\mathcal{O} one has

lim inft+1tlogμψ({xΛ:t(x)𝒪})ψ(μ).\liminf_{t\to+\infty}\frac{1}{t}\log\mu_{\psi}\Big{(}\Big{\{}x\in\Lambda\colon\mathcal{E}_{t}(x)\in\mathcal{O}\Big{\}}\Big{)}\geq-\mathfrak{I}_{\psi}(\mu). (6.3)

Note that since μψ\mu_{\psi} is a Gibbs measure, the set ΛH\Lambda_{H} can be chosen as Λ\Lambda, thus one has ν(ΛH)=ν(Λ)=1\nu(\Lambda_{H})=\nu(\Lambda)=1 for any ν(Λ)\nu\in\mathcal{M}(\Lambda).

If μinv(Λ)\mu\notin\mathcal{M}_{inv}(\Lambda), then ψ(μ)=+\mathfrak{I}_{\psi}(\mu)=+\infty and there is nothing to prove. Hence one just needs to consider where μinv(Λ)\mu\in\mathcal{M}_{inv}(\Lambda). Since by assumption 1(Λ)¯=inv(Λ)\overline{\mathcal{M}_{1}(\Lambda)}=\mathcal{M}_{inv}(\Lambda), Proposition 4.13 guarantees that inv(Λ){\mathcal{M}_{inv}(\Lambda)} admits the horseshoe approximation property. Thus, for any ε>0\varepsilon>0 there exists νεerg(Λ)\nu_{\varepsilon}\in\mathcal{M}_{erg}(\Lambda) so that Λ=Supp(νε)\Lambda^{\prime}=\operatorname{\,Supp}(\nu_{\varepsilon}) is a horseshoe, and the followings are satisfied:

d(νε,μ)<εLandhνε(X)>hμ(X)ε,d^{*}(\nu_{\varepsilon},\mu)<\frac{\varepsilon}{L}~{}~{}~{}\text{and}~{}~{}~{}h_{\nu_{\varepsilon}}(X)>h_{\mu}(X)-\varepsilon,

where L=ψL=\|\psi\| is the supremum norm of ψ\psi. In particular one has |ψ𝑑μψ𝑑νε|<ε\displaystyle\left|\int\psi\,d\mu-\int\psi\,d\nu_{\varepsilon}\right|<\varepsilon. The following estimation holds

ψ(νε)\displaystyle-\mathfrak{I}_{\psi}(\nu_{\varepsilon}) =hνε(X)+ψdνεPtop(X,ψ)\displaystyle=h_{\nu_{\varepsilon}}(X)+\int{\psi}\,{\rm d}\nu_{\varepsilon}-P_{\rm top}(X,\psi)
>hμ(X)+ψdμPtop(X,ψ)2ε\displaystyle>h_{\mu}(X)+\int{\psi}\,{\rm d}\mu-P_{\rm top}(X,\psi)-2\varepsilon
=ψ(μ)2ε.\displaystyle=-\mathfrak{I}_{\psi}(\mu)-2\varepsilon.

Shrink ε\varepsilon so that νε𝒪\nu_{\varepsilon}\in\mathcal{O}. Note that since νε\nu_{\varepsilon} is ergodic and ΛH=Λ\Lambda_{H}=\Lambda, one applies Proposition 6.2 to 𝒪\mathcal{O} and νε\nu_{\varepsilon} and obtains the following

lim inft+1tlogμψ({xΛ:t(x)𝒪})ψ(νε)>ψ(μ)2ε.\liminf_{t\to+\infty}\frac{1}{t}\log\mu_{\psi}\Big{(}\Big{\{}x\in\Lambda:\mathcal{E}_{t}(x)\in\mathcal{O}\Big{\}}\Big{)}\geq-\mathfrak{I}_{\psi}(\nu_{\varepsilon})>-\mathfrak{I}_{\psi}(\mu)-2\varepsilon.

Corollary 6.3 is concluded since ε\varepsilon can be taken arbitrarily small. ∎

Proposition 6.2 together with Corollary 6.3 prove item 23 of Theorem 6.1.

