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

A Mañé-Manning formula for expanding
measures for endomorphisms of k\mathbb{P}^{k}

Fabrizio Bianchi CNRS, Univ. Lille, UMR 8524 - Laboratoire Paul Painlevé, F-59000 Lille, France fabrizio.bianchi@@univ-lille.fr  and  Yan Mary He Department of Mathematics
University of Oklahoma
Norman, OK 73019
[email protected]
Abstract.

Let k1k\geq 1 be an integer and ff a holomorphic endomorphism of k()\mathbb{P}^{k}(\mathbb{C}) of algebraic degree d2d\geq 2. We introduce a volume dimension for ergodic ff-invariant probability measures with strictly positive Lyapunov exponents. In particular, this class of measures includes all ergodic measures whose measure-theoretic entropy is strictly larger than (k1)logd(k-1)\log d, a natural generalization of the class of measures of positive measure-theoretic entropy in dimension 1. The volume dimension is equivalent to the Hausdorff dimension when k=1k=1, but depends on the dynamics of ff to incorporate the possible failure of Koebe’s theorem and the non-conformality of holomorphic endomorphisms for k2k\geq 2.

If ν\nu is an ergodic ff-invariant probability measure with strictly positive Lyapunov exponents, we prove a generalization of the Mañé-Manning formula relating the volume dimension, the measure-theoretic entropy, and the sum of the Lyapunov exponents of ν\nu. As a consequence, we give a characterization of the first zero of a natural pressure function for such expanding measures in terms of their volume dimensions. For hyperbolic maps, such zero also coincides with the volume dimension of the Julia set, and with the exponent of a natural (volume-)conformal measure. This generalizes results by Denker-Urbański and McMullen in dimension 1 to any dimension k1k\geq 1.

Our methods mainly rely on a theorem by Berteloot-Dupont-Molino, which gives a precise control on the distortion of inverse branches of endomorphisms along generic inverse orbits with respect to measures with strictly positive Lyapunov exponents.

1. Introduction

Let f:1()1()f:\mathbb{P}^{1}(\mathbb{C})\to\mathbb{P}^{1}(\mathbb{C}) be a rational map of degree d2d\geq 2 and ν\nu an ergodic ff-invariant probability measure whose Lyapunov exponent is strictly positive. Such a measure is necessarily supported on the Julia set J(f)J(f) of ff. There is a well-known relation between the Hausdorff dimension HD(ν)\operatorname{HD}(\nu), the measure-theoretic entropy hν(f)h_{\nu}(f), and the Lyapunov exponent χν(f)\chi_{\nu}(f) of ν\nu; namely, we have

(1.1) HD(ν)=hν(f)χν(f).\operatorname{HD}(\nu)=\frac{h_{\nu}(f)}{\chi_{\nu}(f)}.

This formula is usually referred to as the Mañé-Manning formula; see [Man84, Mañ88]. Hofbauer and Raith [HR92] proved a version of (1.1) for piecewise monotone maps on the unit interval with bounded variation; see also [Led81]. The fact that (1.1) holds in one-dimensional complex dynamics crucially relies on distortion estimates for univalent holomorphic maps coming from Koebe’s theorem; see Section 1.2.

For smooth dynamical systems in higher dimensions, related formulas are known to hold in a number of settings. If f:MMf:M\to M is a diffeomorphism of a compact manifold MM and ν\nu is an ergodic probability measure on MM which is absolutely continuous with respect to the Lebesgue measure, Pesin [Pes77] proved that

hν(f)=χν+(f)h_{\nu}(f)=\chi^{+}_{\nu}(f)

where χν+(f)\chi^{+}_{\nu}(f) is the sum of the non-negative Lyapunov exponents of ff counted with multiplicity; see also [Mañ81]. When MM is a surface, Young [You82] proved that

HD(ν)=hν(f)χ++hν(f)|χ|\operatorname{HD}(\nu)=\frac{h_{\nu}(f)}{\chi_{+}}+\frac{h_{\nu}(f)}{|\chi_{-}|}

when ν\nu is ergodic and χ<0<χ+\chi_{-}<0<\chi_{+} are its Lyapunov exponents. This formula has been generalized to the case of diffeomorphisms in any dimension; see [LY85] and [BPS99]. Such systems display attracting and repelling directions, and one decomposes the problem into two problems, one for ff (along unstable manifolds) and one for f1f^{-1} (along stable manifolds). The Mañé-Manning formula (1.1) can be seen as a version of Young’s result in (complex) dimension 1 where the system is not invertible. In this paper, we address the validity of (1.1) in several complex variables, and more specifically for expanding measures for (non-invertible) holomorphic endomorphisms of projective spaces in any dimension.

Let k1k\geq 1 be an integer and denote k . . =k()\mathbb{P}^{k}\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\mathbb{P}^{k}(\mathbb{C}). If f:kkf:\mathbb{P}^{k}\to\mathbb{P}^{k} is a holomorphic endomorphism of algebraic degree d2d\geq 2, it is not hard to find examples where (1.1), with χν\chi_{\nu} replaced by the sum of the Lyapunov exponents of ν\nu (the natural generalization of the expansion rate along generic orbits), does not hold. For instance, one can consider product self-maps of 2\mathbb{C}^{2} of the form (z,w)(z2+a1,w2+a2)(z,w)\mapsto(z^{2}+a_{1},w^{2}+a_{2}), where aia_{i}\in\mathbb{C} are such that the measures of maximal entropy of each component have different Hausdorff dimensions.

In [BD03], Binder-DeMarco proposed a conjectural formula for the Hausdorff dimension of the measure of maximal entropy μ\mu of an endomorphism of k\mathbb{P}^{k} as follows:

HD(μ)=logdχ1++logdχk.\operatorname{HD}(\mu)=\frac{\log d}{\chi_{1}}+\dots+\frac{\log d}{\chi_{k}}.

This conjecture has been partially settled [BD03, DD04, Dup11], and also versions of it have been proposed (and partially proved) for more general invariant measures [DD04, Dup11, Dup12, dV15, DR20]. In this paper, we introduce a natural dimension VD(ν)\operatorname{VD}(\nu) for ergodic ff-invariant measures ν\nu with strictly positive Lyapunov exponents and show that this dimension satisfies a natural generalization of (1.1), where χν\chi_{\nu} is replaced by (two times) the sum of the Lyapunov exponents.

1.1. Statement of results

Let f:kkf:\mathbb{P}^{k}\to\mathbb{P}^{k} be a holomorphic endomorphism of algebraic degree d2d\geq 2. The Julia set J(f)J(f) of ff is the support of the unique measure of maximal entropy of ff [Lyu83, BD01, DS10]. Let +(f)\mathcal{M}^{+}(f) (resp. J+(f)\mathcal{M}^{+}_{J}(f)) be the set of ergodic invariant probability measures on k\mathbb{P}^{k} (resp. on J(f)J(f)) with strictly positive Lyapunov exponents. The set J+(f)\mathcal{M}_{J}^{+}(f) contains the set e+(f)\mathcal{M}_{e}^{+}(f) of all ergodic probability measures whose measure-theoretic entropy is strictly larger than (k1)logd(k-1)\log d [deT08, Dup12], which are the natural generalization of the ergodic measures with strictly positive entropy in dimension 11. Large classes of examples of measures in e+(f)\mathcal{M}^{+}_{e}(f) were constructed and studied in [Dup12, UZ13, SUZ14, BD23, BD22].

We introduce a volume dimension for measures ν+(f)\nu\in\mathcal{M}^{+}(f); see Section 1.2 for an overview and Section 4 for precise definitions. The volume dimension is dynamical in nature and generalizes the notion of Hausdorff dimension in dimension 1 to higher dimensions to incorporate the failure of Koebe’s theorem and the non-conformality of holomorphic endomorphisms.

For ν+(f)\nu\in\mathcal{M}^{+}(f), we denote by VD(ν)\operatorname{VD}(\nu) the volume dimension, hν(f)h_{\nu}(f) the measure-theoretic entropy, and Lν(f)L_{\nu}(f) the sum of the Lyapunov exponents of ν\nu. The main result of this paper relates these three quantities and generalizes the Mañé-Manning formula to any k1k\geq 1.

Theorem 1.1.

Let f:kkf:\mathbb{P}^{k}\to\mathbb{P}^{k} be a holomorphic endomorphism of algebraic degree d2d\geq 2. For every ν+(f)\nu\in\mathcal{M}^{+}(f) we have

VD(ν)=hν(f)2Lν(f).\operatorname{VD}(\nu)=\frac{h_{\nu}(f)}{2L_{\nu}(f)}.

When k=1k=1, Theorem 1.1 reduces to the Mañé-Manning formula (1.1), as in this case we have 2VD(ν)=HD(ν)2\operatorname{VD}(\nu)=\operatorname{HD}(\nu); see Proposition 4.20. The factor 2=2k/k2=2k/k is due to the fact that we weight open sets of covers by their volume instead of their diameter and we have kk Lyapunov exponents, counting multiplicities.

As an application of Theorem 1.1, we study a number of natural dimensions and quantities associated to an endomorphism ff. In dimension 1, these quantities are already defined and well studied; see for example [DU91, DU91a, PU10, McM00]. We first define a dynamical dimension DDJ+(f)\operatorname{DD}^{+}_{J}(f) of ff as

DDJ+(f)\displaystyle\operatorname{DD}^{+}_{J}(f) . . =sup{VD(ν):νJ+(f)}.\displaystyle\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\sup\left\{\operatorname{VD}(\nu)\colon\nu\in\mathcal{M}_{J}^{+}(f)\right\}.

For k=1k=1, recall that the pressure function is defined as

P(t) . . =supν{hν(f)tχν(f)},P(t)\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\sup_{\nu}\left\{h_{\nu}(f)-t\chi_{\nu}(f)\right\},

where tt\in\mathbb{R} and the supremum is taken over the set of invariant probability measures on J(f)J(f). In fact, the supremum can be taken over νJ+(f)=+(f)\nu\in\mathcal{M}^{+}_{J}(f)=\mathcal{M}^{+}(f). This can be seen by combining Ruelle’s inequality [Rue78] with a theorem of Przytycki [Prz93] stating that all invariant measures supported on the Julia set of a rational map have non-negative Lyapunov exponent.

For any k1k\geq 1, we define in a similar way a pressure function PJ+(t)P_{J}^{+}(t) as

PJ+(t) . . =sup{hν(f)tLν(f):νJ+(f)}.P^{+}_{J}(t)\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\sup\left\{h_{\nu}(f)-tL_{\nu}(f)\colon\nu\in\mathcal{M}_{J}^{+}(f)\right\}.

By the above, we have PJ+(t)=P(t)P^{+}_{J}(t)=P(t) when k=1k=1. We remark that, for any k2k\geq 2, there may exist ergodic probability measures ν\nu on J(f)J(f) with Lν(f)<0L_{\nu}(f)<0; see Section 2.4 for examples and further comments. However, as in the case of k=1k=1, the pressure function PJ+(t)P_{J}^{+}(t) is still non-increasing and convex for all k1k\geq 1; see Lemma 2.14. We define

pJ+(f) . . =inf{t>0:PJ+(t)0}.p_{J}^{+}(f)\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\inf\left\{t>0\colon P_{J}^{+}(t)\leq 0\right\}.

As a consequence of Theorem 1.1, we have the following result which generalizes a theorem due to Denker-Urbański [DU91, DU91a] in the case of rational maps to any dimension.

Theorem 1.2.

Let f:kkf\colon\mathbb{P}^{k}\to\mathbb{P}^{k} be a holomorphic endomorphism of algebraic degree d2d\geq 2. Then we have

2DDJ+(f)=pJ+(f).2\operatorname{DD}_{J}^{+}(f)=p_{J}^{+}(f).

Finally, in the spirit of the celebrated Bowen-Ruelle formula for hyperbolic maps [Bow79, Rue82], we give an interpretation of pJ+(f)p_{J}^{+}(f), when ff is hyperbolic (i.e., uniformly expanding on J(f)J(f); see Section 2.1) in terms of (volume-)conformal measures. Given t0t\geq 0, we say that a probability measure ν\nu on J(f)J(f) is tt-volume-conformal on J(f)J(f) if, for every Borel subset AJ(f)A\subset J(f) on which ff is invertible, we have

ν(f(A))=A|Jacf|t𝑑ν\nu(f(A))=\int_{A}|\operatorname{Jac}f|^{t}d\nu

and define

δJ(f)\displaystyle\delta_{J}(f) . . =inf{t0: there exists a t-volume-conformal measure on J(f)}.\displaystyle\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\inf\left\{t\geq 0\colon\mbox{ there exists a $t$-volume-conformal measure on }J(f)\right\}.

For k=1k=1, the definitions of tt-volume-conformal measures and δJ(f)\delta_{J}(f) reduce to those of conformal measures and conformal dimension for rational maps; see [DU91, DU91a, McM00, PU10]. In this case, owing to Bowen [Bow79], one sees that

δJ(f)=pJ+(f)=HD(J(f))\delta_{J}(f)=p_{J}^{+}(f)=\operatorname{HD}(J(f))

for every hyperbolic rational map ff on 1()\mathbb{P}^{1}(\mathbb{C}), and that there exists a unique ergodic measure ν\nu on J(f)J(f) such that HD(ν)=HD(J(f))\operatorname{HD}(\nu)=\operatorname{HD}(J(f)). We have here the following result in any dimension, which further motivates the definition of the volume dimension as a natural generalization of the Hausdorff dimension for all k1k\geq 1. Observe that, if ff is hyperbolic, every invariant probability measure ν\nu on J(f)J(f) belongs to +(f)\mathcal{M}^{+}(f).

Theorem 1.3.

Let f:kkf\colon\mathbb{P}^{k}\to\mathbb{P}^{k} be a hyperbolic holomorphic endomorphism of algebraic degree d2d\geq 2. Then we have

δJ(f)=pJ+(f)=2VD(J(f))\delta_{J}(f)=p_{J}^{+}(f)=2\operatorname{VD}(J(f))

and there exists a unique ergodic measure ν\nu on J(f)J(f) such that VD(ν)=VD(J(f))\operatorname{VD}(\nu)=\operatorname{VD}(J(f)).

Remark 1.4.

As all our arguments will be local, our results apply more generally to the setting of polynomial-like maps in any dimension, i.e., proper holomorphic maps of the form f:UVf\colon U\to V, with UVkU\Subset V\Subset\mathbb{C}^{k} and VV convex [DS03, DS10]. For a large class of such maps (i.e., those whose topological degree dominates all the other dynamical degrees [BDR23]), an analogue of the inclusion e+(f)J+(f)\mathcal{M}_{e}^{+}(f)\subset\mathcal{M}^{+}_{J}(f) in this more general context has been proved in [BR22].

As every endomorphism of k\mathbb{P}^{k} lifts to a homogeneous polynomial endomorphism of k+1\mathbb{C}^{k+1}, we can assume for simplicity that the maps we consider are polynomials. Observe that the Lyapunov exponents of every lifted measure are the same as those of the original measure, with the addition of an extra exponent logd\log d. Since logd>0\log d>0 when d2d\geq 2, this does not change the condition on the positivity of the Lyapunov exponents.

1.2. Volume dimensions and strategy of the proofs

Let us first recall the idea of the proof of the Mañé-Manning formula (1.1) in dimension 1. It essentially consists of two steps.

  1. (1)

    The first step consists of defining a local dimension at a point xx by setting

    δx . . =limr0logν(B(x,r))logr\delta_{x}\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\lim_{r\to 0}\frac{\log\nu(B(x,r))}{\log r}

    (whenever the limit exists), where B(x,r)B(x,r) denotes the balls of radius rr centred and xx, and proving that the limit is well-defined and equal to the ratio hν(f)/χν(f)h_{\nu}(f)/\chi_{\nu}(f) for ν\nu-almost every xx. In particular, ν\nu is exact-dimensional.

  2. (2)

    The second step is to prove that the Hausdorff dimension of ν\nu must be equal to the common value of the local dimensions found in the first step [You82].

Let us describe how the one-dimensional setting plays a crucial role in Step (1). By [BK83] and [Mañ81], for ν\nu-almost every xx we have

hν(f)=limκ0limnlogν(Bn(x,κ))n,h_{\nu}(f)=\lim_{\kappa\to 0}\lim_{n\to\infty}\frac{-\log\nu(B_{n}(x,\kappa))}{n},

where Bn(x,κ)B_{n}(x,\kappa) is the Bowen ball of radius κ\kappa and depth nn. This is defined as

Bn(x,κ) . . ={y:|fj(y)fj(x)|<κ,0jn}.B_{n}(x,\kappa)\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\left\{y\colon|f^{j}(y)-f^{j}(x)|<\kappa,0\leq j\leq n\right\}.

The crucial observation is that, for large nn, the Bowen ball Bn(x,κ)B_{n}(x,\kappa) is comparable (up to precisely quantifiable errors) to the ball B(x,κenχν(f))B(x,\kappa\,e^{-n\chi_{\nu}(f)}) of the same center and radius κenχν(f)\kappa\,e^{-n\chi_{\nu}(f)}. Fixing a κ0\kappa_{0} for simplicity, and setting n(r)|logr|/χν(f)n(r)\sim|\log r|/\chi_{\nu}(f), it then follows that

limr0logν(B(x,r))logr=limr0logν(Bn(r)(x,κ0))n(r)n(r)logr=hν(f)χν,\lim_{r\to 0}\frac{\log\nu(B(x,r))}{\log r}=\lim_{r\to 0}\frac{-\log\nu(B_{n(r)}(x,\kappa_{0}))}{n(r)}\frac{n(r)}{-\log r}=\frac{h_{\nu}(f)}{\chi_{\nu}},

which in particular shows that δx\delta_{x} is well-defined. The precise relation between geometric balls and Bowen balls is a consequence of Koebe’s theorem and related distortion estimates, which imply that images of balls by holomorphic maps (and in particular by their inverse branches) are still comparable to balls. As a consequence, in complex dimension 11, there is a natural interplay between the Hausdorff dimension and the dynamics of a rational map. Observe in particular that one may define the Hausdorff dimension of ν\nu by using covers consisting of Bowen balls, indexed over their depth nn, and sending nn to infinity; see also [CPZ19].

All the above is in sharp contrast with the higher-dimensional situation, where, due to the lack of conformality of holomophic maps, preimages of balls can be arbitrarily distorted, and far from being balls. In the best possible scenario (e.g., for hyperbolic product maps), the preimages of balls are approximately ellipses whose axes reflect the contraction rate of the inverse branches in the different directions.

On the other hand, when ν+(f)\nu\in\mathcal{M}^{+}(f), a result by Berteloot-Dupont-Molino (see [BDM08, BD19] and Theorem 2.1 below) states that the best possible scenario described above is actually true, in an infinitesimal sense, for preimages of balls along generic orbits of ν\nu. More precisely, there exists an increasing (as ϵ0\epsilon\to 0) measurable exhaustion {Zν(ϵ)}ϵ\{Z^{\star}_{\nu}(\epsilon)\}_{\epsilon} of a full-measure subset ZνZ^{\star}_{\nu} of the space of orbits for ff such that the preimages of sufficiently small balls along orbits in Zν(ϵ)Z^{\star}_{\nu}(\epsilon) are approximately ellipses, and the contraction rate for their volume is essentially given (up to further controllable error terms) by enLν(f)+nO(ϵ)e^{-nL_{\nu}(f)+n{O(\epsilon)}}. This is a consequence of very refined estimates on the convexity defect of such preimages. Such property was already exploited in [BB18] to give bounds on the Hausdorff dimension of the bifurcation locus of families of endomorphisms of k\mathbb{P}^{k} [BBD18, Bia19], and in particular to prove that this is maximal near isolated Lattés maps, i.e., maps for which all the Lyapunov exponents are equal and minimal, i.e., equal to (logd)/2(\log d)/2 [BD99, BD05].

Fix ν+(f)\nu\in\mathcal{M}^{+}(f). Denote by π:Zνk\pi\colon Z^{\star}_{\nu}\to\mathbb{P}^{k} the projection associating to any orbit z^={zn}n\hat{z}=\{z_{n}\}_{n\in\mathbb{Z}} its element z0z_{0}. For xπ(Zν(ϵ))x\in\pi(Z^{\star}_{\nu}(\epsilon)), κ>0\kappa>0, and NN\in\mathbb{N}, we consider (when well-defined) the neighbourhood U=U(N,x,κ,ϵ)U=U(N,x,\kappa,\epsilon) of xx satisfying

fN(U)=B(fN(x),κeNMϵ)f^{N}(U)=B(f^{N}(x),\kappa\,e^{-NM\epsilon})

where eMe^{M} is a bound for the expansion of ff and we require that fN|Uf^{N}|_{U} is injective. It follows from the above result by Berteloot-Dupont-Molino, and by further estimates that we develop in Section 2, that there exist some r(ϵ)r(\epsilon) and n(ϵ)n(\epsilon) such that, for all xπ(Zν(ϵ))x\in\pi(Z^{\star}_{\nu}(\epsilon)), 0<κ<r(ϵ)0<\kappa<r(\epsilon), and Nn(ϵ)N\geq n(\epsilon) the sets U(N,x,κ,ϵ)U(N,x,\kappa,\epsilon) are indeed well-defined and approximately ellipses, of controlled geometry. We see these sets U(N,x,κ,ϵ)U(N,x,\kappa,\epsilon) as a suitable version of the Bowen balls Bn(x,κ)B_{n}(x,\kappa) in any dimension. Let us set

δx(ϵ,κ,N) . . =logν(U(N,x,κ,ϵ))logVol(U(N,x,κ,ϵ)),\delta_{x}(\epsilon,\kappa,N)\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\frac{\log\nu(U(N,x,\kappa,\epsilon))}{\log{\rm Vol}(U(N,x,\kappa,\epsilon))},

where Vol\rm Vol denotes the volume with respect to the Fubini-Study metric. As a first step (which corresponds to Step (1) above) towards proving Theorem 1.1, we show that every ν+(f)\nu\in\mathcal{M}^{+}(f) is exact (volume-)dimensional; namely, for ν\nu-almost every xx, we have

lim supϵ0lim supκ0lim supNδx(ϵ,κ,N)=lim infϵ0lim infκ0lim infNδx(ϵ,κ,N)=hν(f)2Lν(f);\limsup_{\epsilon\to 0}\limsup_{\kappa\to 0}\limsup_{N\to\infty}\delta_{x}(\epsilon,\kappa,N)=\liminf_{\epsilon\to 0}\liminf_{\kappa\to 0}\liminf_{N\to\infty}\delta_{x}(\epsilon,\kappa,N)=\frac{h_{\nu}(f)}{2L_{\nu}(f)};

see Theorem 3.2 and Corollary 3.4. We adapt here the approach of Mañé [Mañ88] in higher dimensions, thanks to the distortion estimates developed in Section 2.

Once the local dimension of every ν+(f)\nu\in\mathcal{M}^{+}(f) is well-defined as above, we give a global interpretation of this quantity by defining a volume dimension for these measures. The idea is to use the sets U(N,x,κ,ϵ)U(N,x,\kappa,\epsilon) to cover the “slice” Xπ(Zν(ϵ))X\cap\pi(Z^{\star}_{\nu}(\epsilon)) of every set XZνX\subseteq Z^{\star}_{\nu}. More precisely, for every Xπ(Zν)X\subseteq\pi(Z^{\star}_{\nu}) and ϵ>0\epsilon>0, setting Xϵ . . =Xπ(Zν(ϵ))X^{\epsilon}\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=X\cap\pi(Z^{\star}_{\nu}(\epsilon)), we define the quantity VDνϵ(Xϵ)\operatorname{VD}^{\epsilon}_{\nu}(X^{\epsilon}) as

VDνϵ(Xϵ) . . =sup{α:Λαϵ(Xϵ)=}=inf{α:Λαϵ(Xϵ)=0},\operatorname{VD}^{\epsilon}_{\nu}(X^{\epsilon})\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\sup\left\{\alpha:\Lambda^{\epsilon}_{\alpha}(X^{\epsilon})=\infty\right\}=\inf\left\{\alpha:\Lambda^{\epsilon}_{\alpha}(X^{\epsilon})=0\right\},

where

Λαϵ(Xϵ) . . =limκ0limNinf{Ui}i1Vol(Ui)α.\Lambda^{\epsilon}_{\alpha}(X^{\epsilon})\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\lim_{\kappa\to 0}\lim_{N^{\star}\to\infty}\inf_{\{U_{i}\}}\sum_{i\geq 1}\operatorname{Vol}(U_{i})^{\alpha}.

Here the infimum is taken over the covers consisting of sets UiU_{i} of the form Ui=U(Ni,x,κ,ϵ)U_{i}=U(N_{i},x,\kappa,\epsilon), for some xπ(Zν(ϵ))x\in\pi(Z_{\nu}(\epsilon)) and NiNN_{i}\geq N^{\star}. The volume dimensions of XX and ν\nu are then respectively defined as

VDν(X) . . =lim supϵ0VDνϵ(Xϵ) and VD(ν) . . =inf{VDν(X):Xπ(Zν),ν(X)=1},\operatorname{VD}_{\nu}(X)\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\limsup_{\epsilon\to 0}\operatorname{VD}_{\nu}^{\epsilon}(X^{\epsilon})\quad\mbox{ and }\quad\operatorname{VD}(\nu)\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\inf\left\{\operatorname{VD}_{\nu}(X)\colon X\subseteq\pi(Z_{\nu}),\nu(X)=1\right\},

and the lim supϵ0\limsup_{\epsilon\to 0} is actually a limit; see Section 4.2. We prove in Proposition 4.26 a version of Young’s criterion [You82, Proposition 2.1], relating the local volume dimensions δx\delta_{x} with the volume dimensions VDν(X)\operatorname{VD}_{\nu}(X) and VD(ν)\operatorname{VD}(\nu). This corresponds to Step (2) above and, together with the exact volume-dimensionality of ν\nu proved in the first step, completes the proof of Theorem 1.1.

1.3. Organization of the paper

The paper is organized as follows. In Section 2, we derive from the distortion theorem [BDM08, BD19] the estimates that we will need, and we introduce the volume-conformal measures and the pressure function tPJ+(t)t\mapsto P_{J}^{+}(t). We prove the exact dimensionality of every ν+(f)\nu\in\mathcal{M}^{+}(f) in Section 3. In Section 4, we define and study the volume dimensions of sets and measures. We conclude the proof of Theorem 1.1 and prove Theorems 1.2 and 1.3 in Section 5.

Acknowledgements

The authors would like to thank the University of Lille and the University of Oklahoma for the warm welcome and for the excellent work conditions.

This project has received funding from the French government through the Programme Investissement d’Avenir (LabEx CEMPI ANR-11-LABX-0007-01, ANR QuaSiDy ANR-21-CE40-0016, ANR PADAWAN ANR-21-CE40-0012-01) managed by the Agence Nationale de la Recherche.

2. Definitions and preliminary results

After fixing some notations in Section 2.1, in Section 2.2 we recall the distortion theorem by Berteloot-Dupont-Molino [BDM08, BD19] and deduce the estimates which will be essential ingredients in the proof of Theorem 1.1. We define and study basic properties about volume-conformal measures in Section 2.3, and a pressure function in Section 2.4.

2.1. Notations

Let f:kkf:\mathbb{P}^{k}\to\mathbb{P}^{k} be a holomorphic endomorphism and ν\nu an ergodic ff-invariant probability measure. By Oseledets’ theorem [Ose68], one can associate to ν\nu its Lyapunov exponents χmin . . =χl<<χ1\chi_{\min}\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\chi_{l}<\ldots<\chi_{1}, where 1lk1\leq l\leq k. For ν\nu-almost every xkx\in\mathbb{P}^{k}, there exists a stratification in complex linear subspaces {0}=:(Ll+1)x(Ll)x(L1)x=Txk\{0\}=:{(L_{l+1})_{x}}\subset{(L_{l})_{x}}\subset\ldots\subset(L_{1})_{x}=T_{x}\mathbb{P}^{k} of the complex tangent space TxkT_{x}\mathbb{P}^{k} such that Dfx(Lj)x=(Lj)f(x)Df_{x}(L_{j})_{x}=(L_{j})_{f(x)} and limnn1logDfxnv=χj\lim_{n\to\infty}n^{-1}\log||Df_{x}^{n}v||=\chi_{j} for all v(Lj)x(Lj+1)xv\in(L_{j})_{x}\setminus(L_{j+1})_{x} for all 1jl1\leq j\leq l.