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 Λ\Lambda be an invariant compact set of a vector field X𝒳1(M)X\in\mathcal{X}^{1}(M) and let D(Λ)D\subset\mathcal{M}(\Lambda) be a non-empty set. If sD(t,ε)s_{D}(t,\varepsilon) denotes the maximal cardinality of (t,ε)(t,\varepsilon)-separated sets in {xM:t(x)D}\Big{\{}x\in M\colon\mathcal{E}_{t}(x)\in D\Big{\}} then,

lim supt1tlogsD(t,ε)supηD¯coinv(Λ)hη(X),for every ε>0,\limsup_{t\to\infty}\frac{1}{t}\log s_{D}(t,\varepsilon)\leq\sup_{\eta\in\overline{D}^{co}\cap\;\mathcal{M}_{inv}(\Lambda)}\,h_{\eta}(X),\quad\text{for every $\varepsilon>0$},

where D¯co\overline{D}^{co} denotes the closed convex hull of DD.
If, in addition, the entropy function inv(Λ)μhμ(X)\mathcal{M}_{inv}(\Lambda)\ni\mu\mapsto h_{\mu}(X) is upper semicontinuous, then

lim supt1tlogsD(t,ε)supηD¯inv(Λ)hη(X),for every ε>0,\limsup_{t\to\infty}\frac{1}{t}\log s_{D}(t,\varepsilon)\leq\sup_{\eta\in\overline{D}\cap\;\mathcal{M}_{inv}(\Lambda)}\,h_{\eta}(X),\quad\text{for every $\varepsilon>0$},

where D¯\overline{D} denotes the closure of DD.

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 D(Λ)D\subset\mathcal{M}(\Lambda) be a non-empty set and let ε>0\varepsilon>0. For each t>0t>0 let Et{xΛ:t(x)D}E_{t}\subset\{x\in\Lambda\colon\mathcal{E}_{t}(x)\in D\} be a (t,ε)(t,\varepsilon)-separated set (with respect to the flow Φ=(ϕt)t\Phi=(\phi_{t})_{t}) with cardinality sD(t,ε)s_{D}(t,\varepsilon). Choose a sequence (tn)n1(t_{n})_{n\geq 1} so that

lim supt1tlogsD(t,ε)=lim supn1tnlogsD(tn,ε),\limsup_{t\to\infty}\frac{1}{t}\log s_{D}(t,\varepsilon)=\limsup_{n\to\infty}\frac{1}{t_{n}}\log s_{D}(t_{n},\varepsilon), (6.4)

and consider the probability measures

σn:=1sD(tn,ε)xEtnδxandμn:=1sD(tn,ε)xEtntn(x).\sigma_{n}:=\frac{1}{s_{D}(t_{n},\varepsilon)}\sum_{x\in E_{t_{n}}}\delta_{x}\quad\text{and}\quad\mu_{n}:=\frac{1}{s_{D}(t_{n},\varepsilon)}\sum_{x\in E_{t_{n}}}\mathcal{E}_{t_{n}}(x).

Up to consider some convergent subsequence, we may assume without loss of generality that (μn)n1(\mu_{n})_{n\geq 1} is weak convergent to μ\mu. It is clear that μinv(Λ)\mu\in\mathcal{M}_{inv}(\Lambda). Moreover, as the sequence (μn)n1(\mu_{n})_{n\geq 1} is a convex combination of probability measures in DD then μD¯co\mu\in\overline{D}^{co}. Therefore, using (6.4), in order to prove the first statement in the lemma it is enough to show that

lim supn1tnlogsD(tn,ε)hμ(X).\limsup_{n\to\infty}\frac{1}{t_{n}}\log s_{D}(t_{n},\varepsilon)\leq h_{\mu}(X). (6.5)

By Gronwall’s inequality, there exists C>0C>0 so that

C1esXd(x,y)d(ϕs(x),ϕs(y))CesXd(x,y)C^{-1}e^{-s\,\|X\|_{\infty}}\,d(x,y)\leq d(\phi_{s}(x),\phi_{s}(y))\leq Ce^{s\,\|X\|_{\infty}}\,d(x,y)

for every x,yΛx,y\in\Lambda and s[0,1]s\in[0,1]. Since EtnE_{t_{n}} is a (tn,ε)(t_{n},\varepsilon)-separated set with respect to the flow Φ=(ϕt)t\Phi=(\phi_{t})_{t}, for any x,yEtnx,y\in E_{t_{n}}, there exists t[0,tn]t\in[0,t_{n}] such that d(ϕt(x),ϕt(y))>εd(\phi_{t}(x),\phi_{t}(y))>\varepsilon. Thus by the fact that tt[0,1]t-\lfloor t\rfloor\in[0,1], one has

d(ϕt(x),ϕt(y))C1eXd(ϕt(x),ϕt(y))>γd(\phi_{\lfloor t\rfloor}(x),\phi_{\lfloor t\rfloor}(y))\geq C^{-1}e^{-\|X\|_{\infty}}d(\phi_{t}(x),\phi_{t}(y))>\gamma

where γ=C1eXε\gamma=C^{-1}e^{-\|X\|_{\infty}}\varepsilon. Note that t[0,tn]\lfloor t\rfloor\in[0,\lfloor t_{n}\rfloor], so EtnE_{t_{n}} is a a (tn,γ)(\lfloor t_{n}\rfloor,\gamma)-separated set with respect to the time-one map ϕ1\phi_{1}. Choosing a partition 𝒫\mathcal{P} of Λ\Lambda with Diam(𝒫)<γ\operatorname{Diam}(\mathcal{P})<\gamma and μ(𝒫)=0\mu(\partial\mathcal{P})=0, one concludes that each element of the partition j=0tn1ϕj(𝒫)\bigvee_{j=0}^{\lfloor t_{n}\rfloor-1}\phi_{-j}(\mathcal{P}) contains at most one element of EtnE_{t_{n}}, and so

Hσn(j=0tn1ϕj(𝒫))=log#Etn=logsD(tn,ε).H_{\sigma_{n}}\Big{(}\bigvee_{j=0}^{\lfloor t_{n}\rfloor-1}\phi_{-j}(\mathcal{P})\Big{)}=\log\#E_{t_{n}}=\log{s_{D}(t_{n},\varepsilon)}.

It is not hard to check that the probability measures μ^n:=1sD(tn,ε)xEtntn(x)\hat{\mu}_{n}:=\frac{1}{s_{D}(t_{n},\varepsilon)}\sum_{x\in E_{t_{n}}}\mathcal{E}_{\lfloor t_{n}\rfloor}(x) converge to μ\mu 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

lim supn1tnlogsD(tn,ε)=lim supn1tnlogsD(tn,ε)hμ(ϕ1,𝒫)hμ(ϕ1)=hμ(X).\limsup_{n\to\infty}\frac{1}{t_{n}}\log{s_{D}(t_{n},\varepsilon)}=\limsup_{n\to\infty}\frac{1}{\lfloor t_{n}\rfloor}\log{s_{D}(t_{n},\varepsilon)}\leq h_{\mu}(\phi_{1},\mathcal{P})\leq h_{\mu}(\phi_{1})=h_{\mu}(X).

This proves  (6.5), and the first statement in the lemma.

Now, assume that D(Λ)D\subset\mathcal{M}(\Lambda) is a non-empty set and that inv(Λ)μhμ(X)\mathcal{M}_{inv}(\Lambda)\ni\mu\mapsto h_{\mu}(X) is upper semicontinuous. As D¯\overline{D} is compact, for each δ>0\delta>0 there exists a finite open cover {B(ηi,δ)}\{B(\eta_{i},\delta)\} of DD by balls of radius δ\delta. In particular there exists ηiδ(Λ)\eta_{i_{\delta}}\in\mathcal{M}(\Lambda) so that

lim supt1tlogsD(t,ε)=lim supt1tlogsDB(ηiδ,δ)(t,ε).\limsup_{t\to\infty}\frac{1}{t}\log s_{D}(t,\varepsilon)=\limsup_{t\to\infty}\frac{1}{t}\log s_{D\cap B(\eta_{i_{\delta}},\delta)}(t,\varepsilon).