Let us first assume that all the χj\chi_{j}’s are distinct, i.e., that we have l=kl=k and χmin=χk<<χ1\chi_{\min}=\chi_{k}<\ldots<\chi_{1}. Then, (Lj)x(L_{j})_{x} has dimension kj+1k-j+1 for all 1jk1\leq j\leq k. We denote by 𝒪\mathcal{O} a full measure subset of the support of ν\nu given by Oseledets’ theorem. Take x𝒪x\in\mathcal{O}. Fix a basis (j)x(\ell_{j})_{x} of the complex tangent space TxkT_{x}\mathbb{P}^{k} with the property that (Lj)x(L_{j})_{x} is equal to the span of {(j)x,,(k)x}\{(\ell_{j})_{x},\dots,(\ell_{k})_{x}\}. Denote by {ej}j=1k\{e_{j}\}_{j=1}^{k} the standard basis of kk\mathbb{R}^{k}\subset\mathbb{C}^{k}. For every r1,,rkr_{1},\dots,r_{k}\in\mathbb{R} (sufficiently small), we denote by x(r1,,rk)\mathcal{E}_{x}(r_{1},\dots,r_{k}) the image of the unit ball 𝔹kk\mathbb{B}^{k}\subset\mathbb{C}^{k} (in a given local chart at xx) under the composition eΦ:kke\circ\Phi:\mathbb{C}^{k}\to\mathbb{P}^{k}, where e:Txkke\colon T_{x}\mathbb{P}^{k}\to\mathbb{P}^{k} is the standard exponential map and Φ:kTxkk\Phi\colon\mathbb{C}^{k}\to T_{x}\mathbb{P}^{k}\simeq\mathbb{C}^{k} is a linear map such that Φ((ej)x)=rj(j)x\Phi\big{(}(e_{j})_{x}\big{)}=r_{j}(\ell_{j})_{x}.

If, for all 1jk1\leq j\leq k, the argument rjr_{j} of x\mathcal{E}_{x} is a function ϕ(j)\phi(j) depending on jj, we write

x(ϕ(j)) . . =x(ϕ(1),,ϕ(k))\mathcal{E}_{x}(\phi(j))\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\mathcal{E}_{x}(\phi(1),\dots,\phi(k))

for brevity. In particular, we will often have χmin>0\chi_{\min}>0 and take ϕ(j)\phi(j) of the form ϕ(j)=c1en(χj±c2ϵ)\phi(j)=c_{1}e^{-n(\chi_{j}\pm c_{2}\epsilon)} for some nn\in\mathbb{N}, 0<ϵχmin0<\epsilon\ll\chi_{\min} sufficiently small, and some positive constants c1c_{1} (independent of jj and nn) and c2c_{2} (independent of jj, nn, and ϵ\epsilon). We will call the sets x\mathcal{E}_{x} dynamical ellipses in this case.

If l<kl<k, i.e., some Lyapunov exponent χj\chi_{j} has multiplicity larger than 1, the above construction generalizes by taking into account the corresponding rjr_{j} with the same multiplicity. Namely, we assign the same rjr_{j} to all the directions associated to the same Lyapunov exponent which has multiplicity larger than 1.

Let XkX\subseteq\mathbb{P}^{k} be a closed invariant set for ff. We denote by X+(f)\mathcal{M}_{X}^{+}(f) the set of all ergodic ff-invariant measures supported on XX with strictly positive Lyapunov exponents. We drop the index XX if X=kX=\mathbb{P}^{k}. We say that XX is uniformly expanding if there exist η>1\eta>1 and C>0C>0 such that Dfxn(v)>Cηnv||Df_{x}^{n}(v)||>C\eta^{n}||v|| for every xXx\in X, vTxkv\in T_{x}\mathbb{P}^{k}, and nn\in\mathbb{N}. We say that ff is hyperbolic if J(f)J(f) is uniformly expanding.

We will consider the Fubini-Study metric on k\mathbb{P}^{k}. We will denote by dist\operatorname{dist} the corresponding distance, and by B(x,r)B(x,r) the open ball centred at xx and of radius rr. For an open set VkV\subset\mathbb{P}^{k}, we denote by Vol(V)\operatorname{Vol}(V) the volume of VV with respect to the Fubini-Study metric. Given a holomorphic map g:Vkg\colon V\to\mathbb{P}^{k}, we denote by Jacg(x)\operatorname{Jac}g(x) the Jacobian of gg at xVx\in V, i.e., the determinant of the differential DgxDg_{x}.

We also fix the positive constant M>0M>0 defined as

(2.1) M . . =logsupxksupvkDfx(v)vM\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\log\sup_{x\in\mathbb{P}^{k}}\sup_{v\in\mathbb{C}^{k}}\frac{||Df_{x}(v)||}{||v||}

and observe that eMe^{M} dominates the Lipschitz constant of ff. In particular, we have dist(f(x1),f(x2))eMdist(x1,x2)\operatorname{dist}(f(x_{1}),f(x_{2}))\leq e^{M}\operatorname{dist}(x_{1},x_{2}) for every x1,x2kx_{1},x_{2}\in\mathbb{P}^{k}, and f(B(x,r))B(f(x),eMr)f(B(x,r))\subseteq B(f(x),e^{M}r) for every xkx\in\mathbb{P}^{k} and r>0r>0.

2.2. Distortion estimates along generic inverse branches

We fix in this subsection a holomorphic endomorphism f:kkf:\mathbb{P}^{k}\to\mathbb{P}^{k} of algebraic degree d2d\geq 2 and a measure ν+(f)\nu\in\mathcal{M}^{+}(f). All the objects and the constants that we introduce in this subsection depend on ff and ν\nu. We denote by χ1>>χl=χmin>0\chi_{1}>\ldots>\chi_{l}=\chi_{\min}>0 the (distinct) Lyapunov exponents of ν\nu, by k1,,klk_{1},\dots,k_{l} their respective multiplicities, and by Lν=Lν(f) . . =j=1lkjχjL_{\nu}=L_{\nu}(f)\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\sum_{j=1}^{l}k_{j}\chi_{j} their sum. Recall that we have Lν=log|Jacf|dνL_{\nu}=\int\log|\operatorname{Jac}f|d\nu by Birkhoff’s ergodic theorem.

Consider the orbit space of ff

O . . ={x^={xn}n(k):xn+1=f(xn)n}O\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\left\{\hat{x}=\{x_{n}\}_{n\in\mathbb{Z}}\in(\mathbb{P}^{k})^{\mathbb{Z}}\colon x_{n+1}=f(x_{n})\quad\forall n\in\mathbb{Z}\right\}

and the right shift map T:OOT\colon O\to O defined as T(x^)={xn+1}nT(\hat{x})=\{x_{n+1}\}_{n\in\mathbb{Z}} for x^={xn}n\hat{x}=\{x_{n}\}_{n\in\mathbb{Z}}. Given η>0\eta>0, a function ϕ:O(0,1]\phi\colon O\to(0,1] is said to be η\eta-slow if for any x^O\hat{x}\in O we have

eηϕ(x^)ϕ(T(x^))eηϕ(x^).e^{-\eta}\phi(\hat{x})\leq\phi(T(\hat{x}))\leq e^{\eta}\phi(\hat{x}).

We now recall the construction of the lift ν^\hat{\nu} of ν\nu to OO; see [CFS12, Section 10.4] and [PU10, Section 2.7]. For nn\in\mathbb{Z}, we let πn:Ok\pi_{n}\colon O\to\mathbb{P}^{k} be the projection map defined by πn(x^)=xn\pi_{n}(\hat{x})=x_{n}, where x^={xn}n\hat{x}=\{x_{n}\}_{n\in\mathbb{Z}}. We write π . . =π0\pi\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\pi_{0} for brevity. Observe that πnT=fπn\pi_{n}\circ T=f\circ\pi_{n} for all nn\in\mathbb{Z}.

Consider the σ\sigma-algebra ^\hat{\mathcal{B}} on OO generated by the sets of the form

An,B . . =πn1(B)={x^:xnB}A_{n,B}\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\pi_{n}^{-1}(B)=\left\{\hat{x}\colon x_{n}\in B\right\}

with n0n\leq 0 and BkB\subseteq\mathbb{P}^{k} a Borel set. For all such sets An,BA_{n,B}, set

ν^(An,B) . . =ν(B).\hat{\nu}(A_{n,B})\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\nu(B).

Then, by the invariance of ν\nu and the fact that xnBx_{n}\in B if and only if xnmfm(B)x_{n-m}\in f^{-m}(B) with m0m\geq 0, we see that ν^\hat{\nu} is well-defined on the sets An,BA_{n,B} as it satisfies ν^(An,B)=ν^(Anm,B)\hat{\nu}(A_{n,B})=\hat{\nu}(A_{n-m,B}) for all m0m\geq 0. Similarly, for all m0m\geq 0 and Borel sets B0,,BmkB_{0},\dots,B_{-m}\subseteq\mathbb{P}^{k}, we have

ν^({x^:x0B0,,xmBm})=ν(fm(B0)fm+1(B1)Bm).\hat{\nu}\big{(}\{\hat{x}\colon x_{0}\in B_{0},\ldots,x_{-m}\in B_{-m}\}\big{)}=\nu\left(f^{-m}(B_{0})\cap f^{-m+1}(B_{1})\cap\ldots\cap B_{-m}\right).

We can then extend ν^\hat{\nu} to a probability measure on ^\hat{\mathcal{B}}, that we still denote by ν^\hat{\nu}. By construction, ν^\hat{\nu} is TT-invariant and satisfies π(ν^)=ν\pi_{*}(\hat{\nu})=\nu. As ν\nu is ergodic, one can prove that ν^\hat{\nu} is also ergodic.

Recall that the critical set C(f)C(f) of ff is the set of points xkx\in\mathbb{P}^{k} at which the differential DfxDf_{x} is not invertible. As all the Lyapunov exponents of ν\nu are finite, and their sum is equal to log|Df|dν\int\log|Df|d\nu, we have in particular ν(C(f))=0\nu(C(f))=0. Set

Z . . ={x^O:xnC(f)n}.Z\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\big{\{}\hat{x}\in O:x_{n}\notin C(f)\quad\forall n\in\mathbb{Z}\big{\}}.

Then the set ZZ is TT-invariant and satisfies ν^(Z)=1\hat{\nu}(Z)=1. For every x^Z\hat{x}\in Z, we denote by fx^nf_{\hat{x}}^{-n} the inverse branch of fnf^{n} defined in a neighbourhood of x0x_{0} and such that fx^n(x0)=xnf^{-n}_{\hat{x}}(x_{0})=x_{-n}.

The following result is stated in [BD19, Theorem A] (see also [BDM08, Theorem 1.4]) in the case where ν\nu is the measure of maximal entropy of ff. The same statement and proof hold for any measure in ν+(f)\nu\in\mathcal{M}^{+}(f), as stated at the end of the Introduction – and used in later sections – of the same paper.

Theorem 2.1.

For every 0<2η<γχmin0<2\eta<\gamma\ll\chi_{\min} and ν^\hat{\nu}-almost every x^Z\hat{x}\in Z, there exist

  1. (1)

    an integer nx^1n_{\hat{x}}\geq 1 and real numbers hx^1h_{\hat{x}}\geq 1 and 0<rx^,ρx^10<r_{\hat{x}},\rho_{\hat{x}}\leq 1,

  2. (2)

    a sequence {φx^,n}n0\{\varphi_{\hat{x},n}\}_{n\geq 0} of injective holomorphic maps

    φx^,n:B(xn,rx^en(γ+2η))𝔻k(ρx^enη)\varphi_{\hat{x},n}:B(x_{-n},r_{\hat{x}}e^{-n(\gamma+2\eta)})\to\mathbb{D}^{k}(\rho_{\hat{x}}e^{n\eta})

    sending xnx_{-n} to 0 and satisfying

    en(γ2η)dist(u,v)|φx^,n(u)φx^,n(v)|en(γ+3η)hx^dist(u,v)e^{n(\gamma-2\eta)}\operatorname{dist}(u,v)\leq|\varphi_{\hat{x},n}(u)-\varphi_{\hat{x},n}(v)|\leq e^{n(\gamma+3\eta)}h_{\hat{x}}\operatorname{dist}(u,v)

    for every nn\in\mathbb{N} and u,vB(xn,rx^en(γ+2η))u,v\in B(x_{-n},r_{\hat{x}}e^{-n(\gamma+2\eta)});

  3. (3)

    a sequence {x^,n}n0\{\mathcal{L}_{\hat{x},n}\}_{n\geq 0} of linear maps from k\mathbb{C}^{k} to k\mathbb{C}^{k} which stabilize each

    Hj . . ={0}××kj××{0},H_{j}\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\{0\}\times\ldots\times\mathbb{C}^{k_{j}}\times\ldots\times\{0\},

    satisfy

    enχj+n(γη)|v||x^,n(v)|enχj+n(γ+η)|v|for all n and vHj,e^{-n\chi_{j}+n(\gamma-\eta)}|v|\leq|\mathcal{L}_{\hat{x},n}(v)|\leq e^{-n\chi_{j}+n(\gamma+\eta)}|v|\quad\mbox{for all }n\in\mathbb{N}\mbox{ and }v\in H_{j},

    and such that the diagram

    B(x0,rx^){B(x_{0},r_{\hat{x}})}B(xn,rx^en(γ+2η)){B(x_{-n},r_{\hat{x}}e^{-n(\gamma+2\eta)})}𝔻k(ρx^){\mathbb{D}^{k}(\rho_{\hat{x}})}𝔻k(ρx^enη){\mathbb{D}^{k}(\rho_{\hat{x}}e^{n\eta})}fx^n\scriptstyle{f_{\hat{x}}^{-n}}φx^,0\scriptstyle{\varphi_{\hat{x},0}}φx^,n\scriptstyle{\varphi_{\hat{x},n}}x^,n\scriptstyle{\mathcal{L}_{\hat{x},n}}

    commutes for all nnx^n\geq n_{\hat{x}}.

Moreover, the functions x^hx^1,rx^,ρx^\hat{x}\mapsto h_{\hat{x}}^{-1},r_{\hat{x}},\rho_{\hat{x}} are measurable and η\eta-slow on ZZ.

In particular, for every nn\in\mathbb{N} and x^\hat{x} as in the statement, the inverse branch fx^nf_{\hat{x}}^{-n} is well-defined on the ball B(x0,rx^)B(x_{0},r_{\hat{x}}).

Corollary 2.2.

With the same assumptions and notations as in Theorem 2.1 and Section 2.1, for ν^\hat{\nu}-almost all x^Z\hat{x}\in Z and all t,t1,,tk(0,1]t,t_{1},\dots,t_{k}\in(0,1], nnx^n\geq n_{\hat{x}}, and y,wB(x0,rx^)y,w\in B(x_{0},r_{\hat{x}}), we have

  1. (1)

    en(Lν+10kη)|Jacfx^n(y)|en(Lν10kη)e^{-n(L_{\nu}+10k\eta)}\leq|\operatorname{Jac}f^{-n}_{\hat{x}}(y)|\leq e^{-n(L_{\nu}-10k\eta)};

  2. (2)

    e20knη|Jacfx^n(y)||Jacfx^n(w)|1e20knηe^{-20kn\eta}\leq|\operatorname{Jac}f^{-n}_{\hat{x}}(y)|\cdot|\operatorname{Jac}f^{-n}_{\hat{x}}(w)|^{-1}\leq e^{20kn\eta};

  3. (3)

    xn(tjrx^hx^1en(χj+10η))fx^n(x0(tjrx^))xn(tjrx^hx^en(χj10η))\mathcal{E}_{x_{-n}}(t_{j}r_{\hat{x}}h^{-1}_{\hat{x}}e^{-n(\chi_{j}+10\eta)})\subset f^{-n}_{\hat{x}}(\mathcal{E}_{x_{0}}(t_{j}r_{\hat{x}}))\subset\mathcal{E}_{x_{-n}}(t_{j}r_{\hat{x}}h_{\hat{x}}e^{-n(\chi_{j}-10\eta)});

  4. (4)

    (trx^)2ke2n(Lν+10kη)Vol(xn(trx^enχj))(trx^)2ke2n(Lν10kη)(tr_{\hat{x}})^{2k}e^{-2n(L_{\nu}+10k\eta)}\leq\operatorname{Vol}(\mathcal{E}_{x_{-n}}(tr_{\hat{x}}e^{-n\chi_{j}}))\leq(tr_{\hat{x}})^{2k}e^{-2n(L_{\nu}-10k\eta)};

  5. (5)

    (trx^hx^1)2ke2n(Lν+20kη)Vol(fx^n(B(x0,trx^)))(trx^hx^)2ke2n(Lν20kη)(tr_{\hat{x}}h_{\hat{x}}^{-1})^{2k}e^{-2n(L_{\nu}+20k\eta)}\leq\operatorname{Vol}(f^{-n}_{\hat{x}}(B({x_{0}},tr_{\hat{x}})))\leq(tr_{\hat{x}}h_{\hat{x}})^{2k}e^{-2n(L_{\nu}-20k\eta)}.

Proof.

The assertions (1) and (3) follow directly from Theorem 2.1 (2) and (3). The assertion (2) follows from (1). The assertion (4) follows from the fact that the distances in (Txnk)k1\mathbb{P}(T_{x_{-n}}\mathbb{P}^{k})\simeq\mathbb{P}^{k-1} between the directions associated to distinct Lyapunov exponents at xnx_{-n} are larger (up to a multiplicative constant independent of nn) than e5nηe^{-5n\eta}, again by Theorem 2.1 (2). This allows one to compare the volume of xn(trx^enχj)\mathcal{E}_{x_{-n}}(tr_{\hat{x}}e^{-n\chi_{j}}) with that of an ellipse in k\mathbb{C}^{k}, whose axes are parallel to the coordinate planes. The assertion (5) is a consequence of (3), applied with tj=tt_{j}=t for all j=1,,kj=1,\ldots,k, and (4). ∎

Definition 2.3.

We define ZνZZ_{\nu}\subseteq Z to be the full ν^\hat{\nu}-measure set of elements x^Z\hat{x}\in Z satisfying the conditions in Theorem 2.1 and Corollary 2.2. For α>0\alpha>0, we also define

Zν,α . . ={x^Zν:nx^<α1,rx^>α,hx^<α1}.Z_{\nu,\alpha}\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\big{\{}\hat{x}\in Z_{\nu}\colon n_{\hat{x}}<\alpha^{-1},r_{\hat{x}}>\alpha,h_{\hat{x}}<\alpha^{-1}\big{\}}.

It follows from the definition that, as α0\alpha\to 0, the sets Zν,αZ_{\nu,\alpha} increase to ZνZ_{\nu}. In particular, we have ν^(Zν,α)1\hat{\nu}(Z_{\nu,\alpha})\to 1 as α0\alpha\to 0.

Corollary 2.4.

For every 0<ϵχmin0<\epsilon\ll\chi_{\min} sufficiently small, there exist Zν(ϵ)ZνZ^{\prime}_{\nu}(\epsilon)\subseteq Z_{\nu}, n(ϵ)1n^{\prime}(\epsilon)\geq 1, and r(ϵ)(0,1)r(\epsilon)\in(0,1) such that

  1. (1)

    ν^(Z(ϵ))>1ϵ\hat{\nu}(Z(\epsilon))>1-\epsilon;

  2. (2)

    nx^n(ϵ)n_{\hat{x}}\leq n(\epsilon) and rx^r(ϵ)r_{\hat{x}}\geq r(\epsilon) for all x^Zν(ϵ)\hat{x}\in Z^{\prime}_{\nu}(\epsilon);

  3. (3)

    for all t,t1,,tk(0,1]t,t_{1},\ldots,t_{k}\in(0,1], nn(ϵ)n\geq n(\epsilon), x^Zν(ϵ)\hat{x}\in Z^{\prime}_{\nu}(\epsilon), and y,wB(x0,r(ϵ))y,w\in B(x_{0},r(\epsilon)) we have

    1. (a)

      en(Lν+kϵ)|Jacfx^n(y)|en(Lνkϵ)e^{-n(L_{\nu}+k\epsilon)}\leq|\operatorname{Jac}f^{-n}_{\hat{x}}(y)|\leq e^{-n(L_{\nu}-k\epsilon)};

    2. (b)

      eknϵ|Jacfx^n(y)||Jacfx^n(w)|1eknϵe^{-kn\epsilon}\leq|\operatorname{Jac}f^{-n}_{\hat{x}}(y)|\cdot|\operatorname{Jac}f^{-n}_{\hat{x}}(w)|^{-1}\leq e^{kn\epsilon};

    3. (c)

      xn(tjr(ϵ)en(χj+ϵ))fx^n(x0(tjr(ϵ)))xn(tjr(ϵ)en(χjϵ))\mathcal{E}_{x_{-n}}(t_{j}r(\epsilon)e^{-n(\chi_{j}+\epsilon)})\subset f^{-n}_{\hat{x}}(\mathcal{E}_{x_{0}}(t_{j}r(\epsilon)))\subset\mathcal{E}_{x_{-n}}(t_{j}r(\epsilon)e^{-n(\chi_{j}-\epsilon)});

    4. (d)

      (tr(ϵ))2ke2n(Lν+kϵ)Vol(xn(tr(ϵ)enχj))(tr(ϵ))2ke2n(Lνkϵ)(tr(\epsilon))^{2k}e^{-2n(L_{\nu}+k\epsilon)}\leq\operatorname{Vol}(\mathcal{E}_{x_{-n}}(tr(\epsilon)e^{-n\chi_{j}}))\leq(tr(\epsilon))^{2k}e^{-2n(L_{\nu}-k\epsilon)};

    5. (e)

      (tr(ϵ))2ke2n(Lν+kϵ)Vol(fx^n(B(x0,tr(ϵ))))(tr(ϵ))2ke2n(Lνkϵ)(tr(\epsilon))^{2k}e^{-2n(L_{\nu}+k\epsilon)}\leq\operatorname{Vol}(f^{-n}_{\hat{x}}(B({x_{0}},tr(\epsilon))))\leq(tr(\epsilon))^{2k}e^{-2n(L_{\nu}-k\epsilon)},

where nx^n_{\hat{x}}, hx^h_{\hat{x}}, and rx^r_{\hat{x}} are as in Theorem 2.1.

Proof.

By choosing α=α(ϵ)\alpha=\alpha(\epsilon) sufficiently small, Corollary 2.2 and the Definition 2.3 of Zν,αZ_{\nu,\alpha} give the existence of a set Zν′′(ϵ) . . =Zν,α(ϵ)Z^{\prime\prime}_{\nu}(\epsilon)\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=Z_{\nu,\alpha(\epsilon)} and numbers r(ϵ)r(\epsilon), n(ϵ)n^{\prime}(\epsilon) satisfying the properties in the statement, with (3c) and (3e) replaced by

xn(tjr(ϵ)α(ϵ)en(χj+ϵ/2))fx^n(x0(tjr(ϵ)))xn(tjr(ϵ)α(ϵ)1en(χjϵ/2))\mathcal{E}_{x_{-n}}(t_{j}r(\epsilon)\alpha(\epsilon)e^{-n(\chi_{j}+\epsilon/2)})\subset f^{-n}_{\hat{x}}(\mathcal{E}_{x_{0}}(t_{j}r(\epsilon)))\subset\mathcal{E}_{x_{-n}}(t_{j}r(\epsilon)\alpha(\epsilon)^{-1}e^{-n(\chi_{j}-\epsilon/2)})

and

(tr(ϵ)α(ϵ))2ke2n(Lν+kϵ/2)Vol(fx^n(B(x0,tr(ϵ))))(tr(ϵ)α(ϵ)1)2ke2n(Lνkϵ/2),(tr(\epsilon)\alpha(\epsilon))^{2k}e^{-2n(L_{\nu}+k\epsilon/2)}\leq\operatorname{Vol}(f^{-n}_{\hat{x}}(B({x_{0}},tr(\epsilon))))\leq(tr(\epsilon)\alpha(\epsilon)^{-1})^{2k}e^{-2n(L_{\nu}-k\epsilon/2)},

respectively. Since all the Lyapunov exponents of ν\nu are strictly positive, the assertion follows up to increasing n(ϵ)n^{\prime}(\epsilon). ∎

Lemma 2.5.

For every 0<ϵχmin0<\epsilon\ll\chi_{\min} sufficiently small, there exist n(ϵ)n(\epsilon)\in\mathbb{N}, a subset Zν(ϵ)Zν(ϵ)Z_{\nu}(\epsilon)\subseteq Z^{\prime}_{\nu}(\epsilon) with ν^(Zν(ϵ)Zν(ϵ))<ϵ\hat{\nu}(Z_{\nu}^{\prime}(\epsilon)\setminus Z_{\nu}(\epsilon))<\epsilon, and, for all x^Zν(ϵ)\hat{x}\in Z_{\nu}(\epsilon), a sequence {nl}l0={nl(x^)}l0\{n_{l}\}_{l\geq 0}=\{n_{l}(\hat{x})\}_{l\geq 0} such that

  1. (1)

    n(ϵ)n0n(ϵ)n^{\prime}(\epsilon)\leq n_{0}\leq n(\epsilon);

  2. (2)

    nl+1nl<ϵnln_{l+1}-n_{l}<\epsilon n_{l} for all l0l\geq 0;

  3. (3)

    Tnl(x^)Zν(ϵ)T^{n_{l}}(\hat{x})\in Z^{\prime}_{\nu}(\epsilon) for all l0l\geq 0,

where Zν(ϵ)Z^{\prime}_{\nu}(\epsilon) and n(ϵ)n^{\prime}(\epsilon) are as in Corollary 2.4.

A version of Lemma 2.5 is essentially proved in [PU10, Section 11.4] in the case of k=1k=1. We will need here to further get a uniform upper bound for the element n0n_{0} associated to any x^Zν(ϵ)\hat{x}\in Z_{\nu}(\epsilon).

Proof.

We first show the existence of a set Zν′′′(ϵ)Zν(ϵ)Z^{\prime\prime\prime}_{\nu}(\epsilon)\subseteq Z^{\prime}_{\nu}(\epsilon) with ν^(Zν(ϵ)Zν′′′(ϵ))=0\hat{\nu}(Z^{\prime}_{\nu}(\epsilon)\setminus Z^{\prime\prime\prime}_{\nu}(\epsilon))=0 and, for any x^Zν′′′(ϵ)\hat{x}\in Z^{\prime\prime\prime}_{\nu}(\epsilon), of a sequence {nl}l0={nl(x^)}l0\{n_{l}\}_{l\geq 0}=\{n_{l}(\hat{x})\}_{l\geq 0} satisfying (2), (3), and n0n(ϵ)n_{0}\geq n^{\prime}(\epsilon).

For every nn\in\mathbb{N} and x^Z\hat{x}\in Z, set Sn(x^) . . =j=1n1𝟏Zν(ϵ)Tj(x^)S_{n}(\hat{x})\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\sum_{j=1}^{n-1}{\bf 1}_{Z^{\prime}_{\nu}(\epsilon)}\circ T^{j}(\hat{x}), where 𝟏V{\bf 1}_{V} denotes the characteristic function of VkV\subset\mathbb{P}^{k}. Since ν^\hat{\nu} is ergodic, by Birkhoff’s ergodic theorem there exists a measurable set Zν′′′(ϵ)Zν(ϵ)Z^{\prime\prime\prime}_{\nu}(\epsilon)\subseteq Z^{\prime}_{\nu}(\epsilon) such that ν^(Zν′′′(ϵ))=ν^(Zν(ϵ))\hat{\nu}(Z^{\prime\prime\prime}_{\nu}(\epsilon))=\hat{\nu}(Z^{\prime}_{\nu}(\epsilon)) and

(2.2) limnn1Sn(x^)=ν^(Zν(ϵ)) for every x^Zν′′′(ϵ).\lim_{n\to\infty}n^{-1}S_{n}(\hat{x})=\hat{\nu}(Z^{\prime}_{\nu}(\epsilon))\quad\mbox{ for every }\hat{x}\in Z^{\prime\prime\prime}_{\nu}(\epsilon).

Take x^Zν′′′(ϵ)\hat{x}\in Z^{\prime\prime\prime}_{\nu}(\epsilon). By (2.2) and the fact that ν^(Zν(ϵ))>1ϵ\hat{\nu}(Z^{\prime}_{\nu}(\epsilon))>1-\epsilon, there exists n=n(x^)>10/ϵn^{\star}=n^{\star}(\hat{x})>10/\epsilon such that |n1Sn(x^)ν^(Zν(ϵ))|ϵ/20|n^{-1}S_{n}(\hat{x})-\hat{\nu}(Z^{\prime}_{\nu}(\epsilon))|\leq\epsilon/20 for all nnn\geq n^{\star}. We define {nl}l0={nl(x^)}l0\{n_{l}\}_{l\geq 0}=\{n_{l}(\hat{x})\}_{l\geq 0} to be the sequence of integers nmax{n,n(ϵ)}n\geq\max\{n^{\star},n^{\prime}(\epsilon)\} such that Tn(x^)Zν(ϵ)T^{n}(\hat{x})\in Z^{\prime}_{\nu}(\epsilon).