Using the first statement of the lemma, there exists μδinv(Λ)B(ηiδ,δ)¯\mu_{\delta}\in\mathcal{M}_{inv}(\Lambda)\cap\overline{B(\eta_{i_{\delta}},\delta)} so that

lim supt1tlogsD(t,ε)hμδ(X).\limsup_{t\to\infty}\frac{1}{t}\log s_{D}(t,\varepsilon)\leq h_{\mu_{\delta}}(X).

In particular, any weak accumulation point μ\mu of (μδ)δ>0(\mu_{\delta})_{\delta>0} belongs to D¯\overline{D} and, by upper semicontinuity of the entropy, lim supt1tlogsD(t,ε)hμ(X).\limsup_{t\to\infty}\frac{1}{t}\log s_{D}(t,\varepsilon)\leq h_{\mu}(X). This completes the proof of the lemma. ∎

The previous result will allow to obtain the desired large deviations upper bounds. We observe that hμ(X)=hμ(f)h_{\mu}(X)=h_{\mu}(f) where f=ϕ1f=\phi_{1} denotes the time-1 of the flow Φ=(ϕt)t\Phi=(\phi_{t})_{t}. 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 Λ\Lambda be an invariant compact set of a vector field X𝒳1(M)X\in\mathcal{X}^{1}(M) and let 𝒦(Λ)\mathcal{K}\subset\mathcal{M}(\Lambda) be a closed and convex subset so that 𝒦inv(Λ)\mathcal{K}\cap\mathcal{M}_{inv}(\Lambda)\neq\emptyset. Assume μψinv(Λ)\mu_{\psi}\in\mathcal{M}_{inv}(\Lambda) is a weak Gibbs measure with respect to a Hölder continuous potential ψ:Λ\psi\colon\Lambda\rightarrow\mathbb{R} and ΛHΛ\Lambda_{H}\subset\Lambda be the μψ\mu_{\psi}-full measure set satisfied for (2.1). Consider the non-positive real number

c:=lim supδ0lim supt1tlogμψ({xΛ:Ct(x,ε)>eδt}).c_{\infty}:=\limsup_{\delta\to 0}\;\limsup_{t\to\infty}\frac{1}{t}\log\mu_{\psi}\Big{(}\big{\{}x\in\Lambda\colon C_{t}(x,\varepsilon)>e^{\delta t}\big{\}}\Big{)}. (6.6)

Then

lim supt\displaystyle\limsup_{t\to\infty} 1tlogμψ({xΛ:t(x)𝒦})\displaystyle\frac{1}{t}\log\mu_{\psi}\big{(}\{x\in\Lambda\colon\mathcal{E}_{t}(x)\in\mathcal{K}\}\big{)}
max{supμ𝒦inv(Λ)(Ptop(Λ,ψ)+hν(X)+ψdν),c}.\displaystyle\leq\max\Big{\{}\sup_{\mu\in\mathcal{K}\cap\mathcal{M}_{inv}(\Lambda)}\Big{(}\displaystyle-P_{\rm top}(\Lambda,\psi)+h_{\nu}(X)+\int{\psi}\,{\rm d}\nu\Big{)}\;,\;c_{\infty}\Big{\}}. (6.7)

Furthermore, if the entropy function inv(Λ)μhμ(X)\mathcal{M}_{inv}(\Lambda)\ni\mu\mapsto h_{\mu}(X) is upper-semicontinuous then  (6.7) holds even if 𝒦\mathcal{K} is not convex.

Proof.

Assume first that 𝒦(Λ)\mathcal{K}\subset\mathcal{M}(\Lambda) is a closed and convex subset. As ψ:Λ\psi\colon\Lambda\rightarrow\mathbb{R} is Hölder continuous, then it is bounded and, given δ>0\delta>0, one can write 𝒦=j=0Nδ𝒦j\mathcal{K}=\bigcup_{j=0}^{N_{\delta}}\mathcal{K}_{j} where

𝒦j={η𝒦:ψ𝑑η[ψ+jδ,ψ+(j+1)δ]},\mathcal{K}_{j}=\Big{\{}\eta\in\mathcal{K}\colon\int\psi\,d\eta\in[-\|\psi\|_{\infty}+j\delta,-\|\psi\|_{\infty}+(j+1)\delta]\Big{\}},

Nδ=2δψ+1N_{\delta}=\lfloor\frac{2}{\delta}\|\psi\|_{\infty}\rfloor+1. Note that some of the sets 𝒦j\mathcal{K}_{j}, which are closed and convex, may be empty. For each non-empty 𝒦j\mathcal{K}_{j}, either there exists t>0t_{*}>0 so that μψ({xΛ:t(x)𝒦j})=0\mu_{\psi}(\{x\in\Lambda\colon\mathcal{E}_{t}(x)\in\mathcal{K}_{j}\})=0 for every t>tt>t_{*} or 𝒦jinv(Λ)\mathcal{K}_{j}\cap\mathcal{M}_{inv}(\Lambda)\neq\emptyset. For that reason we will assume, without loss of generality, that all 𝒦j\mathcal{K}_{j}\neq\emptyset intersect the space of invariant probability measures. Now, for each 0jNδ0\leq j\leq N_{\delta} so that 𝒦j\mathcal{K}_{j}\neq\emptyset, one has

{xΛ:t(x)𝒦j}{xΛ:t(x)𝒦j&Ct(x,ε)eδt}{xΛ:Ct(x,ε)>eδt}.\big{\{}x\in\Lambda\colon\mathcal{E}_{t}(x)\in\mathcal{K}_{j}\big{\}}\subset\big{\{}x\in\Lambda\colon\mathcal{E}_{t}(x)\in\mathcal{K}_{j}\,\&\,C_{t}(x,\varepsilon)\leq e^{\delta t}\big{\}}\cup\big{\{}x\in\Lambda\colon C_{t}(x,\varepsilon)>e^{\delta t}\big{\}}. (6.8)

The maximal cardinality of a (t,ε)(t,\varepsilon)-separated set in the first set in the right hand-side above is bounded above by s𝒦j(t,ε)s_{\mathcal{K}_{j}}(t,\varepsilon) which, by Lemma 6.4, satisfies

lim supt1nlogs𝒦j(t,ε)supη𝒦jinv(X)hη(X),for every ε>0.\limsup_{t\to\infty}\frac{1}{n}\log s_{\mathcal{K}_{j}}(t,\varepsilon)\leq\sup_{\eta\in{\mathcal{K}_{j}}\cap\;\mathcal{M}_{inv}(X)}\,h_{\eta}(X),\quad\text{for every $\varepsilon>0$}.

Given ε>0\varepsilon>0 small and fixed, pick a (t,ε)(t,\varepsilon)-maximal separated set Ej,t{xΛ:t(x)𝒦j}E_{j,t}\subset\big{\{}x\in\Lambda\colon\mathcal{E}_{t}(x)\in\mathcal{K}_{j}\big{\}}. If 𝒦j\mathcal{K}_{j}\neq\emptyset then, using the weak Gibbs property, one ensures that