It follows from the definitions of Sn(x^)S_{n}(\hat{x}) and of the sequence {nl}l0\{n_{l}\}_{l\geq 0} that Snl+1(x^)=Snl(x^)+1S_{n_{l+1}}(\hat{x})=S_{n_{l}}(\hat{x})+1 for all l0l\geq 0. Moreover, we also have

Snl(x^)nl(ν^(Zν(ϵ))+ϵ/20) and Snl+1(x^)nl+1(ν^(Zν(ϵ))ϵ/20) for all l0.S_{n_{l}}(\hat{x})\leq n_{l}\big{(}\hat{\nu}(Z^{\prime}_{\nu}(\epsilon))+\epsilon/20\big{)}\quad\mbox{ and }\quad S_{n_{l+1}}(\hat{x})\geq n_{l+1}\big{(}\hat{\nu}(Z^{\prime}_{\nu}(\epsilon))-\epsilon/20\big{)}\quad\mbox{ for all }l\geq 0.

We deduce from these inequalities that, again for all l0l\geq 0,

nl+1nlSnl+1(x^)nl(ν^(Zν(ϵ))ϵ/20)=Snl(x^)+1nl(ν^(Zν(ϵ))ϵ/20)ν^(Zν(ϵ))+ϵ/20+1/nlν^(Zν(ϵ))ϵ/20.\frac{n_{l+1}}{n_{l}}\leq\frac{S_{n_{l+1}}(\hat{x})}{n_{l}\big{(}\hat{\nu}(Z^{\prime}_{\nu}(\epsilon))-\epsilon/20\big{)}}\\ =\frac{S_{n_{l}}(\hat{x})+1}{n_{l}\big{(}\hat{\nu}(Z^{\prime}_{\nu}(\epsilon))-\epsilon/20\big{)}}\leq\frac{\hat{\nu}(Z^{\prime}_{\nu}(\epsilon))+\epsilon/20+1/n_{l}}{\hat{\nu}(Z^{\prime}_{\nu}(\epsilon))-\epsilon/20}.

Hence, since 0<ϵχmin0<\epsilon\ll\chi_{\min} and nln>10/ϵn_{l}\geq n^{\star}>10/\epsilon for all l0l\geq 0, we have

nl+1nlnlν^(Zν(ϵ))+ϵ/20+1/nlν^(Zν(ϵ))ϵ/201ϵ/10+ϵ/101ϵϵ/20<ϵ.\frac{n_{l+1}-n_{l}}{n_{l}}\leq\frac{\hat{\nu}(Z^{\prime}_{\nu}(\epsilon))+\epsilon/20+1/n_{l}}{\hat{\nu}(Z^{\prime}_{\nu}(\epsilon))-\epsilon/20}-1\leq\frac{\epsilon/10+\epsilon/10}{1-\epsilon-\epsilon/20}<\epsilon.

This gives the existence of a set Zν′′′(ϵ)Z^{\prime\prime\prime}_{\nu}(\epsilon) with the properties stated at the beginning of the proof.

For every N>n(ϵ)N>n^{\prime}(\epsilon), set ZνN(ϵ) . . ={x^Zν′′′(ϵ):n0(x^)N}Z^{N}_{\nu}(\epsilon)\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\{\hat{x}\in Z^{\prime\prime\prime}_{\nu}(\epsilon)\colon n_{0}(\hat{x})\leq N\}. The sequence of sets ZνN(ϵ)Z^{N}_{\nu}(\epsilon) is non-decreasing as NN\to\infty, and satisfies NZνN(ϵ)=Zν′′′(ϵ)\cup_{N}Z^{N}_{\nu}(\epsilon)=Z^{\prime\prime\prime}_{\nu}(\epsilon). Fix m=m(ϵ)m^{\star}=m^{\star}(\epsilon) such that ν^(Zν′′′(ϵ)Zνm(ϵ))<ϵ\hat{\nu}(Z^{\prime\prime\prime}_{\nu}(\epsilon)\setminus Z^{m^{\star}}_{\nu}(\epsilon))<\epsilon. The assertion follows setting Zν(ϵ) . . =Zνm(ϵ)Z_{\nu}(\epsilon)\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=Z^{m^{\star}}_{\nu}(\epsilon) and n(ϵ) . . =mn(\epsilon)\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=m^{\star}. ∎

Recall that π:Ok\pi\colon O\to\mathbb{P}^{k} denotes the projection map defined by π(x^)=x0\pi(\hat{x})=x_{0}, where x^={xn}n\hat{x}=\{x_{n}\}_{n\in\mathbb{Z}}.

Remark 2.6.

Observe that the sequence {nl}l0={nl(x^)}l0\{n_{l}\}_{l\geq 0}=\{n_{l}(\hat{x})\}_{l\geq 0} as in Lemma 2.5 only depends on x0=π(x^)x_{0}=\pi(\hat{x}). In particular, the sequence {nl}l0\{n_{l}\}_{l\geq 0} as in Lemma 2.5 is well-defined for every xπ(Zν(ϵ))x\in\pi(Z_{\nu}(\epsilon)).

Definition 2.7.

Given N0N\geq 0, xkx\in\mathbb{P}^{k}, κ>0\kappa>0, and ϵ>0\epsilon>0, we denote by U=U(N,x,κ,ϵ)U=U(N,x,\kappa,\epsilon) the (necessarily unique) set UU, if it exists, satisfying fN(U)=B(fN(x),κeNMϵ)f^{N}(U)=B(f^{N}(x),\kappa\,e^{-NM\epsilon}) (where MM is as in (2.1)) and such that fN|Uf^{N}|_{U} is injective. We call z(U)z(U) the center of UU.

By Definition 2.7, we have U(N,x0,κ,ϵ) . . =fTN(z^)N(B(fN(x0),κeNMϵ))U(N,x_{0},\kappa,\epsilon)\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=f_{T^{N}(\hat{z})}^{-N}(B(f^{N}(x_{0}),\kappa\,e^{-NM\epsilon})) for every x^\hat{x} with π(x^)=x0\pi(\hat{x})=x_{0} whenever the inverse branch fTN(z^)Nf_{T^{N}(\hat{z})}^{-N} is well-defined on B(fN(x0),κeNMϵ)B(f^{N}(x_{0}),\kappa\,e^{-NM\epsilon}). Lemma 2.5 says that, for all ϵ>0\epsilon>0 sufficiently small, this happens for every x0π(Zν(ϵ))x_{0}\in\pi(Z_{\nu}(\epsilon)), 0<κ<r(ϵ)0<\kappa<r(\epsilon), and l0l\geq 0 if we take N=nlN=n_{l}, where the sequence {nl}l0\{n_{l}\}_{l\geq 0} is given by that statement (and depends on x0x_{0}; see Remark 2.6). In particular, U(nl,x0,κ,ϵ)U(n_{l},x_{0},\kappa,\epsilon) is well-defined under such conditions for all l0l\geq 0, and Corollary 2.4 can be applied with any y^\hat{y} such that y0=fnl(x0)y_{0}=f^{n_{l}}(x_{0}) and ynl=x0y_{-n_{l}}=x_{0} (observe that the factor eNMϵe^{-NM\epsilon} is not necessary to get this). We now aim at getting similar estimates valid for all Nn(ϵ)N\geq n(\epsilon). The factor eNMϵe^{-NM\epsilon} will need to be introduced for this reason.

In the next lemma and in the rest of the paper, we will only consider 0<ϵχmin0<\epsilon\ll\chi_{\min} sufficiently small as above, and Zν(ϵ)Z^{\prime}_{\nu}(\epsilon), Zν(ϵ)Z_{\nu}(\epsilon), r(ϵ)r(\epsilon), and n(ϵ)n(\epsilon) will be as in Corollary 2.4 and Lemma 2.5. Similarly, for every z^Zν(ϵ)\hat{z}\in Z_{\nu}(\epsilon) (and zπ(Zν(ϵ)))z\in\pi(Z_{\nu}(\epsilon))), the sequence {nl}l0\{n_{l}\}_{l\geq 0} is given by Lemma 2.5 (and Remark 2.6). Recall that all these definitions depend on ff and ν+(f)\nu\in\mathcal{M}^{+}(f). The sets U(N,x,κ,ϵ)U(N,x,\kappa,\epsilon) are as in Definition 2.7.

Lemma 2.8.

Fix 0<ϵχmin0<\epsilon\ll\chi_{\min}. Then, for all zπ(Zν(ϵ))z\in\pi(Z_{\nu}(\epsilon)), Nn(ϵ)N\geq n(\epsilon), and 0<κ<r(ϵ)0<\kappa<r(\epsilon), the set U(N,x,κ,ϵ)U(N,x,\kappa,\epsilon) is well-defined and we have

  1. (1)

    z(κeN(χj+(2M+1)ϵ))U(N,x,κ,ϵ)z(κeN(χjϵ))\mathcal{E}_{z}(\kappa\,e^{-N(\chi_{j}+(2M+1)\epsilon)})\subseteq U(N,x,\kappa,\epsilon)\subseteq\mathcal{E}_{z}(\kappa\,e^{-N(\chi_{j}-\epsilon)});

  2. (2)

    κ2ke2N(Lν+k(2M+1)ϵ)Vol(U(N,x,κ,ϵ))κ2ke2N(Lνkϵ)\kappa^{2k}e^{-2N(L_{\nu}+k(2M+1)\epsilon)}\leq\operatorname{Vol}(U(N,x,\kappa,\epsilon))\leq\kappa^{2k}e^{-2N(L_{\nu}-k\epsilon)};

  3. (3)

    eN(2M+1)kϵ|JacfN(y)||JacfN(w)|1eN(2M+1)kϵe^{-N(2M+1)k\epsilon}\leq|\operatorname{Jac}f^{N}(y)|\cdot|\operatorname{Jac}f^{N}(w)|^{-1}\leq e^{N(2M+1)k\epsilon} for every y,wU(N,x,κ,ϵ)y,w\in U(N,x,\kappa,\epsilon),

where MM is as in (2.1).

Proof.

Fix xπ(Zν(ϵ))x\in\pi(Z_{\nu}(\epsilon)) and 0<κ<r(ϵ)0<\kappa<r(\epsilon). Then xx corresponds to (at least) an orbit x^Zν(ϵ)\hat{x}\in Z_{\nu}(\epsilon) with π(x^)=x\pi(\hat{x})=x. By Lemma 2.5 and Remark 2.6, there exists a sequence {nl}l0\{n_{l}\}_{l\geq 0} with n0n(ϵ)n_{0}\leq n(\epsilon) and such that nl+1<(1+ϵ)nln_{l+1}<(1+\epsilon)n_{l} and fnl(x)π(Zν(ϵ))f^{n_{l}}(x)\in\pi(Z^{\prime}_{\nu}(\epsilon)) for all l0l\geq 0. By Corollary 2.4 (3b), (3c), and (3e), for all l0l\geq 0 and 0<t10<t\leq 1 the set

U~(nl,x,tκ) . . =fTnl(x^)nl(B(fnl(x0),tκ))\widetilde{U}(n_{l},x,t\kappa)\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=f_{T^{n_{l}}(\hat{x})}^{-n_{l}}\big{(}B(f^{n_{l}}(x_{0}),t\kappa)\big{)}

is well-defined and we have

(2.3) x(tκenl(χj+ϵ))U~(nl,x,tκ)x(tκenl(χjϵ)),\mathcal{E}_{x}(t\kappa\,e^{-n_{l}(\chi_{j}+\epsilon)})\subseteq\widetilde{U}(n_{l},x,t\kappa)\subseteq\mathcal{E}_{x}(t\kappa\,e^{-n_{l}(\chi_{j}-\epsilon)}),
(2.4) (tκ)2ke2nl(Lν+kϵ)Vol(U~(nl,x,tκ))(tκ)2ke2nl(Lνkϵ),(t\kappa)^{2k}e^{-2n_{l}(L_{\nu}+k\epsilon)}\leq\operatorname{Vol}(\widetilde{U}(n_{l},x,t\kappa))\leq(t\kappa)^{2k}e^{-2n_{l}(L_{\nu}-k\epsilon)},

and

(2.5) eknlϵ|Jacfnl(y)||Jacfnl(w)|1eknlϵ for every y,wU~(nl,x,κ).e^{-kn_{l}\epsilon}\leq|\operatorname{Jac}f^{n_{l}}(y)|\cdot|\operatorname{Jac}f^{n_{l}}(w)|^{-1}\leq e^{kn_{l}\epsilon}\quad\mbox{ for every }y,w\in\widetilde{U}(n_{l},x,\kappa).

Consider now any Nn(ϵ)N\geq n(\epsilon) and fix l=l(N)l^{\star}=l^{\star}(N) such that nlN<nl+1n_{l^{\star}}\leq N<n_{l^{\star}+1}. Such ll^{\star} exists since n0n(ϵ)n_{0}\leq n(\epsilon) and nln_{l}\to\infty as ll\to\infty. It follows from the definition (2.1) of MM that

(2.6) B(fN(z),κe(Nnl)M)fNnl(B(fnl(x),κe2(Nnl)M))B\big{(}f^{N}(z),\kappa\,e^{-(N-n_{l^{\star}})M}\big{)}\supseteq f^{N-n_{l^{\star}}}\big{(}B(f^{n_{l^{\star}}}(x),\kappa\,e^{-2(N-n_{l^{\star}})M})\big{)}

and

(2.7) fnl+1N(B(fN(x),κe(nl+1N)M))B(fnl+1(x),κ).f^{n_{l^{\star}+1}-N}\left(B\big{(}f^{N}(x),\kappa\,e^{-(n_{l^{\star}+1}-N)M}\big{)}\right)\subseteq B(f^{n_{l^{\star}+1}}(x),\kappa).

It follows from (2.7), the second inequality in (2.3) applied with l=l+1l=l^{\star}+1 and t=1t=1, and the fact that 0nl+1NϵnlϵN0\leq n_{l^{\star}+1}-N\leq\epsilon n_{l^{\star}}\leq\epsilon N, that U(N,x,κ,ϵ)U(N,x,\kappa,\epsilon) is well-defined and satisfies

(2.8) U(N,x,κ,ϵ)U~(nl+1,x,κ)x(κenl+1(χjϵ))x(κeN(χjϵ)).U(N,x,\kappa,\epsilon)\subseteq\widetilde{U}(n_{l^{\star}+1},x,\kappa)\subseteq\mathcal{E}_{x}(\kappa\,e^{-n_{l^{\star}+1}(\chi_{j}-\epsilon)})\subseteq\mathcal{E}_{x}(\kappa\,e^{-N(\chi_{j}-\epsilon)}).

Similarly, from (2.6), the first inequality in (2.3) applied with l=ll=l^{\star} and t=e2(Nnl)Mt=e^{-2(N-n_{l^{\star}})M}, and the fact that 0NnlϵnlϵN0\leq N-n_{l^{\star}}\leq\epsilon n_{l^{\star}}\leq\epsilon N we deduce that

U(N,x,κ,ϵ)\displaystyle U(N,x,\kappa,\epsilon) fTnl(x^)nl(B(fnl(x),κe2(Nnl)M))\displaystyle\supseteq f^{-n_{l^{\star}}}_{T^{n_{l^{\star}}}(\hat{x})}\left(B\big{(}f^{n_{l^{\star}}}(x),\kappa\,e^{-2(N-n_{l^{\star}})M}\big{)}\right)
x(κenl(χj(2M+1)ϵ))x(κeN(χj(2M+1)ϵ)),\displaystyle\supseteq\mathcal{E}_{x}(\kappa\,e^{-n_{l^{\star}}(\chi_{j}-(2M+1)\epsilon)})\supseteq\mathcal{E}_{x}(\kappa\,e^{-N(\chi_{j}-(2M+1)\epsilon)}),

which completes the proof of the first item.

The second and the third assertions follow from similar arguments, combining (2.6) and (2.7) with (2.4) and (2.5), respectively. ∎

The following corollary records a special case of the above lemma when all the Lyapunov exponents of ν+(f)\nu\in\mathcal{M}^{+}(f) are equal.

Corollary 2.9.

Assume that all the Lyapunov exponents of ν+(f)\nu\in\mathcal{M}^{+}(f) are equal to χ>0\chi>0. Then, for all 0<ϵχ0<\epsilon\ll\chi, xπ(Zν(ϵ))x\in\pi(Z_{\nu}(\epsilon)), Nn(ϵ)N\geq n(\epsilon), and 0<κ<r(ϵ)0<\kappa<r(\epsilon), we have

  1. (1)

    B(x,κeN(χ+(2M+1)ϵ))U(N,x,κ,ϵ)B(x,κeN(χϵ))B(x,\kappa\,e^{-N(\chi+(2M+1)\epsilon)})\subseteq U(N,x,\kappa,\epsilon)\subseteq B(x,\kappa\,e^{-N(\chi-\epsilon)});

  2. (2)

    κ2ke2kN(χ+(2M+1)ϵ)Vol(U(N,x,κ,ϵ))κ2ke2kN(χϵ)\kappa^{2k}e^{-2kN(\chi+(2M+1)\epsilon)}\leq\operatorname{Vol}(U(N,x,\kappa,\epsilon))\leq\kappa^{2k}e^{-2kN(\chi-\epsilon)}.

Remark 2.10.

If ff is hyperbolic, then we have Zνπ1(J(f))=Zπ1(J(f))Z_{\nu}\cap\pi^{-1}(J(f))=Z\cap\pi^{-1}(J(f)) for any νJ+(f)\nu\in\mathcal{M}^{+}_{J}(f). Moreover, observe that any ergodic probability measure on J(f)J(f) belongs to J+(f)\mathcal{M}^{+}_{J}(f). In particular, we have π(Zν)J(f)\pi(Z_{\nu})\supseteq J(f) for any ergodic probability measure on J(fJ(f). We also have Zν,απ1(J(f))=Zνπ1(J(f))Z_{\nu,\alpha}\cap\pi^{-1}(J(f))=Z_{\nu}\cap\pi^{-1}(J(f)) for all α\alpha sufficiently small, which implies that we can take Z(ϵ)=Z(ϵ)=ZZ(\epsilon)=Z^{\prime}(\epsilon)=Z for all ϵ\epsilon sufficiently small.

More generally, let XkX\subseteq\mathbb{P}^{k} be a closed invariant uniformly expanding set. Take νX+(f)\nu\in\mathcal{M}^{+}_{X}(f). Denoting by OXO_{X} the set of orbits {xn}nX\{x_{n}\}_{n\in\mathbb{Z}}\in X^{\mathbb{Z}}, it follows from the definition of ν^\hat{\nu} that ν^(OX)=1\hat{\nu}(O_{X})=1 and that we can assume that Xπ(Zν(ϵ))X\subseteq\pi(Z_{\nu}(\epsilon)) for all ϵ>0\epsilon>0. As ν^(OX)=1\hat{\nu}(O_{X})=1, we can also assume that π(Zν(ϵ))X\pi(Z_{\nu}(\epsilon))\subseteq X, hence π(Zν(ϵ))=X\pi(Z_{\nu}(\epsilon))=X, for all ϵ>0\epsilon>0.

2.3. Volume-conformal measures

We again fix in this section a holomorphic endomorphism ff of k\mathbb{P}^{k} of algebraic degree d2d\geq 2, and we let XX be a closed invariant set for ff. Recall that f|X:XXf|_{X}:X\to X is topologically exact if for any open set UkU\subset\mathbb{P}^{k} with UXU\cap X\neq\emptyset there exists n1n\geq 1 such that fn(U)Xf^{n}(U)\supseteq X.

Definition 2.11.

Given any t0t\geq 0, a probability measure μ\mu on XX is tt-volume-conformal on XX if, for every Borel subset AXA\subseteq X on which ff is invertible, we have

μ(f(A))=A|Jacf|t𝑑μ.\mu(f(A))=\int_{A}|\operatorname{Jac}f|^{t}d\mu.

We define

δX(f)\displaystyle\delta_{X}(f) . . =inf{t0:there exists a t-volume-conformal measure on X}.\displaystyle\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\inf\left\{t\geq 0\colon\mbox{there exists a $t$-volume-conformal measure on }X\right\}.
Lemma 2.12.

Assume that f|Xf|_{X} is topologically exact. Let μ\mu be a probability measure on XX which is tt-volume-conformal on XX for some t0t\geq 0. Then

  1. (1)

    the support of μ\mu is equal to XX;

  2. (2)

    for every r>0r>0 there exists constants 0<m=m(μ,r)10<m_{-}=m_{-}(\mu,r)\leq 1 and 0<m+=m+(μ,r)10<m_{+}=m_{+}(\mu,r)\leq 1 such that mμ(B(x,r))m+m_{-}\leq\mu(B(x,r))\leq m_{+} for every xXx\in X.

Proof.

Assume that there exists a point xXx\in X which does not belong to the support of μ\mu. Take a small ball BB centred at xx which is disjoint from the critical set C(f)C(f) of ff and such that μ(B)=0\mu(B)=0. As f|Xf|_{X} is topologically exact, we have Xfn(B)X\subseteq f^{n}(B) for some n1n\geq 1. Hence, it is enough to prove that μ(fn(B))=0\mu(f^{n}(B))=0 for all nn\in\mathbb{N}. Since BC(f)=B\cap C(f)=\emptyset, this is a consequence of the volume-conformality of μ\mu and the fact that μ(B)=0\mu(B)=0 (we need here to partition BB into subsets where fnf^{n} is injective in order to apply Definition 2.11). The first assertion follows.

The second assertion is a consequence of the first and the fact that, for every probability measure μ\mu on k\mathbb{P}^{k} and r>0r>0, there exist constants m±=m±(μ,r)m_{\pm}=m_{\pm}(\mu,r) such that (2) holds for every xx in the support of μ\mu. ∎

Recall that, for every ν+(f)\nu\in\mathcal{M}^{+}(f) and every ϵ\epsilon sufficiently small, Zν(ϵ)Z_{\nu}(\epsilon), r(ϵ)r(\epsilon), and n(ϵ)n(\epsilon) are given by Corollary 2.4 and Lemma 2.5, and the sets U(N,x,κ,ϵ)U(N,x,\kappa,\epsilon) are defined in Definition 2.7.

Lemma 2.13.

Assume that f|Xf|_{X} is topologically exact. Fix νX+(f)\nu\in\mathcal{M}_{X}^{+}(f), t0t\geq 0, and 0<ϵχmin0<\epsilon\ll\chi_{\min}, where χmin>0\chi_{\min}>0 is the smallest Lyapunov exponent of ν\nu. Then, for every 0<κ<r(ϵ)0<\kappa<r(\epsilon), every tt-volume-conformal probability measure μ\mu on XX, every xπ(Zν(ϵ))x\in\pi(Z_{\nu}(\epsilon)), and every Nn(ϵ)N\geq n(\epsilon), the set U=U(N,x,κ,ϵ)U=U(N,x,\kappa,\epsilon) is well-defined and satisfies

m(μ,κeMNϵ)CtκtketNk(5M+2)ϵμ(U)Vol(U)t/2Ctm+(μ,κeMNϵ)κtketNk(5M+2)ϵ,\frac{m_{-}(\mu,\kappa\,e^{-MN\epsilon})}{C^{t}\kappa^{tk}}e^{-tNk(5M+2)\epsilon}\leq\frac{\mu(U)}{\operatorname{Vol}(U)^{t/2}}\leq\frac{C^{t}m_{+}(\mu,\kappa\,e^{-MN\epsilon})}{\kappa^{tk}}e^{tNk(5M+2)\epsilon},

where MM is as in (2.1), the constants mm_{-} and m+m_{+} are as in Lemma 2.12, and CC is a positive constant independent of κ\kappa, ϵ\epsilon, xx, NN, ν\nu, μ\mu, and tt.

Proof.

Fix xπ(Z(ϵ))x\in\pi(Z(\epsilon)), Nn(ϵ)N\geq n(\epsilon), and 0<κ<r(ϵ)0<\kappa<r(\epsilon). The set U . . =U(N,x,κ,ϵ)U\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=U(N,x,\kappa,\epsilon) is well-defined by Lemma 2.8. We denote for simplicity by BB the ball B(fN(x),κeMNϵ)=fN(U).B(f^{N}(x),\kappa\,e^{-MN\epsilon})=f^{N}(U).

Let μ\mu be any tt-volume-conformal probability measure on XX. Since XX has a dense orbit, by Lemma 2.12 the support of μ\mu is equal to XX. Since fNf^{N} is injective on UU, by Definition 2.11 we have

μ(B)=μ(fN(U))=μ(fN(UX))=UX|JacfN|t𝑑μ=U|JacfN|t𝑑μ.\mu(B)=\mu(f^{N}(U))=\mu(f^{N}(U\cap X))=\int_{U\cap X}|\operatorname{Jac}f^{N}|^{t}d\mu=\int_{U}|\operatorname{Jac}f^{N}|^{t}d\mu.

We deduce from Lemma 2.8 (3) that

etkN(2M+1)ϵ|JacfN(x)|tμ(U)μ(B)etkN(2M+1)ϵ|JacfN(x)|tμ(U).e^{-tkN(2M+1)\epsilon}|\operatorname{Jac}f^{N}(x)|^{t}\mu(U)\leq\mu(B)\leq e^{tkN(2M+1)\epsilon}|\operatorname{Jac}f^{N}(x)|^{t}\mu(U).

It follows from the above expression that

(2.9) etkN(2M+1)ϵ|JacfN(x)|tmμ(U)etkN(2M+1)ϵ|JacfN(x)|tm+,\frac{e^{-tkN(2M+1)\epsilon}}{|\operatorname{Jac}f^{N}(x)|^{t}}\cdot m_{-}\leq\mu(U)\leq\frac{e^{tkN(2M+1)\epsilon}}{|\operatorname{Jac}f^{N}(x)|^{t}}\cdot m_{+},

where m . . =m(μ,κeMNϵ)m_{-}\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=m_{-}(\mu,\kappa\,e^{-MN\epsilon}) and m+ . . =m+(μ,κeMNϵ)m_{+}\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=m_{+}(\mu,\kappa\,e^{-MN\epsilon}) are as in Lemma 2.12.

Again by Lemma 2.8 (3), we also have

(2.10) e2kN(2M+1)ϵVol(B)|JacfN(x)|2Vol(U)=B(fN(x),κeMNϵ)|Jacfz^N|2e2kN(2M+1)ϵVol(B)|JacfN(x)|2,\frac{e^{-2kN(2M+1)\epsilon}\operatorname{Vol}(B)}{|\operatorname{Jac}f^{N}(x)|^{2}}\leq\operatorname{Vol}(U)=\int_{B(f^{N}(x),\kappa\,e^{-MN\epsilon})}|\operatorname{Jac}f^{-N}_{\hat{z}}|^{2}\leq\frac{e^{2kN(2M+1)\epsilon}\operatorname{Vol}(B)}{|\operatorname{Jac}f^{N}(x)|^{2}},

where the integral is taken with respect to the Fubini-Study metric, z^\hat{z} is any element in ZZ such that z0=fN(x)z_{0}=f^{N}(x) and zN=xz_{-N}=x, and we observe that fz^Nf^{-N}_{\hat{z}} is well-defined on fN(U)f^{N}(U) by Lemma 2.8.

Combining the inequalities (2.9) and (2.10), we see that

mVol(B)t/2etNk(4M+2)ϵVol(U)t/2μ(U)m+Vol(B)t/2etNk(4M+2)ϵVol(U)t/2.\frac{m_{-}}{\operatorname{Vol}(B)^{t/2}}e^{-tNk(4M+2)\epsilon}\operatorname{Vol}(U)^{t/2}\leq\mu(U)\leq\frac{m_{+}}{\operatorname{Vol}(B)^{t/2}}e^{tNk(4M+2)\epsilon}\operatorname{Vol}(U)^{t/2}.

The assertion follows from the last expression by observing that there exists a positive constant CC such that C2Vol(B(x,r))/r2kC2C^{-2}\leq\operatorname{Vol}(B(x,r))/r^{2k}\leq C^{2} for every xkx\in\mathbb{P}^{k} and 0<r<10<r<1. ∎

2.4. A pressure for expanding measures

Let ff be a holomorphic endomorphism of k\mathbb{P}^{k} of algebraic degree d2d\geq 2. For any invariant probability measure ν\nu and tt\in\mathbb{R}, we define

Pν(t) . . =hν(f)t|Jacf|dν=hν(f)tLν(f).P_{\nu}(t)\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=h_{\nu}(f)-t\int|\operatorname{Jac}f|d\nu=h_{\nu}(f)-tL_{\nu}(f).

Let XkX\subseteq\mathbb{P}^{k} be a closed invariant set for ff. We define a pressure function PX+P^{+}_{X} as

(2.11) PX+(t)\displaystyle P^{+}_{X}(t) . . =sup{Pν(t):νX+(f)}\displaystyle\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\sup\big{\{}P_{\nu}(t)\colon\nu\in\mathcal{M}_{X}^{+}(f)\big{\}}

and set

pX+(f) . . =inf{t:PX+(t)=0}.p^{+}_{X}(f)\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\inf\big{\{}t\colon P_{X}^{+}(t)=0\big{\}}.