μψ({xΛ:t(x)𝒦j&Ct(x,ε)eδt})\displaystyle\mu_{\psi}\Big{(}\big{\{}x\in\Lambda\colon\mathcal{E}_{t}(x)\in\mathcal{K}_{j}\,\&\,C_{t}(x,\varepsilon)\leq e^{\delta t}\big{\}}\Big{)} xEj,tμψ(B(x,t,ε,Φ))\displaystyle\leq\sum_{x\in E_{j,t}}\mu_{\psi}\Big{(}B\big{(}x,t,\varepsilon,\Phi\big{)}\Big{)}
xEj,teδtetPtop(X,ψ)+tψ𝑑t(x)\displaystyle\leq\sum_{x\in E_{j,t}}e^{\delta t}\;e^{-t\,P_{\rm top}(X,\psi)+t\int\psi\,d\mathcal{E}_{t}(x)}
s𝒦j(t,ε)eδtetPtop(X,ψ)+tsupη𝒦jψ𝑑η\displaystyle\leq s_{\mathcal{K}_{j}}(t,\varepsilon)\;e^{\delta t}\;e^{-t\,P_{\rm top}(X,\psi)+t\sup_{\eta\in\mathcal{K}_{j}}\int\psi\,d\eta}
s𝒦j(t,ε)e2δtetPtop(X,ψ)+tsupη𝒦jinv(X)ψ𝑑η\displaystyle\leq s_{\mathcal{K}_{j}}(t,\varepsilon)\;e^{2\delta t}\;e^{-t\,P_{\rm top}(X,\psi)+t\sup_{\eta\in\mathcal{K}_{j}\cap\mathcal{M}_{inv}(X)}\int\psi\,d\eta}

and, consequently,

lim supt1tlogμψ\displaystyle\limsup_{t\to\infty}\frac{1}{t}\log\mu_{\psi} ({xΛ:t(x)𝒦j&Ct(x,ε)eδt})\displaystyle\Big{(}\big{\{}x\in\Lambda\colon\mathcal{E}_{t}(x)\in\mathcal{K}_{j}\,\&\,C_{t}(x,\varepsilon)\leq e^{\delta t}\big{\}}\Big{)}
supη𝒦jinv(Λ){Ptop(X,ψ)+hη(X)+ψ𝑑η}+2δ\displaystyle\leq\sup_{\eta\in{\mathcal{K}_{j}}\cap\mathcal{M}_{inv}(\Lambda)}\,\Big{\{}-P_{\rm top}(X,\psi)+h_{\eta}(X)+\int\psi\,d\eta\Big{\}}+2\delta

This, combined with  (6.8), ensures that lim supt1tlogμψ({xΛ:t(x)𝒦j})\limsup_{t\to\infty}\frac{1}{t}\log\mu_{\psi}\big{(}\{x\in\Lambda\colon\mathcal{E}_{t}(x)\in\mathcal{K}_{j}\}\big{)} is bounded above by

max{\displaystyle\max\Big{\{} supη𝒦inv(Λ){Ptop(X,ψ)+hη(X)+ψ𝑑η}+2δ,\displaystyle\sup_{\eta\in\mathcal{K}\cap\mathcal{M}_{inv}(\Lambda)}\,\Big{\{}-P_{\rm top}(X,\psi)+h_{\eta}(X)+\int\psi\,d\eta\Big{\}}+2\delta\;,
lim supt1tlogμψ({xΛ:Ct(x,ε)>eδt})}\displaystyle\limsup_{t\to\infty}\frac{1}{t}\log\mu_{\psi}\Big{(}\big{\{}x\in\Lambda\colon C_{t}(x,\varepsilon)>e^{\delta t}\big{\}}\Big{)}\Big{\}}

for each small δ>0\delta>0. Taking the lim sup\limsup as δ0\delta\to 0 in each of the terms in the previous expression we conclude that

lim supt\displaystyle\limsup_{t\to\infty} 1tlogμψ({xΛ:t(x)𝒦})\displaystyle\frac{1}{t}\log\mu_{\psi}\big{(}\{x\in\Lambda\colon\mathcal{E}_{t}(x)\in\mathcal{K}\}\big{)}
max{supμ𝒦inv(Λ)(Ptop(Λ,ψ)+hν(X)+ψdν),c},\displaystyle\leq\max\Big{\{}\sup_{\mu\in\mathcal{K}\cap\mathcal{M}_{inv}(\Lambda)}\Big{(}\displaystyle-P_{\rm top}(\Lambda,\psi)+h_{\nu}(X)+\int{\psi}\,{\rm d}\nu\Big{)},c_{\infty}\Big{\}},