We will drop the index XX when X=kX=\mathbb{P}^{k}.

Lemma 2.14.

Let XkX\subseteq\mathbb{P}^{k} be a closed invariant set for ff. Assume that X+(f)\mathcal{M}_{X}^{+}(f) is not empty. Then we have PX+(t)<P_{X}^{+}(t)<\infty for all tt\in\mathbb{R} and the function tPX+(t)t\mapsto P_{X}^{+}(t) is convex and non-increasing.

Proof.

Take νX+(f)\nu\in\mathcal{M}_{X}^{+}(f). As Lν(f)>0L_{\nu}(f)>0 and the topological entropy of ff is bounded by klogdk\log d [Gro03, DS10] we have PX+(t)klogdP^{+}_{X}(t)\leq k\log d for all t0t\geq 0. Take now t<0t<0. Since the function |Jacf||\operatorname{Jac}f| is bounded from above by a constant MM^{\prime} and hν(f)klogdh_{\nu}(f)\leq k\log d, we have PX+(t)klogd+|t|MP^{+}_{X}(t)\leq k\log d+|t|M^{\prime} for every t<0t<0. Hence, PX+(t)<P^{+}_{X}(t)<\infty for all tt\in\mathbb{R}.

For any given measure νX+(f)\nu\in\mathcal{M}_{X}^{+}(f), the function tPν(t)t\mapsto P_{\nu}(t) is non-increasing. It follows from its definition (2.11) that the function tPX+(t)t\mapsto P_{X}^{+}(t) is non-increasing. It is convex as it is a supremum of affine, hence convex, functions. ∎

The following example illustrates that Lemma 2.14 is false (even with X=J(f)X=J(f)) if we take the supremum over the set of all ergodic probability measures, with no requirement on the Lyapunov exponents, in the definition (2.11) of the pressure function PX+(t)P_{X}^{+}(t).

Example 2.15.

It is possible to construct endomorphisms ff of k\mathbb{P}^{k} admitting a saddle fixed point p0p_{0} in the Julia set and with |Jacf(p0)|<1|\operatorname{Jac}f(p_{0})|<1 (and actually also equal to 0). An example of this phenomenon is given for instance by Jonsson in [Jon99, Example 9.1], see also [BDM07, Theorem 6.3], [Taf10], and [BT17, Remark 2.6] for further examples. Consider the polynomial self-map ff of 2\mathbb{C}^{2} defined as

(z,w)(z2,w2+2(1+ηz)w),(z,w)\mapsto\left(z^{2},w^{2}+2(1+\eta-z)w\right),

which extends to 2\mathbb{P}^{2} as a holomorphic endomorphism. As ff preserves the families of the vertical lines parallel to {z=0}\{z=0\}, for every (z0,w0)2(z_{0},w_{0})\in\mathbb{C}^{2} the vertical eigenvalue of Df(z0,w0)Df_{(z_{0},w_{0})} is well-defined. It is immediate to check that, for 0η<1/20\leq\eta<1/2, the point p0=(1,0)p_{0}=(1,0) is a saddle fixed point, with vertical eigenvalue equal to 2η2\eta, and Jacobian equal to 4η4\eta. In particular, the Jacobian of ff at p0p_{0} can take any small non-negative value (including 0). The point p0p_{0} is in J(f)J(f) since J(f)J(f) is closed and, for Lebesgue almost all z0S1z_{0}\in S^{1}, the point (z0,0)(z_{0},0) belongs to J(f)J(f). This follows from a direct computation of the derivatives which, by Birkhoff’s ergodic theorem, gives that

Df(z0,0)n(2n002πlog|1+ηeiθ|dθ)=(2n02nlog|1+η|),Df^{n}_{(z_{0},0)}\sim\begin{pmatrix}2^{n}&0\\ \star&\int_{0}^{2\pi}\log|1+\eta-e^{i\theta}|d\theta\end{pmatrix}=\begin{pmatrix}2^{n}&0\\ \star&2^{n}\log|1+\eta|\end{pmatrix},

and the characterization of the Julia set of ff given in [Jon99, Corollary 4.4].

Consider the function

PJ(t) . . =supνPν(t)P_{J}(t)\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\sup_{\nu}P_{\nu}(t)

where now the supremum is taken over the set of all invariant probability measures supported on J(f)J(f). If ν0=δp0\nu_{0}=\delta_{p_{0}} is the Dirac mass at p0p_{0}, then the function tPν0(t)t\mapsto P_{\nu_{0}}(t) is increasing in tt and Pν0(0)=0P_{\nu_{0}}(0)=0. Hence, for such an endomorphism ff, the function PJ(t)P_{J}(t) is convex but it increases after some t0>0t_{0}>0 and has no zeroes.

Remark 2.16.

One could define P~J+(t)\widetilde{P}_{J}^{+}(t) by considering the set of all ergodic probability measures with positive sum of Lyapunov exponents in the definition of PJ+(t)P_{J}^{+}(t). However, it is unclear to us how to generalize many of the results in this paper, and in particular Theorem 1.1, to this larger class of measures. A priori, it could be possible that the first zero of PJ(t)P_{J}(t) is larger than the first zero of PJ+(t)P_{J}^{+}(t), but (possibly) equal to the first zero of P~J+(t)\widetilde{P}_{J}^{+}(t).

3. Exact volume dimension of measures in +(f)\mathcal{M}^{+}(f)

Let f:kkf:\mathbb{P}^{k}\to\mathbb{P}^{k} be a holomorphic endomorphism of algebraic degree d2d\geq 2. In this section we define a pointwise dynamical volume dimension for every measure ν+(f)\nu\in\mathcal{M}^{+}(f) and prove that it is constant ν\nu-almost everywhere.

Fix a measure ν+(f)\nu\in\mathcal{M}^{+}(f) and let χmin>0\chi_{\min}>0 be the smallest Lyapunov exponent of ν\nu. For every 0<ϵχmin0<\epsilon\ll\chi_{\min}, we fix Zν(ϵ)Z_{\nu}(\epsilon), n(ϵ)n(\epsilon), and r(ϵ)r(\epsilon) as given by Corollary 2.4 and Lemma 2.5. For every xπ(Zν(ϵ))x\in\pi(Z_{\nu}(\epsilon)), the sequence {nl}l0={nl(x)}l0\{n_{l}\}_{l\geq 0}=\{n_{l}(x)\}_{l\geq 0} is also given by Lemma 2.5; see Remark 2.6. For xπ(Zν(ϵ))x\in\pi(Z_{\nu}(\epsilon)), 0<κ<r(ϵ)0<\kappa<r(\epsilon), and Nn(ϵ)N\geq n(\epsilon), we define

(3.1) δx(ϵ,κ,N) . . =logν(U(N,x,κ,ϵ))logVol(U(N,x,κ,ϵ)),\delta_{x}(\epsilon,\kappa,N)\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\frac{\log\nu(U(N,x,\kappa,\epsilon))}{\log{\rm Vol}(U(N,x,\kappa,\epsilon))},

where U(N,x,κ,ϵ)U(N,x,\kappa,\epsilon) is as in Definition 2.7. Observe that, for every ϵ\epsilon, xx, κ\kappa, and NN as above, the definition of δx(ϵ,κ,N)\delta_{x}(\epsilon,\kappa,N) is well-posed by Lemma 2.8.

Recall that the set ZνZ_{\nu} (see Definition 2.3) satisfies Zν=ϵ>0Zν(ϵ)Z_{\nu}=\cup_{\epsilon>0}Z_{\nu}(\epsilon) up to a ν\nu-negligible set, and that the family {Zν(ϵ)}ϵ>0\{Z_{\nu}(\epsilon)\}_{\epsilon>0} is non-decreasing as ϵ0\epsilon\to 0. In particular, ν\nu-almost every xπ(Zν)x\in\pi(Z_{\nu}) belongs to π(Zν(ϵ))\pi(Z_{\nu}(\epsilon)) for every 0<ϵ<ϵ00<\epsilon<\epsilon_{0} for some ϵ0=ϵ0(x)\epsilon_{0}=\epsilon_{0}(x). For every such xx, we define the upper and the lower local volume dimension at xx as

(3.2) δ¯x . . =lim supϵ0lim supκ0lim supNδx(ϵ,κ,N) and δ¯x . . =lim infϵ0lim infκ0lim infNδx(ϵ,κ,N),\overline{\delta}_{x}\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\limsup_{\epsilon\to 0}\limsup_{\kappa\to 0}\limsup_{N\to\infty}\delta_{x}(\epsilon,\kappa,N)\quad\mbox{ and }\quad\underline{\delta}_{x}\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\liminf_{\epsilon\to 0}\liminf_{\kappa\to 0}\liminf_{N\to\infty}\delta_{x}(\epsilon,\kappa,N),

respectively, where δx(ϵ,κ,N)\delta_{x}(\epsilon,\kappa,N) is as in (3.1).

Definition 3.1.

If δ¯x=δ¯x\underline{\delta}_{x}=\overline{\delta}_{x}, we say that δx . . =δ¯x=δ¯x\delta_{x}\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\underline{\delta}_{x}=\overline{\delta}_{x} is the local volume dimension of ν\nu at xx. We say that ν+(f)\nu\in\mathcal{M}^{+}(f) is exact volume-dimensional if the local volume dimension δx\delta_{x} exists for ν\nu-almost every xx.

The main result of this section is the following theorem. Recall that hν(f)h_{\nu}(f), Lν(f)L_{\nu}(f), and χmin\chi_{\min} denote the measure-theoretic entropy, the sum of the Lyapunov exponents, and the smallest Lyapunov exponent of ν\nu, respectively.

Theorem 3.2.

Let f:kkf:\mathbb{P}^{k}\to\mathbb{P}^{k} be a holomorphic endomorphism of algebraic degree d2d\geq 2. Take ν+(f)\nu\in\mathcal{M}^{+}(f) and 0<ϵχmin0<\epsilon\ll\chi_{\min}. Then, for ν\nu-almost all xπ(Zν(ϵ))x\in\pi(Z_{\nu}(\epsilon)) and all 0<κ<r(ϵ)0<\kappa<r(\epsilon), there exists integers m1(ϵ,x)n(ϵ)m_{1}(\epsilon,x)\geq n(\epsilon) and m2(ϵ,κ)0m_{2}(\epsilon,\kappa)\geq 0 such that

hν(f)2Lν(f)cϵδx(ϵ,κ,N)hν(f)2Lν(f)+cϵ for all Nm1(ϵ,x)+m2(ϵ,κ),\frac{h_{\nu}(f)}{2L_{\nu}(f)}-c\epsilon\leq\delta_{x}(\epsilon,\kappa,N)\leq\frac{h_{\nu}(f)}{2L_{\nu}(f)}+c\epsilon\quad\mbox{ for all }N\geq m_{1}(\epsilon,x)+m_{2}(\epsilon,\kappa),

where δx(ϵ,κ,N)\delta_{x}(\epsilon,\kappa,N) is as in (3.1) and c>0c>0 is a constant independent of ϵ\epsilon, xx, and κ\kappa.

Remark 3.3.

Although Theorem 3.2 is stated for points xπ(Zν(ϵ))x\in\pi(Z_{\nu}(\epsilon)), we can associate to ν^\hat{\nu}-almost every x^Zν(ϵ)\hat{x}\in Z_{\nu}(\epsilon) the integer m1(ϵ,x^) . . =m1(ϵ,x0)m_{1}(\epsilon,\hat{x})\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=m_{1}(\epsilon,x_{0}), where, since we have x0π(Zν(ϵ))x_{0}\in\pi(Z_{\nu}(\epsilon)), the number m1(ϵ,x0)m_{1}(\epsilon,x_{0}) is given by Theorem 3.2.

The following consequence of Theorem 3.2 shows that every ν+(f)\nu\in\mathcal{M}^{+}(f) is exact volume-dimensional.

Corollary 3.4.

Let f:kkf:\mathbb{P}^{k}\to\mathbb{P}^{k} be an endomorphism of algebraic degree d2d\geq 2 and take ν+(f).\nu\in\mathcal{M}^{+}(f). For ν\nu-almost every xkx\in\mathbb{P}^{k}, the local volume dimension δx\delta_{x} is well-defined and equal to (2Lν(f))1hν(f)(2L_{\nu}(f))^{-1}h_{\nu}(f).

Proof.

Recall that the family {Zν(ϵ)}ϵ>0\{Z_{\nu}(\epsilon)\}_{\epsilon>0} is non-decreasing for ϵ0\epsilon\to 0, and that we have Zν=ϵ>0Zν(ϵ)Z_{\nu}=\cup_{\epsilon>0}Z_{\nu}(\epsilon) up to a ν\nu-negligible set. In particular, for ν\nu-almost every xkx\in\mathbb{P}^{k} there exists ϵ0=ϵ(x0)>0\epsilon_{0}=\epsilon(x_{0})>0 such that xx belongs to π(Zν(ϵ))\pi(Z_{\nu}(\epsilon)) for every 0<ϵ<ϵ00<\epsilon<\epsilon_{0}. The assertion follows from the definition (3.2) of the upper and lower volume dimensions and Theorem 3.2. ∎

The rest of the section is devoted to the proof of Theorem 3.2. We will follow the general strategy presented in [PU10, Section 11.4] but we will need to use the results in Section 2.2 to replace the distortion estimates for univalent maps in dimension 11.

3.1. Proof of Theorem 3.2: a reduction

Fix a countable measurable partition 𝒫\mathcal{P} of k\mathbb{P}^{k}. Up to taking the elements of the partition sufficiently small, we can assume that the entropy hν(f,𝒫)h_{\nu}(f,\mathcal{P}) of the partition 𝒫\mathcal{P} satisfies hν(f)ϵhν(f,𝒫)hν(f)h_{\nu}(f)-\epsilon\leq h_{\nu}(f,\mathcal{P})\leq h_{\nu}(f). Recall that, by the Shannon-McMillan-Breiman Theorem [Par69, Wal00] for ν\nu-almost every xkx\in\mathbb{P}^{k} we have

limn1nlogν(𝒫n(x))=:hν(f,𝒫).\lim_{n\to\infty}-\frac{1}{n}\log\nu(\mathcal{P}^{n}(x))=:h_{\nu}(f,\mathcal{P}).

Here 𝒫n\mathcal{P}^{n} is the partition generated by 𝒫,f1𝒫,,fn𝒫\mathcal{P},f^{-1}\mathcal{P},\dots,f^{-n}\mathcal{P} (i.e., the partition whose elements are the sets of the form P0f1(P1)fn(Pn)P_{0}\cap f^{-1}(P_{1})\cap\ldots\cap f^{-n}(P_{n}) for P0,,Pn𝒫P_{0},\dots,P_{n}\in\mathcal{P}), and 𝒫n(x)\mathcal{P}^{n}(x) denotes the element of the partition 𝒫n\mathcal{P}^{n} containing xx.

Proposition 3.5.

Fix ν+(f)\nu\in\mathcal{M}^{+}(f). For every 0<ϵχmin0<\epsilon\ll\chi_{\min} there exist two partitions 𝒫1\mathcal{P}_{1} and 𝒫2\mathcal{P}_{2} with hν(f,𝒫1)hν(f)ϵh_{\nu}(f,\mathcal{P}_{1})\geq h_{\nu}(f)-\epsilon and four constants bE,bFb_{E},b_{F} (independent of ϵ\epsilon) and cE,cF>0c_{E},c_{F}>0 (possibly depending on ϵ\epsilon) such that for ν\nu-almost every xπ(Zν(ϵ))x\in\pi(Z_{\nu}(\epsilon)) there exists an integer m(ϵ,x)n(ϵ)m(\epsilon,x)\geq n(\epsilon) such that for all nm(ϵ,x)n\geq m(\epsilon,x), we have

E(n) . . =x(cEen(χj+bEϵ))𝒫1n(x) and 𝒫2n(x)F(n) . . =x(cFen(χjbFϵ)).E(n)\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\mathcal{E}_{x}(c_{E}e^{-n(\chi_{j}+b_{E}\epsilon)})\subseteq\mathcal{P}_{1}^{n}(x)\quad\text{ and }\quad\mathcal{P}_{2}^{n}(x)\subseteq F(n)\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\mathcal{E}_{x}(c_{F}e^{-n(\chi_{j}-b_{F}\epsilon)}).

We prove the existence of the sequences E(n)E(n) and F(n)F(n) and partitions 𝒫1\mathcal{P}_{1} and 𝒫2\mathcal{P}_{2} in the next two subsections. We now show how Theorem 3.2 is a consequence of Proposition 3.5.

Proof of Theorem 3.2 assuming Proposition 3.5.

We fix ϵ>0\epsilon>0 as in the statement, xπ(Zν(ϵ))x\in\pi(Z_{\nu}(\epsilon)), and 0<κ<r(ϵ)0<\kappa<r(\epsilon). For every Nn(ϵ)N\geq n(\epsilon) and 0<κ<r(ϵ)0<\kappa<r(\epsilon), define the integers nE(N,κ)n_{E}(N,\kappa) and nF(N,κ)n_{F}(N,\kappa) as

nE(N,κ) . . =minj(χjϵ)N+logcElogκχj+bEϵn_{E}(N,\kappa)\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\min_{j}\Big{\lfloor}\frac{(\chi_{j}-\epsilon)N+\log c_{E}-\log\kappa}{\chi_{j}+b_{E}\epsilon}\Big{\rfloor}

and

nF(N,κ) . . =maxj[χj+(2M+1)ϵ)]N+logcFlogκχjbFϵ,n_{F}(N,\kappa)\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\max_{j}\Big{\lceil}\frac{[\chi_{j}+(2M+1)\epsilon)]N+\log c_{F}-\log\kappa}{\chi_{j}-b_{F}\epsilon}\Big{\rceil},

where bE,bF,cE,cFb_{E},b_{F},c_{E},c_{F} are as in Proposition 3.5 and we recall that the χj\chi_{j}’s are the Lyapunov exponents of ν\nu, which are strictly positive. Then, by Lemma 2.8 (1) and (2) and Proposition 3.5, for all 0<κ<r(ϵ)0<\kappa<r(\epsilon), we have

(3.3) 𝒫2nF(N,κ)(x)F(nF(N,κ))U(N,x,κ,ϵ)E(nE(N,κ))𝒫1nE(N,κ)(x) for all Nm(ϵ,x),\mathcal{P}^{n_{F}(N,\kappa)}_{2}(x)\subseteq F(n_{F}(N,\kappa))\subseteq U(N,x,\kappa,\epsilon)\subseteq E(n_{E}(N,\kappa))\subseteq\mathcal{P}^{n_{E}(N,\kappa)}_{1}(x)\mbox{ for all }N\geq m(\epsilon,x),

where m(ϵ,x)m(\epsilon,x) is as in Proposition 3.5, and

(3.4) κ2ke2N(Lν+k(2M+1)ϵ)Vol(U(N,x,κ,ϵ))κ2ke2N(Lνkϵ) for all Nn(ϵ),\kappa^{2k}e^{-2N(L_{\nu}+k(2M+1)\epsilon)}\leq\operatorname{Vol}(U(N,x,\kappa,\epsilon))\leq\kappa^{2k}e^{-2N(L_{\nu}-k\epsilon)}\quad\mbox{ for all }N\geq n(\epsilon),

where we recall that MM is as in (2.1).

It follows from (3.3) and the Shannon-McMillan-Breiman Theorem that there exists m(ϵ,x)m(ϵ,x)m^{\prime}(\epsilon,x)\geq m(\epsilon,x) and m′′(ϵ,κ)1m^{\prime\prime}(\epsilon,\kappa)\gg 1 such that

(3.5) (hν(f)2ϵ)(limNnE(N,κ)Nϵ)logν(U(N,x,κ,ϵ))N(hν(f)+ϵ)(limNnF(N,κ)N+ϵ)(h_{\nu}(f)-2\epsilon)\Big{(}\lim_{N\to\infty}\frac{n_{E}(N,\kappa)}{N}-\epsilon\Big{)}\leq-\frac{\log\nu(U(N,x,\kappa,\epsilon))}{N}\leq(h_{\nu}(f)+\epsilon)\Big{(}\lim_{N\to\infty}\frac{n_{F}(N,\kappa)}{N}+\epsilon\Big{)}

for all N>m(ϵ,x)+m′′(ϵ,κ)N>m^{\prime}(\epsilon,x)+m^{\prime\prime}(\epsilon,\kappa). We used here the fact that, since κ<r(ϵ)\kappa<r(\epsilon), the integers nE(N,κ)n_{E}(N,\kappa) and nF(N,κ)n_{F}(N,\kappa) are bounded below by quantities which are independent of 0<κ<r(ϵ)0<\kappa<r(\epsilon). Similarly, it follows from (3.4) that there exists m′′′(ϵ,κ)1m^{\prime\prime\prime}(\epsilon,\kappa)\gg 1 such that

(3.6) 2(Lν(f)kϵ)ϵlogVol(U(N,x,κ,ϵ))N2(Lν(f)+(2M+1)ϵ)+ϵ2(L_{\nu}(f)-k\epsilon)-\epsilon\leq-\frac{\log\operatorname{Vol}(U(N,x,\kappa,\epsilon))}{N}\leq 2(L_{\nu}(f)+(2M+1)\epsilon)+\epsilon

for all N>m′′′(ϵ,κ)+n(ϵ)N>m^{\prime\prime\prime}(\epsilon,\kappa)+n(\epsilon).

Setting

m1(ϵ,x) . . =m(ϵ,x)m(ϵ,x)n(ϵ) and m2(ϵ,κ) . . =max{m′′(ϵ,κ),m′′′(ϵ,κ)},m_{1}(\epsilon,x)\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=m^{\prime}(\epsilon,x)\geq m(\epsilon,x)\geq n(\epsilon)\quad\mbox{ and }\quad m_{2}(\epsilon,\kappa)\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\max\{m^{\prime\prime}(\epsilon,\kappa),m^{\prime\prime\prime}(\epsilon,\kappa)\},

the assertion follows combining (3.5), (3.6), and the definitions of nE(N,κ)n_{E}(N,\kappa) and nF(N,κ)n_{F}(N,\kappa). ∎

3.2. Proof of Proposition 3.5: the existence of E(n)E(n) and 𝒫1\mathcal{P}_{1}

We will need the following lemma, see for instance [PU10, Corollary 9.1.10].

Lemma 3.6.

Let (𝒳,δ)(\mathcal{X},\delta) be a compact metric space, f:𝒳𝒳f:\mathcal{X}\to\mathcal{X} a measurable map with respect to the Borel σ\sigma-algebra on 𝒳\mathcal{X} and ν\nu an ff-invariant Borel probability measure. Then for every r>0r>0, there exists 𝒳0𝒳\mathcal{X}_{0}\subseteq\mathcal{X} with ν(𝒳0)=1\nu(\mathcal{X}_{0})=1 and a finite partition 𝒫\mathcal{P} of 𝒳\mathcal{X} into Borel sets of positive measure ν\nu and of diameter smaller than rr such that, for every ϵ>0\epsilon>0 and every x𝒳0x\in\mathcal{X}_{0}, there exists an integer m0=m0(ϵ,x)m_{0}=m_{0}(\epsilon,x) such that

B𝒳(fn(x),enϵ)𝒫(fn(x)) for every nm0,B_{\mathcal{X}}(f^{n}(x),e^{-n\epsilon})\subset\mathcal{P}(f^{n}(x))\quad\mbox{ for every }n\geq m_{0},

where B𝒳(y,a)B_{\mathcal{X}}(y,a) denotes the open ball in 𝒳\mathcal{X} of radius aa and center y𝒳y\in\mathcal{X}.

Fix 0<ϵχmin0<\epsilon\ll\chi_{\min}. Let 𝒳0\mathcal{X}_{0}, 𝒫\mathcal{P}, and m0(ϵ,x)m_{0}(\epsilon,x) be as given by Lemma 3.6 applied with 𝒳=Supp ν\mathcal{X}=\text{Supp }\nu. Up to taking rr sufficiently small, we can assume that hν(f,𝒫)hν(f)ϵh_{\nu}(f,\mathcal{P})\geq h_{\nu}(f)-\epsilon. Up to replacing Zν(ϵ)Z_{\nu}(\epsilon) with π1(𝒳0)Zν(ϵ)\pi^{-1}(\mathcal{X}_{0})\cap Z_{\nu}(\epsilon), we can assume that the conclusion of Lemma 3.6 holds for all xπ(Zν(ϵ))x\in\pi(Z_{\nu}(\epsilon)).

Fix xπ(Zν(ϵ))x\in\pi(Z_{\nu}(\epsilon)). In particular, there exists x^Zν(ϵ)\hat{x}\in Z_{\nu}(\epsilon) with π(x^)=x\pi(\hat{x})=x. Let {nl}l0\{n_{l}\}_{l\geq 0} be the sequence associated to x^\hat{x} by Lemma 2.5. We fix l0l_{0}\in\mathbb{N} such that nl0m0(ϵ,x)n_{l_{0}}\geq m_{0}(\epsilon,x). Recall that MM is as in (2.1).

Consider an integer nnl0n\geq n_{l_{0}} and the dynamical ellipse

E(n) . . =x(Cr(ϵ)en(χj+(M+2)ϵ)),E(n)\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\mathcal{E}_{x}\big{(}Cr(\epsilon)e^{-n(\chi_{j}+(M+2)\epsilon)}\big{)},

where 0<C<10<C<1 is a constant small enough so that

(3.7) fq(E(n))𝒫(fq(x)) for every qnl0.f^{q}(E(n))\subset\mathcal{P}(f^{q}(x))\quad\mbox{ for every }q\leq n_{l_{0}}.

We now show that fq(E(n))𝒫(fq(x))f^{q}(E(n))\subset\mathcal{P}(f^{q}(x)) for all nl0qnn_{l_{0}}\leq q\leq n. To this end, fix one such qq and let l=l(q)l0l^{\star}=l^{\star}(q)\geq l_{0} be such that nlq<nl+1n_{l^{\star}}\leq q<n_{l^{\star}+1}. Since Tnl(x^)Zν(ϵ)T^{n_{l^{\star}}}(\hat{x})\in Z^{\prime}_{\nu}(\epsilon) by Lemma 2.5 and π(Tnl(x^))=fnl(x)\pi(T^{n_{l^{\star}}}(\hat{x}))=f^{n_{l^{\star}}}(x), Theorem 2.1 and Corollary 2.4 (3c) imply that there exists a holomorphic inverse branch gl . . =ffnl(x)^nl:B(fnl(x),r(ϵ))kg_{l^{\star}}\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=f_{\widehat{f^{n_{l^{\star}}}(x)}}^{-n_{l^{\star}}}:B(f^{n_{l^{\star}}}(x),r(\epsilon))\to\mathbb{P}^{k} of fnlf^{n_{l^{\star}}} such that gl(fnl(x))=xg_{l^{\star}}(f^{n_{l^{\star}}}(x))=x and

x(r(ϵ)enl(χj+ϵ))gl(B(fnl(x),r(ϵ))).\mathcal{E}_{x}\big{(}r(\epsilon)e^{-n_{l^{\star}}(\chi_{j}+\epsilon)}\big{)}\subseteq g_{l^{\star}}\big{(}B(f^{n_{l^{\star}}}(x),r(\epsilon))\big{)}.

Set

E(n) . . =x(r(ϵ)enl(χj+ϵ)).E^{\prime}(n)\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\mathcal{E}_{x}\big{(}r(\epsilon)e^{-n_{l^{\star}}(\chi_{j}+\epsilon)}\big{)}.

Then E(n)E(n)E(n)\subset E^{\prime}(n) (by the choice of CC and the inequality nnln\geq n_{l^{\star}}) and Corollary 2.4 (3c) gives

fnl(E(n))=fnl(x(Cr(ϵ)en(χj+(M+2)ϵ)))fnl(x)(Cr(ϵ)eχj(nnl)eϵnl(M+2)ϵn).f^{n_{l^{\star}}}(E(n))=f^{n_{l^{\star}}}\Big{(}\mathcal{E}_{x}\big{(}Cr(\epsilon)e^{-n(\chi_{j}+(M+2)\epsilon)}\big{)}\Big{)}\subseteq\mathcal{E}_{f^{n_{l^{\star}}}(x)}\left(Cr(\epsilon)e^{-\chi_{j}(n-n_{l^{\star}})}e^{\epsilon n_{l^{\star}}-(M+2)\epsilon n}\right).

Since nnln\geq n_{l^{\star}}, 0qnlϵnl0\leq q-n_{l^{\star}}\leq\epsilon n_{l^{\star}} (by the definition of the sequence {nl}l0\{n_{l}\}_{l\geq 0} in Lemma 2.5), 0<Cr(ϵ)<10<Cr(\epsilon)<1 (as 0<r(ϵ)<10<r(\epsilon)<1 by Corollary 2.4), qnq\leq n, q<nl+1q<n_{l^{\star}+1}, and all the χj\chi_{j}’s are strictly positive, by the definition (2.1) of MM and the above expression we deduce that

fq(E(n))\displaystyle f^{q}(E(n)) =fqnl(fnl(E(n)))fq(x)(Cr(ϵ)eχj(nnl)eϵnl(M+2)ϵne(qnl)M)\displaystyle=f^{q-n_{l^{\star}}}(f^{n_{l^{\star}}}(E(n)))\subseteq\mathcal{E}_{f^{q}(x)}\left(Cr(\epsilon)e^{-\chi_{j}(n-n_{l^{\star}})}e^{\epsilon n_{l^{\star}}-(M+2)\epsilon n}e^{(q-n_{l^{\star}})M}\right)
B(fq(x),eϵnle2ϵnMϵneϵnlM)=B(fq(x),eϵ(nln)M+ϵ(nln)ϵn)\displaystyle\subseteq B\left(f^{q}(x),e^{\epsilon n_{l^{\star}}}e^{-2\epsilon n-M\epsilon n}e^{\epsilon n_{l^{\star}}M}\right)=B\left(f^{q}(x),e^{\epsilon(n_{l^{\star}}-n)M+\epsilon(n_{l^{\star}}-n)-\epsilon n}\right)
B(fq(x),eϵn)B(fq(x),eϵq).\displaystyle\subseteq B\left(f^{q}(x),e^{-\epsilon n}\right)\subseteq B\left(f^{q}(x),e^{-\epsilon q}\right).

As qnl0m0(ϵ,x)q\geq n_{l_{0}}\geq m_{0}(\epsilon,x), by Lemma 3.6 we have B(fq(x),eϵq)𝒫(fq(x))B\left(f^{q}(x),e^{-\epsilon q}\right)\subset\mathcal{P}(f^{q}(x)). It follows that fq(E(n))𝒫(fq(x))f^{q}(E(n))\subset\mathcal{P}(f^{q}(x)) for all nl0qnn_{l_{0}}\leq q\leq n. Together with (3.7), setting 𝒫1 . . =𝒫\mathcal{P}_{1}\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\mathcal{P} this inclusion implies that E(n)𝒫1n(x),E(n)\subseteq\mathcal{P}_{1}^{n}(x), as desired.

3.3. Proof of Proposition 3.5: the existence of F(n)F(n) and 𝒫2\mathcal{P}_{2}

We work with the same setting and notations as in Section 3.2. We need the following lemma, see for instance [PU10, Lemma 11.3.2]. Recall that the entropy of a countable partition 𝒫={Pi}\mathcal{P}=\{P_{i}\} with respect to a probability measure μ\mu is defined as

Hμ(𝒫) . . =iμ(Pi)log(μ(Pi)).H_{\mu}(\mathcal{P})\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\sum_{i}-\mu(P_{i})\log(\mu(P_{i})).
Lemma 3.7.

Let μ\mu be a Borel probability measure on a bounded subset AA of a Euclidean space, and ρ:A(0,1]\rho:A\to(0,1] a measurable function such that logρ\log\rho is integrable with respect to μ\mu. There exists a countable measurable partition 𝒫\mathcal{P} of AA such that Hμ(𝒫)<H_{\mu}(\mathcal{P})<\infty and

diam(𝒫(x))ρ(x) for μ-almost every xA.{\rm diam}(\mathcal{P}(x))\leq\rho(x)\quad\mbox{ for }\mu\mbox{-almost every }x\in A.

Recall that C(f)C(f) denotes the critical set of ff and that ν(C(f))=0\nu(C(f))=0. For xC(f)x\notin C(f), define the function

(3.8) ρ(x) . . =cr(ϵ)min{1,|Jacf|},\rho(x)\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=c\,r(\epsilon)\min\big{\{}{1},|\operatorname{Jac}f|\big{\}},

where c<1c<1 is sufficiently small so that ff is injective on the ball B(x,ρ(x))B(x,\rho(x)) for every xkC(f)x\in\mathbb{P}^{k}\setminus C(f). Such constant exists because, since the function Jacf(x)\operatorname{Jac}f(x) is holomorphic in xx, there exists a positive constant c0c_{0} such that |Jacf(x)|c0dist(x,C(f))|\operatorname{Jac}f(x)|\leq c_{0}\cdot\operatorname{dist}(x,C(f)) for every xkx\in\mathbb{P}^{k}. For the same reason, the function logρ\log\rho is integrable with respect to ν\nu, since by assumption the Lyapunov exponents of ν\nu are not equal to -\infty, hence log|Jacf|\log|\operatorname{Jac}f| is integrable with respect to ν\nu.

Consider a partition 𝒫\mathcal{P} given by Lemma 3.7, applied with μ=ν\mu=\nu, A=Supp νA=\text{Supp }\nu, and the function ρ\rho as in (3.8). In particular, for ν\nu-almost every xSupp νx\in\text{Supp }\nu we have 𝒫(x)B(x,ρ(x))\mathcal{P}(x)\subset B(x,\rho(x)). For every n1n\geq 1, define

Vn(x,ρ) . . =j=0n1fjB(fj(x),ρ(fj(x))).V_{n}(x,\rho)\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\bigcap_{j=0}^{n-1}f^{-j}B(f^{j}(x),\rho(f^{j}(x))).

It follows from the definition of ρ\rho that ff is injective on B(fj(x),ρ(fj(x)))B(f^{j}(x),\rho(f^{j}(x))) for all xkx\in\mathbb{P}^{k} and 0jn10\leq j\leq n-1. As a consequence, for every n1n\geq 1 and for ν\nu-almost every xSupp νx\in\text{Supp }\nu, the map fnf^{n} is injective on Vn(x,ρ)V_{n}(x,\rho) and 𝒫n(x)Vn(x,ρ)\mathcal{P}^{n}(x)\subset V_{n}(x,\rho). It is then enough to show that, for every n>n(ϵ)n>n(\epsilon), the set Vn(x,ρ)V_{n}(x,\rho) is contained in a set F(n) . . =x(cFen(χjbFϵ))F(n)\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\mathcal{E}_{x}(c_{F}e^{-n(\chi_{j}-b_{F}\epsilon)}), for some bFb_{F} and cFc_{F} as in the statement of Proposition 3.5.

Let ll^{\star} be the largest index of the sequence {nl}l0\{n_{l}\}_{l\geq 0} given by Lemma 2.5 such that nln1n_{l^{\star}}\leq n-1 (such ll^{\star} exists since n>n(ϵ)n>n(\epsilon)). As in Section 3.2, set gl . . =ffnl(x)^nlg_{l^{\star}}\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=f^{-n_{l^{\star}}}_{\widehat{f^{n_{l^{\star}}}(x)}}. Then glg_{l^{\star}} is well-defined on B(fnl(x),r(ϵ))B(f^{n_{l^{\star}}}(x),r(\epsilon)). By the above, and in particular by the injectivity of ff on VnV_{n}, we have

Vn(x,ρ)gl(B(fnl(x),ρ(fnl(x)))).V_{n}(x,\rho)\subset g_{l^{\star}}\big{(}B(f^{n_{l^{\star}}}(x),\rho(f^{n_{l^{\star}}}(x)))\big{)}.

By Corollary 2.4 (3c), we deduce that

Vn(x,ρ)x(Kenl(χjϵ))V_{n}(x,\rho)\subset\mathcal{E}_{x}\big{(}Ke^{-n_{l^{\star}}(\chi_{j}-\epsilon)}\big{)}

for some constant K>0K>0 independent of xx and nn. Since n1nl+1(1+ϵ)nln-1\leq n_{l^{\star}+1}\leq(1+\epsilon)n_{l^{\star}}, we deduce that

Vn(x,ρ)F(n) . . =x(Ke(n1)(1+ϵ)1(χjϵ)).V_{n}(x,\rho)\subset F^{\prime}(n)\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\mathcal{E}_{x}\big{(}Ke^{-(n-1)(1+\epsilon)^{-1}(\chi_{j}-\epsilon)}\big{)}.

Set F(n) . . =x(cFen(χjbFϵ))F(n)\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\mathcal{E}_{x}\big{(}c_{F}e^{-n(\chi_{j}-b_{F}\epsilon)}\big{)}, where cF . . =Kc_{F}\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=K and bF . . =(1+ϵ)1minj(χj+1)b_{F}\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=(1+\epsilon)^{-1}\min_{j}(\chi_{j}+1). The assertion follows.

This concludes the proof of Proposition 3.5 and therefore also the proof of Theorem 3.2.

4. Volume dimension of measures in +(f)\mathcal{M}^{+}(f)

Let f:kkf\colon\mathbb{P}^{k}\to\mathbb{P}^{k} be a holomorphic endomorphism of algebraic degree d2d\geq 2. The goal of this section is to define volume dimensions for sets and measures and study their properties. More specifically, in Sections 4.1 and 4.2 we define and study the volume dimension VD(ν)\operatorname{VD}(\nu) for measures ν+(f)\nu\in\mathcal{M}^{+}(f) and VDν(X)\operatorname{VD}_{\nu}(X) for subsets XX of the support of ν\nu. In Section 4.3 we prove a criterion to relate the volume dimension of a set of positive measure to the local volume dimensions defined in Section 3; see Proposition 4.26. This criterion, together with Theorem 3.2, will allow us to prove Theorem 1.1 in the next section.

4.1. Definition of volume dimension and first properties

Given ϵ>0\epsilon>0, κ>0\kappa>0, NN\in\mathbb{N}, and WkW\subseteq\mathbb{P}^{k}, we consider the collection 𝒰Nκ(W,ϵ)\mathcal{U}^{\kappa}_{N}(W,\epsilon) of open subsets of k\mathbb{P}^{k} given by

𝒰Nκ(W,ϵ) . . ={Uk:xW such that U=U(N,x,κ,ϵ)}\mathcal{U}_{N}^{\kappa}(W,\epsilon)\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\{U\subset\mathbb{P}^{k}\colon\exists x\in W\text{ such that }U=U(N,x,\kappa,\epsilon)\}

where U(N,x,κ,ϵ)U(N,x,\kappa,\epsilon) is as in Definition 2.7. Recall in particular that each U(N,x,κ,ϵ)U(N,x,\kappa,\epsilon) (if it exists) is an open neighbourhood of xx. Given ϵ>0\epsilon>0 and Uκ>0,N0𝒰Nκ(k,ϵ)U\in\bigcup_{\kappa>0,N\geq 0}\mathcal{U}_{N}^{\kappa}(\mathbb{P}^{k},\epsilon), we denote by N(U),κ(U)N(U),\kappa(U), and z(U)z(U) the parameters associated to UU as in that definition, i.e., such that

fN(U)(U)=B(fN(U)(z(U)),κ(U)eN(U)Mϵ),f^{N(U)}(U)=B(f^{N(U)}(z(U)),\kappa(U)\,e^{-N(U)M\epsilon}),

where MM is as in (2.1).

Remark 4.1.

Let U1U2U_{1}\neq U_{2} have the same parameters N=N(Ui)N=N(U_{i}) and κ=κ(Ui)\kappa=\kappa(U_{i}) and assume that z(U1)z(U_{1}) and z(U2)z(U_{2}) satisfy fN(z(U1))=fN(z(U2))=wf^{N}(z(U_{1}))=f^{N}(z(U_{2}))=w. Then, we necessarily have U1U2=U_{1}\cap U_{2}=\emptyset, as both U1U_{1} and U2U_{2} correspond to an inverse branch of fNf^{N} defined on a subset of B(w,κ)B(w,\kappa) containing ww.

We fix now ν+(f)\nu\in\mathcal{M}^{+}(f) and let ZνZ_{\nu} be as in Definition 2.3. We denote as before by χmin>0\chi_{\min}>0 the smallest Lyapunov exponent of ν\nu. For every 0<ϵχmin0<\epsilon\ll\chi_{\min}, we fix Zν(ϵ)Z_{\nu}(\epsilon), n(ϵ)n(\epsilon), and r(ϵ)r(\epsilon) as given by Corollary 2.4 and Lemma 2.5.

By Theorem 3.2 and Remark 3.3, for ν^\hat{\nu}-almost every x^Zν(ϵ)\hat{x}\in Z_{\nu}(\epsilon) and every 0<κ<r(ϵ)0<\kappa<r(\epsilon) there exist positive integers m1(ϵ,x^)n(ϵ)m_{1}(\epsilon,\hat{x})\geq n(\epsilon) and m2(ϵ,κ)1m_{2}(\epsilon,\kappa)\geq 1 such that the conclusion of Theorem 3.2 holds for Nm1(ϵ,x)+m2(ϵ,κ)N\geq m_{1}(\epsilon,x)+m_{2}(\epsilon,\kappa). For every mm\in\mathbb{N}, consider the set Zν(ϵ,m) . . ={x^Zν(ϵ):n(ϵ)m1(ϵ,x^)m}Z_{\nu}(\epsilon,m)\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\{\hat{x}\in Z_{\nu}(\epsilon)\colon n(\epsilon)\leq m_{1}(\epsilon,\hat{x})\leq m\}. Since ν^(Zν(ϵ)Zν(ϵ,m))0\hat{\nu}(Z_{\nu}(\epsilon)\setminus Z_{\nu}(\epsilon,m))\to 0 as mm\to\infty, for every 0<ϵχmin0<\epsilon\ll\chi_{\min} there exists m(ϵ)n(ϵ)m(\epsilon)\geq n(\epsilon) such that ν(π(Zν(ϵ))π(Zν(ϵ,m(ϵ))))<ϵ\nu\left(\pi(Z_{\nu}(\epsilon))\setminus\pi(Z_{\nu}(\epsilon,m(\epsilon)))\right)<\epsilon. For every 0<ϵχmin0<\epsilon\ll\chi_{\min}, we define

(4.1) Zν(ϵ) . . =Zν(ϵ,m(ϵ)) and Zν . . =0<ϵχminZν(ϵ).Z^{\star}_{\nu}(\epsilon)\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=Z_{\nu}(\epsilon,m(\epsilon))\quad\mbox{ and }\quad Z^{\star}_{\nu}\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\cup_{0<\epsilon\ll\chi_{\min}}Z_{\nu}^{\star}(\epsilon).

By definition, the conclusion of Theorem 3.2 holds for every xπ(Zν(ϵ))x\in\pi(Z_{\nu}^{\star}(\epsilon)), with m1(ϵ,x)m_{1}(\epsilon,x) independent of xx. This fact will not be used in this subsection, but will be crucial in the proof of Proposition 4.26. Observe also that ν^(Zν)=1\hat{\nu}(Z_{\nu}^{\star})=1 and ν(π(Zν))=1\nu(\pi(Z_{\nu}^{\star}))=1.

Remark 4.2.

As in Remark 2.10, when XX is uniformly expanding, for every νX+(f)\nu\in\mathcal{M}_{X}^{+}(f) we can assume that Zν=Zν=Zν(ϵ)Z_{\nu}=Z^{\star}_{\nu}=Z_{\nu}^{\star}(\epsilon) for all 0<ϵχmin0<\epsilon\ll\chi_{\min}, and that π(Zν)=X\pi(Z_{\nu})=X.

We first fix 0<ϵχmin0<\epsilon\ll\chi_{\min} and define a quantity VDνϵ(Y)\operatorname{VD}_{\nu}^{\epsilon}(Y) for every subset Yπ(Zν(ϵ))Y\subseteq\pi(Z^{\star}_{\nu}(\epsilon)). The definition will depend on both ff and ν\nu.

For 0<κ<r(ϵ)0<\kappa<r(\epsilon) and Nn(ϵ)\mathbb{N}\ni N^{\star}\geq n(\epsilon), we denote by 𝒰(ϵ,κ,N)\mathcal{U}(\epsilon,\kappa,N^{\star}) the collection of open sets

(4.2) 𝒰(ϵ,κ,N) . . =NN𝒰Nκ(π(Zν(ϵ)),ϵ).\mathcal{U}(\epsilon,\kappa,N^{\star})\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\bigcup_{N\geq N^{\star}}\mathcal{U}^{\kappa}_{N}(\pi(Z^{\star}_{\nu}(\epsilon)),\epsilon).
Lemma 4.3.

For every 0<ϵχmin0<\epsilon\ll\chi_{\min}, 0<κ<r(ϵ)0<\kappa<r(\epsilon) and Nn(ϵ)\mathbb{N}\ni N^{\star}\geq n(\epsilon), the collection 𝒰(ϵ,κ,N)\mathcal{U}(\epsilon,\kappa,N^{\star}) is an open cover of π(Zν(ϵ))\pi(Z^{\star}_{\nu}(\epsilon)).

Proof.

It follows from Lemma 2.8 and the fact that Zν(ϵ)Zν(ϵ)Z^{\star}_{\nu}(\epsilon)\subseteq Z_{\nu}(\epsilon) that U(N,x,κ,ϵ)U(N,x,\kappa,\epsilon) is well-defined (and is an open neighbourhood of xx) for all xπ(Zν(ϵ))x\in\pi(Z^{\star}_{\nu}(\epsilon)), 0<κ<r(ϵ)0<\kappa<r(\epsilon), and Nn(ϵ)N\geq n(\epsilon). The assertion follows. ∎

Definition 4.4.

For every 0<ϵχmin0<\epsilon\ll\chi_{\min}, Yπ(Zν(ϵ))Y\subseteq\pi(Z^{\star}_{\nu}(\epsilon)), and 0<κ<r(ϵ)0<\kappa<r(\epsilon), an (ϵ,κ)(\epsilon,\kappa)-cover of YY is a countable cover {Ui}i1\{U_{i}\}_{i\geq 1} of YY with the property that Ui𝒰(ϵ,κ,n(ϵ))U_{i}\in\mathcal{U}(\epsilon,\kappa,n(\epsilon)) for all ii. An ϵ\epsilon-cover is an (ϵ,κ)(\epsilon,\kappa)-cover for some 0<κ<r(ϵ)0<\kappa<r(\epsilon).

For every α0\alpha\geq 0 and Yπ(Zν(ϵ))Y\subseteq\pi(Z^{\star}_{\nu}(\epsilon)), we define Λαϵ(Y)[0,+]\Lambda^{\epsilon}_{\alpha}(Y)\in[0,+\infty] as

(4.3) Λαϵ(Y) . . =lim supκ0Λαϵ,κ(Y),whereΛαϵ,κ(Y) . . =limNinf{Ui}i1Vol(Ui)α\Lambda^{\epsilon}_{\alpha}(Y)\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\limsup_{\kappa\to 0}\Lambda^{\epsilon,\kappa}_{\alpha}(Y),\quad\mbox{where}\quad\Lambda^{\epsilon,\kappa}_{\alpha}(Y)\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\lim_{N^{\star}\to\infty}\inf_{\{U_{i}\}}\sum_{i\geq 1}\operatorname{Vol}(U_{i})^{\alpha}

and the infimum in the second expression is taken over all (ϵ,κ)(\epsilon,\kappa)-covers {Ui}i1\{U_{i}\}_{i\geq 1} of YY with Ui𝒰(ϵ,κ,N)U_{i}\in\mathcal{U}(\epsilon,\kappa,N^{\star}) for all i1i\geq 1. Observe that the limit in the second expression above is well-defined, and equal to a supremum over Nn(ϵ)N^{\star}\geq n(\epsilon), as the 𝒰(ϵ,κ,N)\mathcal{U}(\epsilon,\kappa,N^{\star})’s are decreasing collections of covers for NN^{\star}\to\infty. We will see below that the function αΛαϵ,κ(Y)\alpha\mapsto\Lambda_{\alpha}^{\epsilon,\kappa}(Y) is essentially independent of κ\kappa; see Lemma 4.11. Hence we will be able to use this approximated version of Λαϵ(Y)\Lambda^{\epsilon}_{\alpha}(Y) in order to study its properties.

Lemma 4.5.

For every 0<ϵχmin0<\epsilon\ll\chi_{\min}, 0<κ<r(ϵ)0<\kappa<r(\epsilon), and Yπ(Zν(ϵ))Y\subseteq\pi(Z^{\star}_{\nu}(\epsilon)), the following assertions hold:

  1. (1)

    the functions αΛαϵ,κ(Y)\alpha\mapsto\Lambda^{\epsilon,\kappa}_{\alpha}(Y) and αΛαϵ(Y)\alpha\mapsto\Lambda^{\epsilon}_{\alpha}(Y) are non-increasing;

  2. (2)

    if Λα0ϵ,κ(Y)<\Lambda^{\epsilon,\kappa}_{\alpha_{0}}(Y)<\infty (resp. Λα0ϵ(Y)<\Lambda^{\epsilon}_{\alpha_{0}}(Y)<\infty) for some α00\alpha_{0}\geq 0, then Λαϵ,κ(Y)=0\Lambda^{\epsilon,\kappa}_{\alpha}(Y)=0 (resp. Λαϵ(Y)=0\Lambda^{\epsilon}_{\alpha}(Y)=0) for all α>α0\alpha>\alpha_{0}.

Proof.

By the definition (4.3) of Λαϵ(Y)\Lambda_{\alpha}^{\epsilon}(Y) and Λαϵ,κ(Y)\Lambda_{\alpha}^{\epsilon,\kappa}(Y), it is enough to show the two assertions for Λαϵ,κ(Y)\Lambda^{\epsilon,\kappa}_{\alpha}(Y) for a given κ\kappa as in the statement.

The first property is clear from the definition of Λαϵ,κ(Y)\Lambda^{\epsilon,\kappa}_{\alpha}(Y) and the fact that, up to taking NN^{\star} sufficiently large, we can assume that the volume of all the UiU_{i}’s is less than 1 in the definition of Λαϵ,κ(Y)\Lambda^{\epsilon,\kappa}_{\alpha}(Y); see Lemma 2.8 (2). If α1<α2\alpha_{1}<\alpha_{2}, then, for every η>0\eta>0 and up to taking NN^{\star} sufficiently large, for every U𝒰(ϵ,κ,N)U\in\mathcal{U}(\epsilon,\kappa,N^{\star}) we also have

Vol(U)α2=Vol(U)α1+(α2α1)η(α2α1)Vol(U)α1.\operatorname{Vol}(U)^{\alpha_{2}}=\operatorname{Vol}(U)^{\alpha_{1}+(\alpha_{2}-\alpha_{1})}\leq\eta^{(\alpha_{2}-\alpha_{1})}\cdot\operatorname{Vol}(U)^{\alpha_{1}}.

As η\eta can be taken arbitrarily small and Λα1ϵ,κ(Y)\Lambda^{\epsilon,\kappa}_{\alpha_{1}}(Y) is finite, this gives Λα2ϵ,κ(Y)=0\Lambda^{\epsilon,\kappa}_{\alpha_{2}}(Y)=0. The assertion follows. ∎

Because of Lemma 4.5, the following definition is well-posed.

Definition 4.6.

For every 0<ϵχmin0<\epsilon\ll\chi_{\min} and Yπ(Zν(ϵ))Y\subseteq\pi(Z^{\star}_{\nu}(\epsilon)), we set

VDνϵ(Y) . . =sup{α:Λαϵ(Y)=}=inf{α:Λαϵ(Y)=0}.\operatorname{VD}^{\epsilon}_{\nu}(Y)\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\sup\{\alpha:\Lambda^{\epsilon}_{\alpha}(Y)=\infty\}=\inf\{\alpha:\Lambda^{\epsilon}_{\alpha}(Y)=0\}.

Similarly, for every 0<κ<r(ϵ)0<\kappa<r(\epsilon), we also set

VDνϵ,κ(Y) . . =sup{α:Λαϵ,κ(Y)=}=inf{α:Λαϵ,κ(Y)=0}.\operatorname{VD}^{\epsilon,\kappa}_{\nu}(Y)\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\sup\{\alpha:\Lambda^{\epsilon,\kappa}_{\alpha}(Y)=\infty\}=\inf\{\alpha:\Lambda^{\epsilon,\kappa}_{\alpha}(Y)=0\}.
Remark 4.7.

The definition of VDνϵ,κ(Y)\operatorname{VD}^{\epsilon,\kappa}_{\nu}(Y) will not be needed in this section, but Lemma 4.12 will be used in the proof of Proposition 5.4.

Lemma 4.8.

For every 0<ϵχmin0<\epsilon\ll\chi_{\min}, 0<κ<r(ϵ)0<\kappa<r(\epsilon), and Y1Y2π(Zν(ϵ))Y_{1}\subseteq Y_{2}\subseteq\pi(Z^{\star}_{\nu}(\epsilon)), we have

VDνϵ,κ(Y1)VDνϵ,κ(Y2) and VDνϵ(Y1)VDνϵ(Y2).\operatorname{VD}_{\nu}^{\epsilon,\kappa}(Y_{1})\leq\operatorname{VD}_{\nu}^{\epsilon,\kappa}(Y_{2})\quad\mbox{ and }\quad\operatorname{VD}_{\nu}^{\epsilon}(Y_{1})\leq\operatorname{VD}_{\nu}^{\epsilon}(Y_{2}).
Proof.

For every 0<κ<r(ϵ)0<\kappa<r(\epsilon), every (ϵ,κ)(\epsilon,\kappa)-cover of Y2Y_{2} is also an (ϵ,κ)(\epsilon,\kappa)-cover of Y1Y_{1}. The assertion follows. ∎

In the next lemma, we will use the following form of Besicovitch’s covering theorem [Bes45].

Theorem 4.9.

Let n1n\geq 1 be an integer. There exists a constant b(n)>0b(n)>0 such that the following claim is true. If AA is a bounded subset of n\mathbb{R}^{n}, then for any function r:A(0,)r\colon A\to(0,\infty) there exists a countable subset {xm:m}\{x_{m}\colon m\in\mathbb{N}\} of AA such that the collection of open balls (A,r) . . ={B(xm,r(xm)):m}\mathcal{B}(A,r)\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\{B(x_{m},r(x_{m}))\colon m\in\mathbb{N}\} covers AA and can be decomposed into b(n)b(n) families whose elements are disjoint.

Lemma 4.10.

For every 0<ϵχmin0<\epsilon\ll\chi_{\min}, 0<κ<r(ϵ)0<\kappa<r(\epsilon), and Yπ(Zν(ϵ))Y\subseteq\pi(Z_{\nu}^{\star}(\epsilon)), we have

VDνϵ(Y)1 and VDνϵ,κ(Y)1.\operatorname{VD}^{\epsilon}_{\nu}(Y)\leq 1\quad\mbox{ and }\quad\operatorname{VD}^{\epsilon,\kappa}_{\nu}(Y)\leq 1.
Proof.

By Lemma 4.8, it is enough to show the statement for Y=π(Zν(ϵ))Y=\pi(Z^{\star}_{\nu}(\epsilon)). By (4.3) and the Definition 4.6 of VDνϵ(Y)\operatorname{VD}^{\epsilon}_{\nu}(Y) and VDνϵ,κ(Y)\operatorname{VD}^{\epsilon,\kappa}_{\nu}(Y), it is enough to show that, for any 0<κ<r(ϵ)0<\kappa<r(\epsilon) and N>n(ϵ)N^{\star}>n(\epsilon), there exists an (ϵ,κ)(\epsilon,\kappa)-cover {Ui}\{U_{i}\} of π(Zν(ϵ))\pi(Z_{\nu}^{\star}(\epsilon)) with N(Ui)NN(U_{i})\geq N^{\star} for all i1i\geq 1 and such that

iVol(Ui)<C<\sum_{i}\operatorname{Vol}(U_{i})<C<\infty

for some constant CC independent of κ\kappa. Indeed, by (4.3), this shows that Λ1ϵ,κ(π(Zν(ϵ)))<\Lambda_{1}^{\epsilon,\kappa}(\pi(Z^{\star}_{\nu}(\epsilon)))<\infty, which implies that VDνϵ,κ(π(Zν(ϵ)))1\operatorname{VD}_{\nu}^{\epsilon,\kappa}(\pi(Z^{\star}_{\nu}(\epsilon)))\leq 1. As CC is independent of κ\kappa, this also gives Λ1ϵ(π(Zν(ϵ)))<\Lambda_{1}^{\epsilon}(\pi(Z^{\star}_{\nu}(\epsilon)))<\infty and thus VDνϵ(π(Zν(ϵ)))1\operatorname{VD}_{\nu}^{\epsilon}(\pi(Z^{\star}_{\nu}(\epsilon)))\leq 1, as desired.