thus proving the first statement in the proposition. Finally, if inv(M)μhμ(ϕ1)\mathcal{M}_{inv}(M)\ni\mu\mapsto h_{\mu}(\phi_{1}) is upper semicontinuous then the large deviations upper bound holds for arbitrary closed sets 𝒦\mathcal{K} 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 X𝒳1(M)X\in\mathscr{X}^{1}(M) and Λ\Lambda be a singular hyperbolic homoclinic class such that each pair of periodic orbits in Λ\Lambda are homoclinically related and 1(Λ)¯=inv(Λ)\overline{\mathcal{M}_{1}(\Lambda)}=\mathcal{M}_{inv}(\Lambda). Assume μψ\mu_{\psi} is a Gibbs measure with respect to a Hölder continuous potential ψ:Λ\psi\colon\Lambda\to\mathbb{R}. For any continuous observable g:Λg\colon\Lambda\to\mathbb{R} it holds that

lim supt+1tlogμψ({xΛ:1t0tg(ϕs(x))ds[a,b]})infs[a,b]Iψ,g(s)\displaystyle\limsup_{t\to+\infty}\frac{1}{t}\log\mu_{\psi}\Big{(}\Big{\{}x\in\Lambda\colon\frac{1}{t}\int_{0}^{t}g(\phi_{s}(x))\,{\rm d}s\in[a,b]\Big{\}}\Big{)}\leq-\inf_{s\in[a,b]}I_{\psi,g}(s)

and

lim inft+1tlogμψ({xΛ:1t0tg(ϕs(x))ds(a,b)})infs(a,b)Iψ,g(s)\displaystyle\liminf_{t\to+\infty}\frac{1}{t}\log\mu_{\psi}\Big{(}\Big{\{}x\in\Lambda\colon\frac{1}{t}\int_{0}^{t}g(\phi_{s}(x))\,{\rm d}s\in(a,b)\Big{\}}\Big{)}\geq-\inf_{s\in(a,b)}I_{\psi,g}(s)

where the lower-semicontinuous rate function Iψ,gI_{\psi,g} is given by

Iψ,g(s)=sup{Ptop(Λ,ψ)hη(X)ψdη:ηinv(Λ),gdη=s}.I_{\psi,g}(s)=\sup\Big{\{}P_{\rm top}(\Lambda,\psi)-h_{\eta}(X)-\int{\psi}\,{\rm d}\eta\colon\eta\in\mathcal{M}_{inv}(\Lambda),\;\int g\,{\rm d}\eta=s\Big{\}}.

Moreover, if there exist μ1,μ2inv(Λ)\mu_{1},\mu_{2}\in\mathcal{M}_{inv}(\Lambda) so that gdμ1gdμ2\displaystyle\int g{\rm d}\mu_{1}\neq\int g{\rm d}\mu_{2} and gdμψ[a,b]\displaystyle\int g\,{\rm d}\mu_{\psi}\notin[a,b] then the infima in the right-hand side of the previous inequalities are strictly negative.

Remark.

Similarly as Theorem ABC, the conclusion of Corollary 6.6 holds for Lorenz attractors of vector fields in a residual subset r𝒳r(M3),(r2)\mathcal{R}^{r}\subset\mathcal{X}^{r}(M^{3}),~{}(r\in\mathbb{N}_{\geq 2}), and also holds for singular hyperbolic attractors of vector fields in a residual subset 𝒳1(M)\mathcal{R}\subset\mathcal{X}^{1}(M).

Acknowledgments

The authors would like to thank the anonymous referees for their valuable comments and suggestions. The authors would also like to thank Professor Dejun Feng who pointed out to us the second item of Remark 3.2 and Remark 5.2.

References

  • [1] F. Abdenur, C. Bonatti and S. Crovisier, Nonuniform hyperbolicity for C1C^{1}-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. Afraı˘\breve{\rm\i}movicˇ\check{\rm c}, 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 d\mathbb{Z}^{d}-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 187188, 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 β\beta-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.

Yi Shi

School of Mathematics, Sichuan University, Chengdu, 610000, China

E-mail: [email protected]