Fix Nn(ϵ)N^{\star}\geq n(\epsilon) and 0<κ<r(ϵ)0<\kappa<r(\epsilon). Theorem 4.9 applied with n=2kn=2k, A . . =fN(π(Zν(ϵ)))A\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=f^{N^{\star}}(\pi(Z^{\star}_{\nu}(\epsilon))), and rκeNMϵr\equiv\kappa\,e^{-N^{\star}M\epsilon}, gives b(2k)b(2k) collections j\mathcal{B}_{j}, 1jb(2k)1\leq j\leq b(2k), of disjoint open balls {Bj,l}l1\{B_{j,l}\}_{l\geq 1} centred on fN(π(Zν(ϵ)))f^{N^{\star}}(\pi(Z^{\star}_{\nu}(\epsilon))) and of radius rr such that Aj,lBj,lA\subset\cup_{j,l}B_{j,l}. We work here in local charts; see also Remark 1.4.

Consider an element Bj,lB_{j^{\star},l^{\star}} of the collection {Bj,l}j,l\{B_{j,l}\}_{j,l}. By construction, its center belongs to fN(Zν(ϵ))f^{N^{\star}}(Z^{\star}_{\nu}(\epsilon)). Denote by xj,l1,,xj,lm(j,l)x_{j^{\star},l^{\star}}^{1},\dots,x_{j^{\star},l^{\star}}^{m(j^{\star},l^{\star})} the preimages by fNf^{N^{\star}} of the center of Bj,lB_{j^{\star},l^{\star}} which belong to π(Zν(ϵ))\pi(Z^{\star}_{\nu}(\epsilon)). For each 1qm(j,l)1\leq q\leq m(j^{\star},l^{\star}), choose an orbit x^=x^(j,l,q)Zν(ϵ)\hat{x}=\hat{x}(j^{\star},l^{\star},q)\in Z_{\nu}^{\star}(\epsilon) such that π0(x^(j,l,q))=xj,lq\pi_{0}(\hat{x}(j^{\star},l^{\star},q))=x_{j^{\star},l^{\star}}^{q} (observe that πN(x^(j,l,q))\pi_{N^{\star}}(\hat{x}(j^{\star},l^{\star},q)) is necessarily the center of Bj,lB_{j^{\star},l^{\star}}). The inverse branch fTN(x^(j,l,q))Nf_{T^{N^{\star}}(\hat{x}(j^{\star},l^{\star},q))}^{-N^{\star}} is well-defined on the ball Bj,lB_{j^{\star},l^{\star}} by the choice of the function rr and Lemma 2.8. More precisely, the image of each Bj,lB_{j,l} under any of such branches is of the form U(N,x,κ,ϵ)U(N^{\star},x,\kappa,\epsilon) for some xπ(Zν(ϵ))x\in\pi(Z^{\star}_{\nu}(\epsilon)). The images associated to the same Bj,lB_{j,l} are disjoint; see Remark 4.1. Similarly, any two such images are also disjoint whenever the corresponding balls Bj0,l0B_{j_{0},l_{0}} and Bj1,l1B_{j_{1},l_{1}} are disjoint. Observe that this in particular applies whenever j0=j1j_{0}=j_{1}, since each collection j\mathcal{B}_{j} consists of disjoint balls.

By construction, we have

π(Zν(ϵ))j=1b(2k)l1q=1m(j,l)fTN(x^(j,l,q))N(Bj,l).\pi(Z_{\nu}^{\star}(\epsilon))\subseteq\bigcup_{j=1}^{b(2k)}\bigcup_{l\geq 1}\bigcup_{q=1}^{m(j,l)}f_{T^{N^{\star}}({\hat{x}}(j,l,q))}^{-N^{\star}}(B_{j,l}).

By the arguments above, we have

j=1b(2k)l11qm(j,l)Vol(fTN(x^(j,l,q))N(Bj,l))Cb(2k)Vol(k),\sum_{j=1}^{b(2k)}\sum_{l\geq 1}\sum_{1\leq q\leq m(j,l)}\operatorname{Vol}\big{(}f_{T^{N^{\star}}({\hat{x}}(j,l,q))}^{-N^{\star}}(B_{j,l})\big{)}\leq C^{\prime}b(2k)\operatorname{Vol}(\mathbb{P}^{k}),

where the positive constant CC^{\prime} is due to the use of local charts, and is in particular independent of κ\kappa. This completes the proof. ∎

Observe that, for every 0<γ<10<\gamma<1, there exists a positive integer θ=θγ\theta=\theta_{\gamma} with the property that it is possible to cover any ball of radius rr in k\mathbb{P}^{k} with a finite number θ\theta of open balls of radius γr\gamma r (the constant θ\theta also depends on the dimension kk, but we omit this dependence since kk is fixed).

For every 0<ϵχmin0<\epsilon\ll\chi_{\min}, define also the constant

(4.4) (ϵ) . . =Lν+k(2M+1)ϵLνkϵ.\ell(\epsilon)\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\frac{L_{\nu}+k(2M+1)\epsilon}{L_{\nu}-k\epsilon}.

Observe that (ϵ)>1\ell(\epsilon)>1 for all ϵ\epsilon as above, and we have (ϵ)=1+O(ϵ)\ell(\epsilon)=1+O(\epsilon) for ϵ0\epsilon\to 0.

Lemma 4.11.

Fix 0<ϵχmin0<\epsilon\ll\chi_{\min}, 0<κ1<κ2<r(ϵ)/30<\kappa_{1}<\kappa_{2}<r(\epsilon)/3, and Yπ(Zν(ϵ))Y\subseteq\pi(Z^{\star}_{\nu}(\epsilon)). For every α0\alpha\geq 0, we have

(4.5) (κ1/κ2(ϵ))2kαΛα(ϵ)ϵ,κ2(Y)Λαϵ,κ1(Y)θ3κ2/κ1(κ1(ϵ)/κ2)2kαΛα(ϵ)1ϵ,κ2(Y),\left(\kappa_{1}/\kappa_{2}^{\ell(\epsilon)}\right)^{2k\alpha}\Lambda_{\alpha\ell(\epsilon)}^{\epsilon,\kappa_{2}}(Y)\leq\Lambda_{\alpha}^{\epsilon,\kappa_{1}}(Y)\leq\theta_{3\kappa_{2}/\kappa_{1}}\left(\kappa_{1}^{\ell(\epsilon)}/\kappa_{2}\right)^{2k\alpha}\Lambda_{\alpha\ell(\epsilon)^{-1}}^{\epsilon,\kappa_{2}}(Y),

where (ϵ)>1\ell(\epsilon)>1 is as in (4.4) and θ3κ2/κ1>0\theta_{3\kappa_{2}/\kappa_{1}}>0 is as above. In particular, for every 0<ϵχmin0<\epsilon\ll\chi_{\min}, Yπ(Zν(ϵ))Y\subseteq\pi(Z_{\nu}^{\star}(\epsilon)), and 0<κ1<κ2<r(ϵ)/30<\kappa_{1}<\kappa_{2}<r(\epsilon)/3, we have

(4.6) (ϵ)1VDνϵ,κ2(Y)VDνϵ,κ1(Y)(ϵ)VDνϵ,κ2(Y).\ell(\epsilon)^{-1}\operatorname{VD}^{\epsilon,\kappa_{2}}_{\nu}(Y)\leq\operatorname{VD}^{\epsilon,\kappa_{1}}_{\nu}(Y)\leq\ell(\epsilon)\operatorname{VD}^{\epsilon,\kappa_{2}}_{\nu}(Y).
Proof.

We first show the inequality (κ1/κ2(ϵ))2kΛα(ϵ)ϵ,κ2(Y)Λαϵ,κ1(Y)(\kappa_{1}/\kappa_{2}^{\ell(\epsilon)})^{2k}\Lambda_{\alpha\ell(\epsilon)}^{\epsilon,\kappa_{2}}(Y)\leq\Lambda_{\alpha}^{\epsilon,\kappa_{1}}(Y). We can assume that Λαϵ,κ1(Y)<\Lambda_{\alpha}^{\epsilon,\kappa_{1}}(Y)<\infty. Fix η\eta such that Λαϵ,κ1(Y)<η<\Lambda_{\alpha}^{\epsilon,\kappa_{1}}(Y)<\eta<\infty. There exists an (ϵ,κ1)(\epsilon,\kappa_{1})-cover {Ui}i1\{U_{i}\}_{i\geq 1} of YY, with each UiU_{i} of the form Ui=U(Ni,xi,κ1,ϵ)U_{i}=U(N_{i},x_{i},\kappa_{1},\epsilon), such that

i1Vol(Ui)α<η.\sum_{i\geq 1}\operatorname{Vol}(U_{i})^{\alpha}<\eta.

For each i1i\geq 1, set Vi . . =U(Ni,xi,κ2,ϵ)V_{i}\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=U(N_{i},x_{i},\kappa_{2},\epsilon). Since κ1<κ2<r(ϵ)\kappa_{1}<\kappa_{2}<r(\epsilon), we have UiViU_{i}\subset V_{i} for all ii, and the collection {Vi}i1\{V_{i}\}_{i\geq 1} is an (ϵ,κ2)(\epsilon,\kappa_{2})-cover of YY. Moreover, by Lemma 2.8 (2) and the definition of (ϵ)\ell(\epsilon), for every ii we have

Vol(Vi)(ϵ)Vol(Ui)κ22k(ϵ)κ12ke2Ni(Lνkϵ)(ϵ)e2Ni(Lν+k(2M+1)ϵ)=(κ2(ϵ)/κ1)2k.\frac{\operatorname{Vol}(V_{i})^{\ell(\epsilon)}}{\operatorname{Vol}(U_{i})}\leq\frac{\kappa_{2}^{2k\ell(\epsilon)}}{\kappa_{1}^{2k}}\frac{e^{-2N_{i}(L_{\nu}-k\epsilon)\ell(\epsilon)}}{e^{-2N_{i}(L_{\nu}+k(2M+1)\epsilon)}}=\left(\kappa_{2}^{\ell(\epsilon)}/\kappa_{1}\right)^{2k}.

It follows that

i1Vol(Vi)α(ϵ)<(κ2(ϵ)/κ1)2kαη.\sum_{i\geq 1}\operatorname{Vol}(V_{i})^{\alpha\ell(\epsilon)}<\left(\kappa_{2}^{\ell(\epsilon)}/\kappa_{1}\right)^{2k\alpha}\eta.

By the choice of η\eta, this shows that Λα(ϵ)ϵ,κ2(Y)(κ2(ϵ)/κ1)2kαΛαϵ,κ1(Y)\Lambda^{\epsilon,\kappa_{2}}_{\alpha\ell(\epsilon)}(Y)\leq(\kappa_{2}^{\ell(\epsilon)}/\kappa_{1})^{2k\alpha}\Lambda_{\alpha}^{\epsilon,\kappa_{1}}(Y), as desired. By the definition of VDνϵ,κ(Y)\operatorname{VD}_{\nu}^{\epsilon,\kappa}(Y), this also shows the first inequality in (4.6).

We now show the second inequality in (4.5). As above, this also shows the second inequality in (4.6). We can assume that Λαϵ,κ2(Y)<\Lambda_{\alpha}^{\epsilon,\kappa_{2}}(Y)<\infty. Fix η\eta such that Λαϵ,κ2(Y)<η<\Lambda_{\alpha}^{\epsilon,\kappa_{2}}(Y)<\eta<\infty. There exists an (ϵ,κ2)(\epsilon,\kappa_{2})-cover {Ui}i1\{U_{i}\}_{i\geq 1} of YY, with each UiU_{i} of the form Ui=U(Ni,xi,κ2,ϵ)U_{i}=U(N_{i},x_{i},\kappa_{2},\epsilon), such that

i1Vol(Ui)α<η.\sum_{i\geq 1}\operatorname{Vol}(U_{i})^{\alpha}<\eta.

By Definition 2.7, for each i1i\geq 1, we have Ai . . =fNi(Ui)=B(fNi(xi),κ2eNiMϵ)A_{i}\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=f^{N_{i}}(U_{i})=B(f^{N_{i}}(x_{i}),\kappa_{2}\,e^{-N_{i}M\epsilon}). By the definition of θ3κ2/κ1\theta_{3\kappa_{2}/\kappa_{1}}, one can cover any ball AiA_{i} with θ3κ2/κ1\theta_{3\kappa_{2}/\kappa_{1}} open balls of radius (κ1/3)eNiMϵ(\kappa_{1}/3)e^{-N_{i}M\epsilon}. In particular, these balls cover fNi(YUi)Aif^{N_{i}}(Y\cap U_{i})\subseteq A_{i}. Up to removing from the collection the balls not intersecting fNi(YUif^{N_{i}}(Y\cap U_{i}) and replacing all the other balls with balls of the same center and radius κ1eNiMϵ\kappa_{1}e^{-N_{i}M\epsilon}, we see that we can cover fNi(YUi)f^{N_{i}}(Y\cap U_{i}) with θ3κ2/κ1\theta_{3\kappa_{2}/\kappa_{1}} balls of radius κ1eNiMϵ\kappa_{1}e^{-N_{i}M\epsilon} and centred at points of fNi(YUi)f^{N_{i}}(Y\cap U_{i}). For every ii, we denote by {Bi,j}1jJi\{B_{i,j}\}_{1\leq j\leq J_{i}} (for some 1Jiθ3κ2/κ11\leq J_{i}\leq\theta_{3\kappa_{2}/\kappa_{1}}), the collection of the balls of radius κ1eNiMϵ\kappa_{1}\,e^{-N_{i}M\epsilon} constructed above. By construction, for every ii we have

fNi(YUi)j=1JiBi,j.f^{N_{i}}(Y\cap U_{i})\subseteq\bigcup_{j=1}^{J_{i}}B_{i,j}.

For every i,ji,j, set Vi,j . . =gi(Bi,j)V_{i,j}\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=g_{i}(B_{i,j}), where gig_{i} is the inverse branch of fNif^{N_{i}} defined in a neighbourhood of fNi(zi)f^{N_{i}}(z_{i}) that sends fNi(zi)f^{N_{i}}(z_{i}) to ziz_{i}. It follows that, for each ii, the collection {Vi,j}1jJi\{V_{i,j}\}_{1\leq j\leq J_{i}} is an (ϵ,κ1)(\epsilon,\kappa_{1})-cover of UiYU_{i}\cap Y. By Lemma 2.8 (2) and the definition of (ϵ)\ell(\epsilon), for every ii and 1jJi1\leq j\leq J_{i} we have

Vol(Vi,j)(ϵ)Vol(Ui)κ12k(ϵ)κ22ke2Ni(Lνkϵ)(ϵ)e2Ni(Lν+k(2M+1)ϵ)=(κ1(ϵ)/κ2)2k.\frac{\operatorname{Vol}(V_{i,j})^{\ell(\epsilon)}}{\operatorname{Vol}(U_{i})}\leq\frac{\kappa_{1}^{2k\ell(\epsilon)}}{\kappa_{2}^{2k}}\frac{e^{-2N_{i}(L_{\nu}-k\epsilon)\ell(\epsilon)}}{e^{-2N_{i}(L_{\nu}+k(2M+1)\epsilon)}}=\left({\kappa_{1}^{\ell(\epsilon)}}/{\kappa_{2}}\right)^{2k}.

Summing over ii and jj, we obtain

i1j=1JiVol(Vi,j)α(ϵ)\displaystyle\sum_{i\geq 1}\sum_{j=1}^{J_{i}}\operatorname{Vol}(V_{i,j})^{\alpha\ell(\epsilon)} i1(θ3κ2/κ1(κ1(ϵ)/κ2)2kαVol(Ui)α)\displaystyle\leq\sum_{i\geq 1}\left(\theta_{3\kappa_{2}/\kappa_{1}}\left({\kappa_{1}^{\ell(\epsilon)}}/{\kappa_{2}}\right)^{2k\alpha}\operatorname{Vol}(U_{i})^{\alpha}\right)
θ3κ2/κ1(κ1(ϵ)/κ2)2kαη.\displaystyle\leq\theta_{3\kappa_{2}/\kappa_{1}}\left({\kappa_{1}^{\ell(\epsilon)}}/{\kappa_{2}}\right)^{2k\alpha}\eta.

It follows that Λα(ϵ)ϵ,κ1(Y)θ3κ2/κ1(κ1(ϵ)/κ2)2kαΛαϵ,κ2(Y)\Lambda^{\epsilon,\kappa_{1}}_{\alpha\ell(\epsilon)}(Y)\leq\theta_{3\kappa_{2}/\kappa_{1}}(\kappa_{1}^{\ell(\epsilon)}/\kappa_{2})^{2k\alpha}\Lambda_{\alpha}^{\epsilon,\kappa_{2}}(Y), as desired. This completes the proof. ∎

Lemma 4.12.

For every 0<ϵχmin0<\epsilon\ll\chi_{\min}, 0<κ<r(ϵ)/30<\kappa<r(\epsilon)/3, and Yπ(Zν(ϵ))Y\subseteq\pi(Z^{\star}_{\nu}(\epsilon)), we have

|VDνϵ,κ(Y)VDνϵ(Y)|((ϵ)1)min{VDνϵ(Y),VDνϵ,κ(Y)}(ϵ)1,|\operatorname{VD}^{\epsilon,\kappa}_{\nu}(Y)-\operatorname{VD}^{\epsilon}_{\nu}(Y)|\leq(\ell(\epsilon)-1)\min\big{\{}\operatorname{VD}^{\epsilon}_{\nu}(Y),\operatorname{VD}^{\epsilon,\kappa}_{\nu}(Y)\big{\}}\leq\ell(\epsilon)-1,

where (ϵ)>1\ell(\epsilon)>1 is as in (4.4).

Proof.

Fix 0<ϵχmin0<\epsilon\ll\chi_{\min} and Yπ(Zν(ϵ))Y\subseteq\pi(Z^{\star}_{\nu}(\epsilon)). It follows from Lemma 4.11 that there exists α0=α0(ϵ,Y)[0,+]\alpha_{0}=\alpha_{0}(\epsilon,Y)\in[0,+\infty] such that

(4.7) α0VDνϵ,κ(Y)α0(ϵ) for every 0<κ<r(ϵ).\alpha_{0}\leq\operatorname{VD}^{\epsilon,\kappa}_{\nu}(Y)\leq\alpha_{0}\ell(\epsilon)\quad\mbox{ for every }0<\kappa<r(\epsilon).

Take any η>0\eta>0. It follows from the definition (4.3) of Λαϵ(Y)\Lambda_{\alpha}^{\epsilon}(Y) and Λαϵ,κ(Y)\Lambda_{\alpha}^{\epsilon,\kappa}(Y) that Λα0ηϵ(Y)=+\Lambda_{\alpha_{0}-\eta}^{\epsilon}(Y)=+\infty (assuming α0>0\alpha_{0}>0 and 0<η<α00<\eta<\alpha_{0}) and Λα0(ϵ)+ηϵ(Y)=0\Lambda_{\alpha_{0}\ell(\epsilon)+\eta}^{\epsilon}(Y)=0 (assuming α0<\alpha_{0}<\infty). Since η\eta is arbitrary, we deduce from the definition of VDνϵ(Y)\operatorname{VD}^{\epsilon}_{\nu}(Y) that

(4.8) α0VDνϵ(Y)α0(ϵ).\alpha_{0}\leq\operatorname{VD}^{\epsilon}_{\nu}(Y)\leq\alpha_{0}\ell(\epsilon).

The first inequality in the statement follows from (4.7) and (4.8). The second one follows from Lemma 4.10. ∎

Remark 4.13.

In Definition 4.4, we do not require z(Ui)Yz(U_{i})\in Y for any of the UiU_{i} in an ϵ\epsilon-cover of YY, but we could also define VDνϵ(Y)\operatorname{VD}_{\nu}^{\epsilon}(Y) (and VDνϵ,κ(Y)\operatorname{VD}_{\nu}^{\epsilon,\kappa}(Y)) for Yπ(Zν(ϵ))Y\subseteq\pi(Z^{\star}_{\nu}(\epsilon)) by only using sets UiU_{i} such that z(Ui)Yz(U_{i})\in Y, rather than z(Ui)π(Zν(ϵ))z(U_{i})\in\pi(Z^{\star}_{\nu}(\epsilon)), in the definition of Λαϵ,κ(Y)\Lambda^{\epsilon,\kappa}_{\alpha}(Y). Denoting by Λ¯αϵ(Y)\overline{\Lambda}^{\epsilon}_{\alpha}(Y) and VD¯νϵ(Y)\overline{\operatorname{VD}}^{\epsilon}_{\nu}(Y) the corresponding quantities, it is straightforward to see that Λαϵ(Y)Λ¯αϵ(Y)\Lambda^{\epsilon}_{\alpha}(Y)\leq\overline{\Lambda}^{\epsilon}_{\alpha}(Y) for all α0\alpha\geq 0. Hence, we have VDνϵ(Y)VD¯νϵ(Y)\operatorname{VD}_{\nu}^{\epsilon}(Y)\leq\overline{\operatorname{VD}}_{\nu}^{\epsilon}(Y). On the other hand, take α\alpha such that Λαϵ(Y)<\Lambda^{\epsilon}_{\alpha}(Y)<\infty and consider an (ϵ,κ)(\epsilon,\kappa)-cover {Ui}i1\{U_{i}\}_{i\geq 1} of YY for which the value of i1Vol(Ui)α\sum_{i\geq 1}\operatorname{Vol}(U_{i})^{\alpha} is close to the value of Λαϵ(Y)\Lambda^{\epsilon}_{\alpha}(Y). By the definition of Λαϵ(Y)\Lambda^{\epsilon}_{\alpha}(Y), we can assume that, for all ii, we have Ui𝒰(ϵ,κ,N0)U_{i}\in\mathcal{U}(\epsilon,\kappa,N_{0}) for some κ<r(ϵ)/3\kappa<r(\epsilon)/3 and some N0n(ϵ)N_{0}\geq n(\epsilon). Take i0i_{0} such that z(Ui0)Yz(U_{i_{0}})\notin Y. Observe that there must exist yUi0Yy\in U_{i_{0}}\cap Y and that, since κ<r(ϵ)/3\kappa<r(\epsilon)/3, the set U(N(Ui0),y,3κ,ϵ)U(N(U_{i_{0}}),y,3\kappa,\epsilon) is well-defined and contains Ui0U_{i_{0}}. By similar arguments as in the proof of Lemma 4.11, this shows that Λ¯α/(ϵ)ϵ,3κ(Y)32kα/(ϵ)Λαϵ,κ(Y)\overline{\Lambda}^{\epsilon,3\kappa}_{\alpha/\ell(\epsilon)}(Y)\leq 3^{2k\alpha/\ell(\epsilon)}\Lambda^{\epsilon,\kappa}_{\alpha}(Y) for all 0<κ<r(ϵ)/30<\kappa<r(\epsilon)/3, which gives VD¯νϵ(Y)(ϵ)VDνϵ(Y)\overline{\operatorname{VD}}_{\nu}^{\epsilon}(Y)\leq\ell(\epsilon)\operatorname{VD}_{\nu}^{\epsilon}(Y). Similar arguments and estimates hold for VDνϵ,κ(Y)\operatorname{VD}^{\epsilon,\kappa}_{\nu}(Y).

Take now Xπ(Zν)X\subseteq\pi(Z_{\nu}^{\star}), and recall that ν(π(Zν))=1\nu(\pi(Z_{\nu}^{\star}))=1. For every 0<ϵχmin0<\epsilon\ll\chi_{\min}, we set

(4.9) Xϵ . . =Xπ(Zν(ϵ)).X^{\epsilon}\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=X\cap\pi(Z^{\star}_{\nu}(\epsilon)).

Observe that, since Zν=0<ϵχminZν(ϵ)Z_{\nu}^{\star}=\cup_{0<\epsilon\ll\chi_{\min}}Z^{\star}_{\nu}(\epsilon), we have 0<ϵχminXϵ=X\cup_{0<\epsilon\ll\chi_{\min}}X^{\epsilon}=X.

Definition 4.14.

For every Xπ(Zν)X\subseteq\pi(Z^{\star}_{\nu}), the volume dimension VDν(X)\operatorname{VD}_{\nu}(X) of XX is

VDν(X) . . =lim supϵ0VDνϵ(Xϵ).\operatorname{VD}_{\nu}(X)\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\limsup_{\epsilon\to 0}\operatorname{VD}_{\nu}^{\epsilon}(X^{\epsilon}).
Remark 4.15.

The limsup in Definition 4.14 is actually a limit; see Section 4.2, and in particular Corollary 4.24, and Remark 5.1.

Lemma 4.16.

For every Y1Y2π(Zν)Y_{1}\subseteq Y_{2}\subseteq\pi(Z_{\nu}^{\star}), we have VDν(Y1)VDν(Y2)\operatorname{VD}_{\nu}(Y_{1})\leq\operatorname{VD}_{\nu}(Y_{2}).

Proof.

By Lemma 4.8, for any 0<ϵχmin0<\epsilon\ll\chi_{\min} we have VDνϵ(Y1ϵ)VDνϵ(Y2ϵ)\operatorname{VD}_{\nu}^{\epsilon}(Y^{\epsilon}_{1})\leq\operatorname{VD}_{\nu}^{\epsilon}(Y^{\epsilon}_{2}). The conclusion follows from Definition 4.14. ∎

Definition 4.17.

Take ν+(f)\nu\in\mathcal{M}^{+}(f). The volume dimension VD(ν)\operatorname{VD}(\nu) of ν\nu is

VD(ν) . . =inf{VDν(X):Xπ(Zν),ν(X)=1}.\operatorname{VD}(\nu)\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\inf\big{\{}\operatorname{VD}_{\nu}(X):X\subseteq\pi(Z_{\nu}^{\star}),\nu(X)=1\big{\}}.
Lemma 4.18.

For every ν+(f)\nu\in\mathcal{M}^{+}(f) and Xπ(Zν)X\subseteq\pi(Z_{\nu}^{\star}), we have VDν(X)1\operatorname{VD}_{\nu}(X)\leq 1. In particular, we have VD(ν)1\operatorname{VD}(\nu)\leq 1 for every ν+(f)\nu\in\mathcal{M}^{+}(f).

Proof.

The statement is an immediate consequence of Lemma 4.10 and Definitions 4.14 and 4.17. ∎

When XkX\subseteq\mathbb{P}^{k} is a uniformly expanding closed invariant set for ff, by Remark 4.2 we can assume that π(Zν(ϵ))=π(Zν)=X\pi(Z^{\star}_{\nu}(\epsilon))=\pi(Z^{\star}_{\nu})=X for every invariant measure ν\nu on XX and every 0<ϵχmin0<\epsilon\ll\chi_{\min}. In particular, the following definition is well-posed and defines the term VD(J(f))\operatorname{VD}(J(f)) in Theorem 1.3.

Definition 4.19.

If XkX\subseteq\mathbb{P}^{k} is uniformly expanding, the volume dimension of XX is

VD(X) . . =supνX+(f)VDν(X).\operatorname{VD}(X)\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\sup_{\nu\in\mathcal{M}_{X}^{+}(f)}\operatorname{VD}_{\nu}(X).

We conclude this section with the next proposition, which in particular shows that, when k=1k=1, the volume dimension associated to any ν+(f)\nu\in\mathcal{M}^{+}(f) is equivalent to the Hausdorff dimension.

Proposition 4.20.

If ν+(f)\nu\in\mathcal{M}^{+}(f) is such that all the Lyapunov exponents of ν\nu are equal to χ>0\chi>0, then

  1. (1)

    2kVDν(X)=HD(X)2k\operatorname{VD}_{\nu}(X)=\operatorname{HD}(X) for all Xπ(Zν)X\subseteq\pi(Z^{\star}_{\nu});

  2. (2)

    2kVD(ν)=HD(ν)2k\operatorname{VD}(\nu)=\operatorname{HD}(\nu),

where HD(X)\operatorname{HD}(X) and HD(ν)\operatorname{HD}(\nu) denote the Hausdorff dimension of XX and ν\nu, respectively.

Proof.

Recall that the Hausdorff dimension of XkX\subseteq\mathbb{P}^{k} is defined as

HD(X) . . =inf{α:Hα(X)=0}, where Hα(X) . . =supδ>0inf{Bi}i(diamBi)α.\operatorname{HD}(X)\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\inf\big{\{}\alpha\colon H_{\alpha}(X)=0\big{\}},\quad\mbox{ where }H_{\alpha}(X)\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\sup_{\delta>0}\inf_{\{B_{i}\}}\sum_{i}(\operatorname{diam}B_{i})^{\alpha}.

The infimum in the second expression is taken over all countable covers of XX by open balls {Bi}\{B_{i}\} whose diameter is less than δ\delta. The Hausdorff dimension HD(ν)\operatorname{HD}(\nu) of ν\nu is defined as

(4.10) HD(ν) . . =inf{HD(X):XSupp ν,ν(X)=1}.\operatorname{HD}(\nu)\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\inf\big{\{}\operatorname{HD}(X)\colon X\subseteq\text{Supp }\nu,\,\nu(X)=1\big{\}}.

In order to prove the first assertion, it is enough to show that

(ϵ)1HD(Xϵ)2kVDνϵ(Xϵ)(ϵ)HD(Xϵ) for all Xπ(Zν) and 0<ϵχmin,\ell(\epsilon)^{-1}\operatorname{HD}(X^{\epsilon})\leq 2k\operatorname{VD}_{\nu}^{\epsilon}(X^{\epsilon})\leq\ell(\epsilon)\operatorname{HD}(X^{\epsilon})\quad\mbox{ for all }X\subseteq\pi(Z_{\nu}^{\star})\text{ and }0<\epsilon\ll\chi_{\min},

where we recall that XϵX^{\epsilon} is defined as in (4.9) and the constant (ϵ)\ell(\epsilon) is defined in (4.4). Observe that, as Lν=kχL_{\nu}=k\chi, we have (ϵ)=(χ+(2M+1)ϵ)/(χϵ)\ell(\epsilon)=\big{(}\chi+(2M+1)\epsilon\big{)}/(\chi-\epsilon).

We first prove the inequality HD(Xϵ)2k(ϵ)VDνϵ(Xϵ)\operatorname{HD}(X^{\epsilon})\leq 2k\ell(\epsilon)\operatorname{VD}_{\nu}^{\epsilon}(X^{\epsilon}). Fix α1>VDνϵ(Xϵ)\alpha_{1}>\operatorname{VD}_{\nu}^{\epsilon}(X^{\epsilon}). By Lemma 4.5, we have Λα1ϵ(Xϵ)=lim supκ0Λα1ϵ,κ(Xϵ)=0\Lambda^{\epsilon}_{\alpha_{1}}(X^{\epsilon})=\limsup_{\kappa\to 0}\Lambda^{\epsilon,\kappa}_{\alpha_{1}}(X^{\epsilon})=0. Therefore, for any η>0\eta>0 and up to taking 0<κ<r(ϵ)0<\kappa<r(\epsilon) sufficiently small, there exists an (ϵ,κ)(\epsilon,\kappa)-cover {Ui}i1\{U_{i}\}_{i\geq 1} of XϵX^{\epsilon} such that

(4.11) i1Vol(Ui)α1<η.\sum_{i\geq 1}\operatorname{Vol}(U_{i})^{\alpha_{1}}<\eta.

Setting Ni . . =N(Ui)N_{i}\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=N(U_{i}), by Corollary 2.9 (1) and (2) we have

diam(Ui)2k(ϵ)Vol(Ui)22k(ϵ)κ2k(ϵ)e2kNi(χϵ)(ϵ)κ2ke2kNi(χ+(2M+1)ϵ)22k(ϵ)κ2k((ϵ)1)22k(ϵ),\frac{\text{diam}(U_{i})^{2k\ell(\epsilon)}}{\operatorname{Vol}(U_{i})}\leq\frac{2^{2k\ell(\epsilon)}\kappa^{2k\ell(\epsilon)}e^{-2kN_{i}(\chi-\epsilon)\ell(\epsilon)}}{\kappa^{2k}e^{-2kN_{i}(\chi+(2M+1)\epsilon)}}\\ \leq 2^{2k\ell(\epsilon)}\kappa^{2k(\ell(\epsilon)-1)}\leq 2^{2k\ell(\epsilon)},

where in the last step we used the fact that κ<r(ϵ)<1\kappa<r(\epsilon)<1.

For each ii, define the ball Vi . . =B(z(Ui),diam(Ui))V_{i}\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=B(z(U_{i}),\text{diam}(U_{i})) of center z(Ui)z(U_{i}) and radius diam(Ui)\text{diam}(U_{i}). Then, UiViU_{i}\subseteq V_{i} and {Vi}i1\{V_{i}\}_{i\geq 1} is a cover of XϵX^{\epsilon} by balls. By the above estimates and (4.11), we have

i1diam(Vi)2k(ϵ)α1=i1(2diam(Ui))2k(ϵ)α124k(ϵ)α1i1Vol(Ui)α1<24k(ϵ)α1η.\sum_{i\geq 1}\operatorname{diam}(V_{i})^{2k\ell(\epsilon)\alpha_{1}}=\sum_{i\geq 1}(2\operatorname{diam}(U_{i}))^{2k\ell(\epsilon)\alpha_{1}}\leq 2^{4k\ell(\epsilon)\alpha_{1}}\sum_{i\geq 1}\operatorname{Vol}(U_{i})^{\alpha_{1}}<2^{4k\ell(\epsilon)\alpha_{1}}\eta.

Therefore, we have HD(Xϵ)2k(ϵ)α1\operatorname{HD}(X^{\epsilon})\leq 2k\ell(\epsilon)\alpha_{1} and the conclusion follows by taking α1VDνϵ(Xϵ)\alpha_{1}\searrow\operatorname{VD}_{\nu}^{\epsilon}(X^{\epsilon}).

We now prove the inequality 2k(ϵ)1VDνϵ(Xϵ)HD(Xϵ)2k\ell(\epsilon)^{-1}\operatorname{VD}_{\nu}^{\epsilon}(X^{\epsilon})\leq\operatorname{HD}(X^{\epsilon}). Fix α0\alpha_{0} such that Hα0(Xϵ)=0H_{\alpha_{0}}(X^{\epsilon})=0. Then, for any η>0\eta>0, there exists a cover {Bi=B(xi,ri)}i1\{B_{i}=B(x_{i},r_{i})\}_{i\geq 1} of XϵX^{\epsilon} consisting of open balls such that

(4.12) i1(2ri)α0<η.\sum_{i\geq 1}(2r_{i})^{\alpha_{0}}<\eta.

Fix any 0<κ<r(ϵ)0<\kappa<r(\epsilon). By definition of Hα0(Xϵ)H_{\alpha_{0}}(X^{\epsilon}) we can assume that

(4.13) supiri<κen(ϵ)(χ+(2M+1)ϵ).\sup_{i}r_{i}<\kappa\,e^{-n(\epsilon)(\chi+(2M+1)\epsilon)}.

For each i1i\geq 1, set

Ni . . =logκlogriχ+(2M+1)ϵN_{i}\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\big{\lfloor}\frac{\log\kappa-\log r_{i}}{\chi+(2M+1)\epsilon}\big{\rfloor}

and Ui . . =U(Ni,xi,κ,ϵ)U_{i}\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=U(N_{i},x_{i},\kappa,\epsilon). Observe that Nin(ϵ)N_{i}\geq n(\epsilon) for all ii by (4.13), hence every UiU_{i} is well-defined by Lemma 2.8. By Corollary 2.9 (1), for every ii we also have

B(xi,ri)B(xi,κeNi(χ+(2M+1)ϵ))UiB(xi,κeNi(χϵ)).B(x_{i},r_{i})\subseteq B(x_{i},\kappa\,e^{-N_{i}(\chi+(2M+1)\epsilon)})\subseteq U_{i}\subseteq B(x_{i},\kappa\,e^{-N_{i}(\chi-\epsilon)}).

In particular, the collection {Ui}i1\{U_{i}\}_{i\geq 1} is an (ϵ,κ)(\epsilon,\kappa)-cover of XϵX^{\epsilon} and, for all ii, we also have

(4.14) Vol(Ui)(ϵ)/(2k)κ(ϵ)eNi(χ+(2M+1)ϵ)κ(ϵ)e(1+logrilogκχ+(2M+1)ϵ)(χ+(2M+1)ϵ)eχ+(2M+1)ϵri,\operatorname{Vol}(U_{i})^{\ell(\epsilon)/(2k)}\leq\kappa^{\ell(\epsilon)}e^{-N_{i}(\chi+(2M+1)\epsilon)}\leq\kappa^{\ell(\epsilon)}e^{\left(1+\frac{\log r_{i}-\log\kappa}{\chi+(2M+1)\epsilon}\right)(\chi+(2M+1)\epsilon)}\leq e^{\chi+(2M+1)\epsilon}r_{i},

where we used the facts that xx1\lfloor x\rfloor\geq x-1 for every x>0x>0 and that κ<r(ϵ)<1\kappa<r(\epsilon)<1.

It follows from (4.12) and (4.14) that

i1Vol(Ui)α0(ϵ)/(2k)i1(12eχ+(2M+1)ϵ)α0(2ri)α0<(12eχ+(2M+1)ϵ)α0η.\sum_{i\geq 1}\operatorname{Vol}(U_{i})^{\alpha_{0}\ell(\epsilon)/(2k)}\leq\sum_{i\geq 1}\left(\frac{1}{2}e^{\chi+(2M+1)\epsilon}\right)^{\alpha_{0}}(2r_{i})^{\alpha_{0}}<\left(\frac{1}{2}e^{\chi+(2M+1)\epsilon}\right)^{\alpha_{0}}\eta.

Therefore, for every 0<κ<r(ϵ)0<\kappa<r(\epsilon), we have

(4.15) Λα0(ϵ)/(2k)ϵ,κ(Xϵ)<(12eχ+(2M+1)ϵ)α0η<(12eχ+(2M+1)ϵ)α0η.\Lambda^{\epsilon,\kappa}_{\alpha_{0}\ell(\epsilon)/(2k)}(X^{\epsilon})<\left(\frac{1}{2}e^{\chi+(2M+1)\epsilon}\right)^{\alpha_{0}}\eta<\left(\frac{1}{2}e^{\chi+(2M+1)\epsilon}\right)^{\alpha_{0}}\eta.

Taking the limsup over κ\kappa in the left hand side of (4.15), by (4.3) we obtain Λα0(ϵ)/(2k)ϵ(Xϵ)<\Lambda^{\epsilon}_{\alpha_{0}\ell(\epsilon)/(2k)}(X^{\epsilon})<\infty. Therefore, we have 2k(ϵ)1VDνϵ(Xϵ)HD(Xϵ)2k\ell(\epsilon)^{-1}\operatorname{VD}^{\epsilon}_{\nu}(X^{\epsilon})\leq\operatorname{HD}(X^{\epsilon}). This completes the proof of the first assertion.

We now prove the second assertion. We first show the inequality HD(ν)2kVD(ν)\operatorname{HD}(\nu)\leq 2k\operatorname{VD}(\nu). For every Xπ(Zν)X\subseteq\pi(Z_{\nu}^{\star}), we have 2kVDν(X)=HD(X)2k\operatorname{VD}_{\nu}(X)=\operatorname{HD}(X) by the first assertion. By Definition 4.17, we obtain 2kVD(ν)=inf{HD(X):Xπ(Zν),ν(X)=1}2k\operatorname{VD}(\nu)=\inf\{\operatorname{HD}(X):X\subseteq\pi(Z_{\nu}^{\star}),\nu(X)=1\}. By the definition (4.10) of HD(ν)\operatorname{HD}(\nu), this implies that HD(ν)2kVD(ν)\operatorname{HD}(\nu)\leq 2k\operatorname{VD}(\nu).

We now prove the inequality HD(ν)2kVD(ν)\operatorname{HD}(\nu)\geq 2k\operatorname{VD}(\nu). Take XkX\subseteq\mathbb{P}^{k} with ν(X)=1\nu(X)=1. Since ν(π(Zν))=1\nu(\pi(Z_{\nu}^{\star}))=1, we have ν(Xπ(Zν))=1\nu(X\cap\pi(Z_{\nu}^{\star}))=1. Then, by the first assertion and the monotonicity of the Hausdorff dimension, we have 2kVD(Xπ(Zν))=HD(Xπ(Zν))HD(X)2k\operatorname{VD}(X\cap\pi(Z_{\nu}^{\star}))=\operatorname{HD}(X\cap\pi(Z_{\nu}^{\star}))\leq\operatorname{HD}(X). By the definition (4.10) of HD(ν)\operatorname{HD}(\nu), we deduce that inf{2kVD(Xπ(Zν)):ν(X)=1}HD(ν)\inf\{2k\operatorname{VD}(X\cap\pi(Z_{\nu}^{\star})):\nu(X)=1\}\leq\operatorname{HD}(\nu). By Definition 4.17, this gives 2kVD(ν)HD(ν)2k\operatorname{VD}(\nu)\leq\operatorname{HD}(\nu) and completes the proof. ∎

4.2. An equivalent definition of VD(ν)\operatorname{VD}(\nu)

We present here an equivalent definition of the volume dimension for sets Xπ(Zν)X\subset\pi(Z_{\nu}^{\star}). This definition in particular allows us to prove that the lim supϵ0\limsup_{\epsilon\to 0} in Definition 4.14 is always a limit; see also Remark 5.1 for sets Xπ(Zν)X\subseteq\pi(Z_{\nu}^{\star}) with ν(X)>0\nu(X)>0. The advantage of this definition is that we will not have the small exponential terms eNMϵe^{-NM\epsilon} in the definition of the sets of the covers. In particular, we work with sets which are more similar to actual Bowen balls of fixed radius. On the other hand, the collection of neighbourhoods associated to any xx will in some sense depend on xx. This section is not necessary in order to obtain the main results of the paper.

For every 0<ϵχmin0<\epsilon\ll\chi_{\min}, 0<κ<r(ϵ)0<\kappa<r(\epsilon), ll\in\mathbb{N}, and Wπ(Zν(ϵ))W\subseteq\pi(Z_{\nu}^{\star}(\epsilon)), we consider the collection 𝒰~lκ(W,ϵ)\widetilde{\mathcal{U}}^{\kappa}_{l}(W,\epsilon) of open subsets of k\mathbb{P}^{k} given by

𝒰~lκ(W,ϵ) . . ={U~k:xW, such that U~=U~(nl,x,κ)}.\widetilde{\mathcal{U}}_{l}^{\kappa}(W,\epsilon)\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\big{\{}\widetilde{U}\subset\mathbb{P}^{k}\colon\exists x\in W,\text{ such that }\widetilde{U}=\widetilde{U}(n_{l},x,\kappa)\big{\}}.

Here, {nl}l0\{n_{l}\}_{l\geq 0} is the sequence associated to xπ(Zν(ϵ))x\in\pi(Z_{\nu}(\epsilon)) by Lemma 2.5, and, letting x^\hat{x} be any element of Zν(ϵ)Z_{\nu}(\epsilon) with x0=xx_{0}=x, we set

U~(nl,x,κ) . . =fTnl(x^)nl(B(fnl(x0),κ)),\widetilde{U}(n_{l},x,\kappa)\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=f_{T^{n_{l}}(\hat{x})}^{-n_{l}}\big{(}B(f^{n_{l}}(x_{0}),\kappa)\big{)},

where the right hand side of the above expression is well-defined by Corollary 2.4. For every ϵ\epsilon and κ\kappa as above and Nn(ϵ)\mathbb{N}\ni N^{\star}\geq n(\epsilon), we denote by 𝒰~(ϵ,κ,N)\widetilde{\mathcal{U}}(\epsilon,\kappa,N^{\star}) the collection of open sets

𝒰~(ϵ,κ,N) . . =nlN𝒰~lκ(π(Zν(ϵ)),ϵ).\widetilde{\mathcal{U}}(\epsilon,\kappa,N^{\star})\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\bigcup_{n_{l}\geq N^{\star}}\widetilde{\mathcal{U}}^{\kappa}_{l}(\pi(Z^{\star}_{\nu}(\epsilon)),\epsilon).

For every α0\alpha\geq 0 and Yπ(Zν(ϵ))Y\subseteq\pi(Z^{\star}_{\nu}(\epsilon)), we define Λ~αϵ(Y)[0,+]\widetilde{\Lambda}^{\epsilon}_{\alpha}(Y)\in[0,+\infty] as

(4.16) Λ~αϵ(Y) . . =lim supκ0Λ~αϵ,κ(Y), where Λ~αϵ,κ(Y) . . =limNinf{Ui}i1Vol(U~i)α\widetilde{\Lambda}^{\epsilon}_{\alpha}(Y)\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\limsup_{\kappa\to 0}\widetilde{\Lambda}^{\epsilon,\kappa}_{\alpha}(Y),\quad\text{ where }\quad\widetilde{\Lambda}^{\epsilon,\kappa}_{\alpha}(Y)\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\lim_{N^{\star}\to\infty}\inf_{\{U_{i}\}}\sum_{i\geq 1}\operatorname{Vol}(\widetilde{U}_{i})^{\alpha}

and the infimum is taken over all covers {U~i}i1\{\widetilde{U}_{i}\}_{i\geq 1} of YY with U~i𝒰~(ϵ,κ,N)\widetilde{U}_{i}\in\widetilde{\mathcal{U}}(\epsilon,\kappa,N^{\star}) for all i1i\geq 1. As in Lemma 4.5, one can show that, for every 0<ϵχmin0<\epsilon\ll\chi_{\min} and Yπ(Zν(ϵ))Y\subseteq\pi(Z^{\star}_{\nu}(\epsilon)), the function αΛ~αϵ(Y)\alpha\mapsto\widetilde{\Lambda}^{\epsilon}_{\alpha}(Y) is non-increasing and that, if Λ~α0ϵ(Y)<\widetilde{\Lambda}^{\epsilon}_{\alpha_{0}}(Y)<\infty for some α00\alpha_{0}\geq 0, then Λ~αϵ(Y)=0\widetilde{\Lambda}^{\epsilon}_{\alpha}(Y)=0 for all α>α0\alpha>\alpha_{0}. As a consequence, the following definition is well-posed.

Definition 4.21.

For every 0<ϵχmin0<\epsilon\ll\chi_{\min} and YZν(ϵ)Y\subseteq Z^{\star}_{\nu}(\epsilon), we set

VD~νϵ(Y) . . =sup{α:Λ~αϵ(Y)=}=inf{α:Λ~αϵ(Y)=0}.\widetilde{\operatorname{VD}}^{\epsilon}_{\nu}(Y)\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\sup\{\alpha:\widetilde{\Lambda}^{\epsilon}_{\alpha}(Y)=\infty\}=\inf\{\alpha:\widetilde{\Lambda}^{\epsilon}_{\alpha}(Y)=0\}.
Lemma 4.22.

For every 0<ϵ1<ϵ2χmin0<\epsilon_{1}<\epsilon_{2}\ll\chi_{\min} and Yπ(Zν(ϵ))Y\subseteq\pi(Z^{\star}_{\nu}(\epsilon)), we have VD~νϵ1(Y)=VD~νϵ2(Y)\widetilde{\operatorname{VD}}^{\epsilon_{1}}_{\nu}(Y)=\widetilde{\operatorname{VD}}^{\epsilon_{2}}_{\nu}(Y).

Proof.

The statement is clear since the sets U~(nl,x,κ)\widetilde{U}(n_{l},x,\kappa) do not depend on ϵ\epsilon. ∎

Lemma 4.23.

For every 0<ϵχmin0<\epsilon\ll\chi_{\min} and Yπ(Zν(ϵ))Y\subseteq\pi(Z^{\star}_{\nu}(\epsilon)) we have

(ϵ)1VD~νϵ(Y)VDνϵ(Y)β(ϵ)VD~νϵ(Y),\ell(\epsilon)^{-1}\widetilde{\operatorname{VD}}^{\epsilon}_{\nu}(Y)\leq\operatorname{VD}^{\epsilon}_{\nu}(Y)\leq\beta(\epsilon)\widetilde{\operatorname{VD}}^{\epsilon}_{\nu}(Y),

where (ϵ)>1\ell(\epsilon)>1 is as in (4.4) and β(ϵ) . . =Lν+kϵLνkϵ(minjχjϵχj+(2M+1)ϵ1n(ϵ))1\beta(\epsilon)\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\frac{L_{\nu}+k\epsilon}{L_{\nu}-k\epsilon}\cdot\big{(}\min_{j}\frac{\chi_{j}-\epsilon}{\chi_{j}+(2M+1)\epsilon}-\frac{1}{n(\epsilon)}\big{)}^{-1}.

Observe that β(ϵ)\beta(\epsilon) as in the statement above satisfies β(ϵ)>1\beta(\epsilon)>1 for all 0<ϵχmin0<\epsilon\ll\chi_{\min} and β(ϵ)=1+O(ϵ)\beta(\epsilon)=1+O(\epsilon) as ϵ0\epsilon\to 0.

Proof.

We first prove the first inequality. Suppose α0\alpha\geq 0 is such that Λαϵ(Y)=0\Lambda^{\epsilon}_{\alpha}(Y)=0. Then for any η>0\eta>0 and 0<κ<r(ϵ)0<\kappa<r(\epsilon), there exists a (κ,ϵ)(\kappa,\epsilon)-cover {Ui}i1\{U_{i}\}_{i\geq 1} of YY of the form Ui=U(Ni,xi,κ,ϵ)U_{i}=U(N_{i},x_{i},\kappa,\epsilon), with NiNn(ϵ)N_{i}\geq N^{\star}\geq n(\epsilon) for all ii, such that i1Vol(Ui)α<η\sum_{i\geq 1}\operatorname{Vol}(U_{i})^{\alpha}<\eta.

For each i1i\geq 1, let l(i)l(i) be such that nl(i)Ni<nl(i)+1n_{l(i)}\leq N_{i}<n_{l(i)+1}. Such l(i)l(i) exists since, by Lemma 2.5, we have n0(x)n(ϵ)n_{0}(x)\leq n(\epsilon) for all xπ(Zν(ϵ))x\in\pi(Z_{\nu}(\epsilon)). For every ii, we then have UiU~i . . =U~(nl(i)+1,xi,κ)U_{i}\subset\widetilde{U}_{i}\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\widetilde{U}(n_{l(i)+1},x_{i},\kappa), see also (2.8) in the proof of Lemma 2.8. In particular, we have {U~i}𝒰~(ϵ,κ,N)\{\widetilde{U}_{i}\}\subset\widetilde{\mathcal{U}}(\epsilon,\kappa,N^{\star}) and the sets {U~i}\{\widetilde{U}_{i}\} form a cover of YY. It follows from the definition of (ϵ)\ell(\epsilon) that for all i1i\geq 1, we have

Vol(U~(nl(i)+1,xi,κ))(ϵ)Vol(U(Ni,xi,κ,ϵ))κ2k(ϵ)e2nl(i)+1(Lνkϵ)(ϵ)κ2ke2Ni(Lν+k(2M+1)ϵ)=κ2k((ϵ)1)e2(Lν+κϵ(2M+1))(Ninl(i)+1)1,\frac{\operatorname{Vol}(\widetilde{U}(n_{l(i)+1},x_{i},\kappa))^{\ell(\epsilon)}}{\operatorname{Vol}(U(N_{i},x_{i},\kappa,\epsilon))}\leq\frac{\kappa^{2k\ell(\epsilon)}e^{-2n_{l(i)+1}(L_{\nu}-k\epsilon)\ell(\epsilon)}}{\kappa^{2k}e^{-2N_{i}(L_{\nu}+k(2M+1)\epsilon)}}=\kappa^{2k(\ell(\epsilon)-1)}e^{2(L_{\nu}+\kappa\epsilon(2M+1))(N_{i}-n_{l(i)+1})}\leq 1,

where in the last inequality we used the facts the Ni<nl(i)+1N_{i}<n_{l(i)+1} and κ<r(ϵ)<1\kappa<r(\epsilon)<1. In particular, we have

i1Vol(U~(nl(i)+1,xi,κ))α(ϵ)i1Vol(Ui)α<η,\sum_{i\geq 1}\operatorname{Vol}\big{(}\widetilde{U}(n_{l(i)+1},x_{i},\kappa)\big{)}^{\alpha\ell(\epsilon)}\leq\sum_{i\geq 1}\operatorname{Vol}(U_{i})^{\alpha}<\eta,

which gives the inequality Λ~α(ϵ)ϵ,κ(Y)Λαϵ,κ(Y)\widetilde{\Lambda}^{\epsilon,\kappa}_{\alpha\ell(\epsilon)}(Y)\leq\Lambda^{\epsilon,\kappa}_{\alpha}(Y) for any 0<κ<r(ϵ)0<\kappa<r(\epsilon). Taking the limsup over κ\kappa as in the definition of Λ~α(ϵ)ϵ(Y)\widetilde{\Lambda}^{\epsilon}_{\alpha\ell(\epsilon)}(Y), we obtain Λ~α(ϵ)ϵ(Y)<\widetilde{\Lambda}^{\epsilon}_{\alpha\ell(\epsilon)}(Y)<\infty. By the choice of α\alpha, we deduce the desired inequality VD~νϵ(Y)(ϵ)VDνϵ(Y)\widetilde{\operatorname{VD}}^{\epsilon}_{\nu}(Y)\leq\ell(\epsilon)\operatorname{VD}^{\epsilon}_{\nu}(Y).

We now prove the second inequality. Suppose Λ~αϵ(Y)=0\widetilde{\Lambda}^{\epsilon}_{\alpha}(Y)=0. Then for any η>0\eta>0 and 0<κ<r(ϵ)0<\kappa<r(\epsilon), there exists a cover {U~i}i1𝒰~(ϵ,κ,N)\{\widetilde{U}_{i}\}_{i\geq 1}\subset\widetilde{\mathcal{U}}(\epsilon,\kappa,N^{\star}) of YY, which each U~i\widetilde{U}_{i} of the form U~i=U~(nl(i),xi,κ)\widetilde{U}_{i}=\widetilde{U}(n_{l(i)},x_{i},\kappa), such that i1Vol(U~i)α<η\sum_{i\geq 1}\operatorname{Vol}(\widetilde{U}_{i})^{\alpha}<\eta. For each i1i\geq 1, set

Ni . . =nl(i)minjχjϵχj+(2M+1)ϵ.N_{i}\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\big{\lfloor}n_{l(i)}\cdot\min_{j}\frac{\chi_{j}-\epsilon}{\chi_{j}+(2M+1)\epsilon}\big{\rfloor}.

From the definition of NiN_{i} and Lemma 2.8, for all i1i\geq 1, we have

U~(nl(i),xi,κ)U(Ni,xi,κ,ϵ).\widetilde{U}(n_{l(i)},x_{i},\kappa)\subset U(N_{i},x_{i},\kappa,\epsilon).

It follows that, for every i1i\geq 1, we have

Vol(U(Ni,xi,κ,ϵ))β(ϵ)Vol(U~(nl(i),xi,κ))κ2kβ(ϵ)e2Ni(Lνkϵ)β(ϵ)κ2ke2nl(i)(Lν+kϵ)1,\frac{\operatorname{Vol}(U(N_{i},x_{i},\kappa,\epsilon))^{\beta(\epsilon)}}{\operatorname{Vol}(\widetilde{U}(n_{l(i)},x_{i},\kappa))}\leq\frac{\kappa^{2k\beta(\epsilon)}e^{-2N_{i}(L_{\nu}-k\epsilon)\beta(\epsilon)}}{\kappa^{2k}e^{-2n_{l(i)}(L_{\nu}+k\epsilon)}}\leq 1,

where in the last step we used the facts that κ<r(ϵ)<1\kappa<r(\epsilon)<1 and that, since nl(i)n(ϵ)n_{l(i)}\geq n(\epsilon) for all i1i\geq 1 and rr1\lfloor r\rfloor\geq r-1 for all r>0r>0, we have

NiLνkϵLν+kϵβ(ϵ)nl(i)minjχjϵχj+(2M+1)ϵ1minjχjϵχj+(2M+1)ϵ1n(ϵ)nl(i).N_{i}\frac{L_{\nu}-k\epsilon}{L_{\nu}+k\epsilon}\beta(\epsilon)\geq\frac{n_{l(i)}\cdot\min_{j}\frac{\chi_{j}-\epsilon}{\chi_{j}+(2M+1)\epsilon}-1}{\min_{j}\frac{\chi_{j}-\epsilon}{\chi_{j}+(2M+1)\epsilon}-\frac{1}{n(\epsilon)}}\geq n_{l(i)}.

Therefore, we have

i1Vol(U(Ni,xi,κ,ϵ))αβ(ϵ)η,\sum_{i\geq 1}\operatorname{Vol}(U(N_{i},x_{i},\kappa,\epsilon))^{\alpha\beta(\epsilon)}\leq\eta,

which gives the inequality Λαβ(ϵ)ϵ,κ(Y)Λ~αϵ,κ(Y)\Lambda^{\epsilon,\kappa}_{\alpha\beta(\epsilon)}(Y)\leq\widetilde{\Lambda}^{\epsilon,\kappa}_{\alpha}(Y) for any 0<κ<r(ϵ)0<\kappa<r(\epsilon). Taking the limsup over κ\kappa as in the definition of Λαβ(ϵ)ϵ(Y)\Lambda^{\epsilon}_{\alpha\beta(\epsilon)}(Y), we obtain Λαβ(ϵ)ϵ(Y)=0\Lambda^{\epsilon}_{\alpha\beta(\epsilon)}(Y)=0. By the choice of α\alpha, we have VDνϵ(Y)β(ϵ)VD~νϵ(Y)\operatorname{VD}^{\epsilon}_{\nu}(Y)\leq\beta(\epsilon)\widetilde{\operatorname{VD}}^{\epsilon}_{\nu}(Y). The proof is complete. ∎

Thanks to Lemma 4.23, one can see that the lim supϵ0\limsup_{\epsilon\to 0} in the Definition 4.6 is actually a limit. Recall that, for every Xπ(Zν)X\subseteq\pi(Z_{\nu}^{\star}), we denote Xϵ . . =Xπ(Zν(ϵ))X^{\epsilon}\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=X\cap\pi(Z^{\star}_{\nu}(\epsilon)).

Corollary 4.24.

For every Xπ(Zν)X\subseteq\pi(Z^{\star}_{\nu}), we have

VDν(X)=limϵ0VDνϵ(Xϵ).\operatorname{VD}_{\nu}(X)=\lim_{\epsilon\to 0}\operatorname{VD}^{\epsilon}_{\nu}(X^{\epsilon}).
Proof.

For every Xπ(Zν)X\subseteq\pi(Z_{\nu}^{\star}), set VD~ν(X) . . =limϵ0VD~νϵ(Xϵ)\widetilde{\operatorname{VD}}_{\nu}(X)\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\lim_{\epsilon\to 0}\widetilde{\operatorname{VD}}_{\nu}^{\epsilon}(X^{\epsilon}). The limit is well-defined and equal to the supremum over 0<ϵχmin0<\epsilon\ll\chi_{\min} since, for every 0<ϵ1<ϵ2χmin0<\epsilon_{1}<\epsilon_{2}\ll\chi_{\min}, we have VD~νϵ1(Xϵ2)=VD~νϵ2(Xϵ2)\widetilde{\operatorname{VD}}^{\epsilon_{1}}_{\nu}(X^{\epsilon_{2}})=\widetilde{\operatorname{VD}}^{\epsilon_{2}}_{\nu}(X^{\epsilon_{2}}) by Lemma 4.22 and VD~νϵ1(Xϵ1)VD~νϵ1(Xϵ2)\widetilde{\operatorname{VD}}^{\epsilon_{1}}_{\nu}(X^{\epsilon_{1}})\geq\widetilde{\operatorname{VD}}^{\epsilon_{1}}_{\nu}(X^{\epsilon_{2}}) since Xϵ1Xϵ2X^{\epsilon_{1}}\supseteq X^{\epsilon_{2}}. It follows from Lemma 4.23 that limϵ0VDνϵ(Xϵ)\lim_{\epsilon\to 0}\operatorname{VD}^{\epsilon}_{\nu}(X^{\epsilon}) is well-defined and equal to VD~ν(X)\widetilde{\operatorname{VD}}_{\nu}(X). The assertion follows. ∎

Remark 4.25.

A further possible (equivalent) way to define the volume dimension is the following. Take ν+(f)\nu\in\mathcal{M}^{+}(f). Fix 0<ϵχmin0<\epsilon\ll\chi_{\min} and take Yπ(Zν(ϵ))Y\subseteq\pi(Z^{\star}_{\nu}(\epsilon)). For every α0\alpha\geq 0, define

Λˇαϵ(Y) . . =lim supκ0Λˇαϵ,κ(Y), where Λˇαϵ,κ(Y) . . =limNinf{Ui}i1e2N(Ui)Lν(f)ακ2kα,\check{\Lambda}_{\alpha}^{\epsilon}(Y)\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\limsup_{\kappa\to 0}\check{\Lambda}^{\epsilon,\kappa}_{\alpha}(Y),\quad\text{ where }\quad\check{\Lambda}^{\epsilon,\kappa}_{\alpha}(Y)\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\lim_{N^{\star}\to\infty}\inf_{\{U_{i}\}}\sum_{i\geq 1}e^{-2N(U_{i})L_{\nu}(f)\alpha}\kappa^{2k\alpha},

and the infimum is taken over all countable covers {Ui}i1𝒰(ϵ,κ,N)\{U_{i}\}_{i\geq 1}\subset\mathcal{U}(\epsilon,\kappa,N^{\star}) of YY. Recall that Lν(f)L_{\nu}(f) denotes the sum of the Lyapunov exponents of ν\nu. As in Lemma 4.5, one can prove that, for every 0<ϵχmin0<\epsilon\ll\chi_{\min} and Yπ(Zν(ϵ))Y\subseteq\pi(Z^{\star}_{\nu}(\epsilon)), the function αΛˇαϵ(Y)\alpha\mapsto\check{\Lambda}_{\alpha}^{\epsilon}(Y) is non-increasing in α\alpha, and that if Λˇα0ϵ(Y)\check{\Lambda}_{\alpha_{0}}^{\epsilon}(Y) is finite, then Λˇαϵ(Y)=0\check{\Lambda}_{\alpha}^{\epsilon}(Y)=0 for all α>α0\alpha>\alpha_{0}. Hence, the quantity

VDˇνϵ(Y) . . =inf{α:Λˇαϵ(Y)=0}=sup{α:Λˇαϵ(Y)=}\check{\operatorname{VD}}_{\nu}^{\epsilon}(Y)\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\inf\big{\{}\alpha\colon\check{\Lambda}_{\alpha}^{\epsilon}(Y)=0\big{\}}=\sup\big{\{}\alpha\colon\check{\Lambda}_{\alpha}^{\epsilon}(Y)=\infty\big{\}}

is well-defined for all 0<ϵχmin0<\epsilon\ll\chi_{\min} and Yπ(Zν(ϵ))Y\subseteq\pi(Z^{\star}_{\nu}(\epsilon)). For every 0<ϵχmin0<\epsilon\ll\chi_{\min}, define the constants

(ϵ) . . =Lν(f)kϵLν(f) and +(ϵ) . . =Lν(f)+(2M+1)kϵLν(f).\ell_{-}(\epsilon)\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\frac{L_{\nu}(f)-k\epsilon}{L_{\nu}(f)}\quad\mbox{ and }\quad\ell_{+}(\epsilon)\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\frac{L_{\nu}(f)+(2M+1)k\epsilon}{L_{\nu}(f)}.

Observe that, for all 0<ϵχmin0<\epsilon\ll\chi_{\min}, we have (ϵ)<1<+(ϵ)\ell_{-}(\epsilon)<1<\ell_{+}(\epsilon) and (ϵ),+(ϵ)1\ell_{-}(\epsilon),\ell_{+}(\epsilon)\to 1 as ϵ0\epsilon\to 0. One can show in this case that, for every 0<ϵχmin0<\epsilon\ll\chi_{\min} and Yπ(Zν(ϵ))Y\subseteq\pi(Z^{\star}_{\nu}(\epsilon)), we have

(4.17) (ϵ)VDνϵ(Y)VDˇνϵ(Y)+(ϵ)VDνϵ(Y).\ell_{-}(\epsilon)\operatorname{VD}_{\nu}^{\epsilon}(Y)\leq\check{\operatorname{VD}}^{\epsilon}_{\nu}(Y)\leq\ell_{+}(\epsilon)\operatorname{VD}_{\nu}^{\epsilon}(Y).

Take now Xπ(Zν)X\subseteq\pi(Z_{\nu}^{\star}). As before, setting Xϵ . . =Xπ(Zν(ϵ))X^{\epsilon}\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=X\cap\pi(Z^{\star}_{\nu}(\epsilon)) for every 0<ϵχmin0<\epsilon\ll\chi_{\min}, we can define

VDˇν(X)\displaystyle\check{\operatorname{VD}}_{\nu}(X) . . =lim supϵ0VDνϵ(Xϵ) and\displaystyle\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\limsup_{\epsilon\to 0}\operatorname{VD}_{\nu}^{\epsilon}(X^{\epsilon})\quad\mbox{ and }
VDˇ(ν)\displaystyle\check{\operatorname{VD}}(\nu) . . =inf{VDˇν(X):Xπ(Zν) and ν(X)=1}.\displaystyle\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\inf\big{\{}\check{\operatorname{VD}}_{\nu}(X)\colon X\subseteq\pi(Z_{\nu}^{\star})\mbox{ and }\nu(X)=1\big{\}}.

It follows from (4.17), applied with Y=XϵY=X^{\epsilon}, that VDˇν(X)=VDν(X)\check{\operatorname{VD}}_{\nu}(X)=\operatorname{VD}_{\nu}(X) for all Xπ(Zν)X\subseteq\pi(Z_{\nu}^{\star}), and that VDˇ(ν)=VD(ν)\check{\operatorname{VD}}(\nu)=\operatorname{VD}(\nu).

4.3. From local volume dimensions to volume dimensions

Fix a measure ν+(f)\nu\in\mathcal{M}^{+}(f) and 0<ϵχmin0<\epsilon\ll\chi_{\min}. For xπ(Zν(ϵ))x\in\pi(Z^{\star}_{\nu}(\epsilon)), 0<κ<r(ϵ)0<\kappa<r(\epsilon), and Nn(ϵ)N\geq n(\epsilon) recall that δx(ϵ,κ,N)\delta_{x}(\epsilon,\kappa,N) is defined in (3.1) and well-defined by Lemma 2.8. The integer m1(ϵ,x)m_{1}(\epsilon,x) in Theorem 3.2 is uniformly bounded from above for all xπ(Zν(ϵ))x\in\pi(Z^{\star}_{\nu}(\epsilon)) by the definition (4.1) of Zν(ϵ)Z^{\star}_{\nu}(\epsilon). This fact is crucial in the proof of the next statement. Recall that a measure is non-atomic if it does not assign mass to points.

Proposition 4.26.

Let ff be a holomorphic endomorphism of k\mathbb{P}^{k} of algebraic degree d2d\geq 2 and take ν+(f)\nu\in\mathcal{M}^{+}(f). Assume that ν\nu is non-atomic. Fix α1,α20\alpha_{1},\alpha_{2}\geq 0 and 0<ϵχmin0<\epsilon\ll\chi_{\min}. Let Yπ(Zν(ϵ))Y\subseteq\pi(Z^{\star}_{\nu}(\epsilon)) be such that ν(Y)>0\nu(Y)>0. Suppose that for every 0<κ<r(ϵ)0<\kappa<r(\epsilon) there exists m=m(ϵ,κ)n(ϵ)m=m(\epsilon,\kappa)\geq n(\epsilon) such that

(4.18) α1δx(ϵ,κ,N)α2 for all xY and Nm(ϵ,κ).\alpha_{1}\leq\delta_{x}(\epsilon,\kappa,N)\leq\alpha_{2}\quad\mbox{ for all }x\in Y\mbox{ and }N\geq m(\epsilon,\kappa).

Then, we have

α1VDνϵ(Y)α2.\alpha_{1}\leq\operatorname{VD}^{\epsilon}_{\nu}(Y)\leq\alpha_{2}.
Proof.

The proof of the proposition essentially follows the arguments in [You82, Proposition 2.1]. Recall that, for every 0<κ<r(ϵ)0<\kappa<r(\epsilon) and N0N^{\star}\geq 0, the collection 𝒰(ϵ,κ,N)\mathcal{U}(\epsilon,\kappa,N^{\star}) is defined in (4.2) and is a cover of YY; see Lemma 4.3. Define the quantity

α(Y,ν) . . =inf{α:limNinf𝒰U𝒰ν(U)α=0},\alpha(Y,\nu)\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\inf\Big{\{}\alpha:\lim_{N^{\star}\to\infty}\inf_{\mathcal{U}^{\prime}}\sum_{U\in\mathcal{U}^{\prime}}\nu(U)^{\alpha}=0\Big{\}},

where the infimum is taken over all the sub-covers 𝒰𝒰(ϵ,κ,N)\mathcal{U}^{\prime}\subset\mathcal{U}(\epsilon,\kappa,N^{\star}) of YY. Since ν\nu is non-atomic and ν(Y)>0\nu(Y)>0, we have α(Y,ν)=1\alpha(Y,\nu)=1; see [Bil65, Section 14]. We use here the fact that, for every fixed 0<κ<r(ϵ)0<\kappa<r(\epsilon), the sets U(N,x,κ,ϵ)U(N,x,\kappa,\epsilon) shrink to {x}\{x\} as NN\to\infty; see Lemma 2.8.

Fix α>1\alpha>1, 0<κ<r(ϵ)0<\kappa<r(\epsilon), Nm(ϵ,κ)N^{\star}\geq m(\epsilon,\kappa), and η>0\eta>0. Since α(Y,ν)=1\alpha(Y,\nu)=1, there exists a cover 𝒰0𝒰(ϵ,κ,N)\mathcal{U}_{0}\subset\mathcal{U}(\epsilon,\kappa,N^{\star}) of YY such that

U𝒰0ν(U)α<η.\sum_{U\in\mathcal{U}_{0}}\nu(U)^{\alpha}<\eta.

By the assumption (4.18) and the choice Nm(ϵ,κ)N^{\star}\geq m(\epsilon,\kappa), for every U𝒰(ϵ,κ,N)U\in\mathcal{U}(\epsilon,\kappa,N^{\star}) we have Vol(U)α2ν(U)\operatorname{Vol}(U)^{\alpha_{2}}\leq\nu(U). Hence, we have

U𝒰0Vol(U)α2αU𝒰0ν(U)α<η.\sum_{U\in\mathcal{U}_{0}}\operatorname{Vol}(U)^{\alpha_{2}\alpha}\leq\sum_{U\in\mathcal{U}_{0}}\nu(U)^{\alpha}<\eta.

This shows the inequality Λα2αϵ,κ(Y)<η\Lambda^{\epsilon,\kappa}_{\alpha_{2}\alpha}(Y)<\eta for any 0<κ<r(ϵ)0<\kappa<r(\epsilon). Therefore, we have lim supκ0Λα2αϵ,κ(Y)=Λα2αϵ(Y)<η\limsup_{\kappa\to 0}\Lambda^{\epsilon,\kappa}_{\alpha_{2}\alpha}(Y)=\Lambda^{\epsilon}_{\alpha_{2}\alpha}(Y)<\eta. Taking α1\alpha\searrow 1, we obtain the inequality VDνϵ(Y)α2\operatorname{VD}_{\nu}^{\epsilon}(Y)\leq\alpha_{2}.

For the other inequality, again by the assumption (4.18), for all 0<κ<r(ϵ)0<\kappa<r(\epsilon) and Nm(ϵ,κ)N^{\star}\geq m(\epsilon,\kappa) we have Vol(U)α1ν(U)\operatorname{Vol}(U)^{\alpha_{1}}\geq\nu(U) for every U𝒰(ϵ,κ,N)U\in\mathcal{U}(\epsilon,\kappa,N^{\star}). Hence, for any cover 𝒰0𝒰(ϵ,κ,N)\mathcal{U}_{0}\subset\mathcal{U}(\epsilon,\kappa,N^{\star}), we have

U𝒰0Vol(U)α1U𝒰0ν(U)ν(Y).\sum_{U\in\mathcal{U}_{0}}\operatorname{Vol}(U)^{\alpha_{1}}\geq\sum_{U\in\mathcal{U}_{0}}\nu(U)\geq\nu(Y).

Therefore, we have Λα1ϵ,κ(Y)ν(Y)>0\Lambda^{\epsilon,\kappa}_{\alpha_{1}}(Y)\geq\nu(Y)>0 for any 0<κ<r(ϵ)0<\kappa<r(\epsilon), which gives

Λα1ϵ(Y)=lim supκ0Λα1ϵ,κ(Y)ν(Y)>0.\Lambda^{\epsilon}_{\alpha_{1}}(Y)=\limsup_{\kappa\to 0}\Lambda^{\epsilon,\kappa}_{\alpha_{1}}(Y)\geq\nu(Y)>0.

Hence, we have VDνϵ(Y)α1\operatorname{VD}_{\nu}^{\epsilon}(Y)\geq\alpha_{1}. The proof is complete. ∎

Remark 4.27.

Let ν+(f)\nu\in\mathcal{M}^{+}(f) be non-atomic. Take Xπ(Zν)X\subseteq\pi(Z_{\nu}^{\star}) with ν(X)>0\nu(X)>0. Setting Xϵ . . =Xπ(Zν(ϵ))X^{\epsilon}\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=X\cap\pi(Z^{\star}_{\nu}(\epsilon)), assume that for every 0<ϵχmin0<\epsilon\ll\chi_{\min} and 0<κ<r(ϵ)0<\kappa<r(\epsilon) there exists m=m(ϵ,κ)m=m(\epsilon,\kappa) and α10,α20\alpha_{1}^{0},\alpha_{2}^{0}\in\mathbb{R} such that

(4.19) α1(ϵ)δx(ϵ,κ,N)α2(ϵ) for all xXϵ and Nm(ϵ,κ)\alpha_{1}(\epsilon)\leq\delta_{x}(\epsilon,\kappa,N)\leq\alpha_{2}(\epsilon)\quad\mbox{ for all }x\in X^{\epsilon}\mbox{ and }N\geq m(\epsilon,\kappa)

for some functions α1(ϵ)=α10+O(ϵ)\alpha_{1}(\epsilon)=\alpha_{1}^{0}+O(\epsilon) and α2(ϵ)=α20+O(ϵ)\alpha_{2}(\epsilon)=\alpha_{2}^{0}+O(\epsilon). Applying Proposition 4.26 to XϵX^{\epsilon} instead of YY we see that, for every 0<ϵχmin0<\epsilon\ll\chi_{\min}, we have α1(ϵ)VDνϵ(Xϵ)α2(ϵ)\alpha_{1}(\epsilon)\leq\operatorname{VD}_{\nu}^{\epsilon}(X^{\epsilon})\leq\alpha_{2}(\epsilon), which gives α10VDν(X)α20\alpha_{1}^{0}\leq\operatorname{VD}_{\nu}(X)\leq\alpha_{2}^{0}.

5. Proofs of Theorems 1.1, 1.2, and 1.3

In this section, f:kkf\colon\mathbb{P}^{k}\to\mathbb{P}^{k} is a holomorphic endomorphism of algebraic degree d2d\geq 2.

5.1. Proof of Theorem 1.1

For ν+(f)\nu\in\mathcal{M}^{+}(f), recall that ZνZ_{\nu} is defined in Definition 2.3 and for every 0<ϵχmin0<\epsilon\ll\chi_{\min} (where χmin>0\chi_{\min}>0 is the smallest Lyapunov exponent of ν\nu), the set Zν(ϵ)Z^{\star}_{\nu}(\epsilon) is defined in (4.1).

Assume first that ν\nu is atomic. Since ν\nu is ergodic, it gives mass only to a finite number of points,. hence it satisfies hν(f)=0h_{\nu}(f)=0. It also follows from the Definition 4.17 of VD(ν)\operatorname{VD}(\nu) that VD(ν)=0\operatorname{VD}(\nu)=0, since the support SνS_{\nu} of ν\nu satisfies ν(Sν)=1\nu(S_{\nu})=1 and VDν(Sν)=0\operatorname{VD}_{\nu}(S_{\nu})=0, being finite. Therefore, the conclusion follows in this case.

We can then assume that ν\nu is non-atomic. By Theorem 3.2 and Proposition 4.26, for every 0<ϵχmin0<\epsilon\ll\chi_{\min} we have

VDνϵ(π(Zν(ϵ)))hν(f)2Lν(f)+cϵ,\operatorname{VD}^{\epsilon}_{\nu}(\pi(Z^{\star}_{\nu}(\epsilon)))\leq\frac{h_{\nu}(f)}{2L_{\nu}(f)}+c\epsilon,

where the constant cc is independent of ϵ\epsilon. By Definition 4.14, taking ϵ0\epsilon\searrow 0, we obtain the inequality VDν(π(Zν))(2Lν(f))1hν(f)\operatorname{VD}_{\nu}(\pi(Z_{\nu}))\leq(2L_{\nu}(f))^{-1}h_{\nu}(f). As ν(π(Zν))=1\nu(\pi(Z_{\nu}))=1, by Definition 4.17, we deduce the inequality VD(ν)(2Lν(f))1hν(f)\operatorname{VD}(\nu)\leq(2L_{\nu}(f))^{-1}h_{\nu}(f).

In order to prove the reversed inequality, let Y0π(Zν)Y_{0}\subseteq\pi(Z^{\star}_{\nu}) be such that ν(Y0)=1\nu(Y_{0})=1. For any 0<ϵχmin0<\epsilon\ll\chi_{\min}, applying Proposition 4.26 to Y0π(Zν(ϵ))Y_{0}\cap\pi(Z^{\star}_{\nu}(\epsilon)), we deduce from Theorem 3.2 that

VDνϵ(Y0π(Zν(ϵ)))hν(f)2Lν(f)cϵ,\operatorname{VD}^{\epsilon}_{\nu}(Y_{0}\cap\pi(Z^{\star}_{\nu}(\epsilon)))\geq\frac{h_{\nu}(f)}{2L_{\nu}(f)}-c\epsilon,

where again the constant cc is independent of ϵ\epsilon; see also Remark 4.27. By Definition 4.14, we have the inequality VDν(Y0)(2Lν(f))1hν(f)\operatorname{VD}_{\nu}(Y_{0})\geq(2L_{\nu}(f))^{-1}h_{\nu}(f). As Y0Y_{0} is arbitrary, it follows from Definition 4.17 that VD(ν)(2Lν(f))1hν(f).\operatorname{VD}(\nu)\geq(2L_{\nu}(f))^{-1}h_{\nu}(f). The proof of Theorem 1.1 is complete.

Remark 5.1.

Let ν+(f)\nu\in\mathcal{M}^{+}(f) be non-atomic and take Xπ(Zν)X\subseteq\pi(Z_{\nu}^{\star}) with ν(X)>0\nu(X)>0. By Remark 4.27 and with similar arguments as in the proof of Theorem 3.2, it follows that the lim supϵ0\limsup_{\epsilon\to 0} in Definition 4.14 is actually a limit.

5.2. Proof of Theorem 1.2

Let XkX\subseteq\mathbb{P}^{k} be a closed ff-invariant set. Define

(5.1) DDX+(f) . . =sup{VD(ν):νX+(f)}\operatorname{DD}^{+}_{X}(f)\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\sup\big{\{}\operatorname{VD}(\nu)\colon\nu\in\mathcal{M}_{X}^{+}(f)\big{\}}

and recall that δX(f)\delta_{X}(f), PX+(t)P^{+}_{X}(t), and pX+(f)p^{+}_{X}(f) are defined in Sections 2.3 and 2.4. Theorem 1.2 follows from the following proposition applied with X=J(f)X=J(f).

Proposition 5.2.

We have pX+(f)=2DDX+(f)p_{X}^{+}(f)=2\operatorname{DD}_{X}^{+}(f). In particular, the set {t:PX+(t)=0}\big{\{}t\colon P_{X}^{+}(t)=0\big{\}} is non-empty.

Proof.

We first prove the inequality pX+(f)2DDX+(f)p_{X}^{+}(f)\geq 2\operatorname{DD}_{X}^{+}(f). We can assume that DDX+(f)>0\operatorname{DD}_{X}^{+}(f)>0. Fix 0<t<2DDX+(f)0<t<2\operatorname{DD}_{X}^{+}(f). By the definition (5.1) of DDX+(f)\operatorname{DD}_{X}^{+}(f), there exists νX+(f)\nu\in\mathcal{M}^{+}_{X}(f) such that VD(ν)>t/2\operatorname{VD}(\nu)>t/2. Since VD(ν)=(2Lν(f))1hν(f)\operatorname{VD}(\nu)=(2L_{\nu}(f))^{-1}h_{\nu}(f) by Theorem 1.1, we have hν(f)/Lν(f)>th_{\nu}(f)/L_{\nu}(f)>t. It follows that hν(f)tLν(f)>0h_{\nu}(f)-tL_{\nu}(f)>0; that is, PX+(t)>0P_{X}^{+}(t)>0. Therefore we have pX+(f)>tp_{X}^{+}(f)>t. Since tt is arbitrary, we obtain pX+(f)2DDX+(f)p_{X}^{+}(f)\geq 2\operatorname{DD}_{X}^{+}(f).

Let us now prove that 2DDX+(f)pX+(f)2\operatorname{DD}_{X}^{+}(f)\geq p_{X}^{+}(f). Suppose that 2DDX+(f)<pX+(f)2\operatorname{DD}_{X}^{+}(f)<p_{X}^{+}(f). Then there exists t(2DDX+(f),pX+(f))t\in(2\operatorname{DD}_{X}^{+}(f),p_{X}^{+}(f)) such that PX+(t)>0P_{X}^{+}(t)>0. In particular, there exists a measure νX+(f)\nu\in\mathcal{M}^{+}_{X}(f) with hν(f)tLν(f)>0h_{\nu}(f)-tL_{\nu}(f)>0. We deduce from Theorem 1.1 that

VD(ν)=hν(f)2Lν(f)>t2>DDX+(f).\operatorname{VD}(\nu)=\frac{h_{\nu}(f)}{2L_{\nu}(f)}>\frac{t}{2}>\operatorname{DD}_{X}^{+}(f).

This contradicts the definition of DDX+(f)\operatorname{DD}_{X}^{+}(f). Hence, we have 2DDX+(f)pX+(f)2\operatorname{DD}_{X}^{+}(f)\geq p_{X}^{+}(f).

By Lemma 4.18, we have DDX+(f)1\operatorname{DD}_{X}^{+}(f)\leq 1. Since the function tPX+(t)t\mapsto P_{X}^{+}(t) is convex and non-increasing, the equality pX+(f)=2DDX+(f)2p_{X}^{+}(f)=2\operatorname{DD}_{X}^{+}(f)\leq 2 implies that the set {t:PX+(t)=0}\{t\colon P_{X}^{+}(t)=0\} is non-empty. The proof is complete. ∎

Remark 5.3.

One can also define

DDe+(f)\displaystyle\operatorname{DD}^{+}_{e}(f) . . =sup{VD(ν):νe+(f)},\displaystyle\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\sup\{\operatorname{VD}(\nu)\colon\nu\in\mathcal{M}_{e}^{+}(f)\},
Pe+(t)\displaystyle P^{+}_{e}(t) . . =sup{hν(f)tLν(f):νe+(f)}, and\displaystyle\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\sup\{h_{\nu}(f)-tL_{\nu}(f)\colon\nu\in\mathcal{M}_{e}^{+}(f)\},\quad\mbox{ and }
pe+(f)\displaystyle p^{+}_{e}(f) . . =inf{t:Pe+(f)0},\displaystyle\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\inf\{t:P^{+}_{e}(f)\leq 0\},

where we recall that e+(f)\mathcal{M}^{+}_{e}(f) is the set of ergodic ff-invariant measures whose measure-theoretic entropy is strictly larger than (k1)logd(k-1)\log d. Since e+(f)+(f)\mathcal{M}_{e}^{+}(f)\subseteq\mathcal{M}^{+}(f) for every ff [deT08, Dup12], Theorem 1.1 applies in particular to every νe+(f)\nu\in\mathcal{M}^{+}_{e}(f). The same proof as Proposition 5.2 gives pe+(f)=2DDe+(f)p_{e}^{+}(f)=2\operatorname{DD}_{e}^{+}(f).

5.3. Proof of Theorem 1.3

Recall that if XkX\subset\mathbb{P}^{k} is a uniformly expanding closed invariant set for ff, the volume dimension VD(X)\operatorname{VD}(X) is defined as VD(X) . . =supνX+(f)VDν(X)\operatorname{VD}(X)\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\sup_{\nu\in\mathcal{M}_{X}^{+}(f)}\operatorname{VD}_{\nu}(X); see Definition 4.19.

Proposition 5.4.

Let XkX\subseteq\mathbb{P}^{k} be a uniformly expanding closed invariant set for ff containing a dense orbit. We have δX(f)2VD(X)\delta_{X}(f)\geq 2\operatorname{VD}(X). In particular, if ff is hyperbolic, we have δJ(f)2VD(J(f))\delta_{J}(f)\geq 2\operatorname{VD}(J(f)).

Proof.

We can assume that a volume-conformal measure on XX exists, otherwise we have δX(f)=+\delta_{X}(f)=+\infty and the assertion is trivial. Let μ\mu be a tt-volume-conformal measure on XX, for some tδX(f)t\geq\delta_{X}(f). Since XX contains a dense orbit, we have Supp μ=X\text{Supp }\mu=X. It suffices to prove the inequality t2VDν(X)t\geq 2\operatorname{VD}_{\nu}(X) for any measure νX+(f)\nu\in\mathcal{M}_{X}^{+}(f). By Definition 4.19, this implies that t2VD(X)t\geq 2\operatorname{VD}(X) and the conclusion follows by taking the infimum over tt as above.

Fix νX+(f)\nu\in\mathcal{M}_{X}^{+}(f). We can assume that VDν(X)>0\operatorname{VD}_{\nu}(X)>0, since otherwise the assertion is trivial. In particular, recalling that all measures in +(f)\mathcal{M}^{+}(f) are ergodic, we can assume that ν\nu is non-atomic. Fix a constant γ>1\gamma>1. Since VDν(X)=lim supϵ0VDνϵ(Xϵ)\operatorname{VD}_{\nu}(X)=\limsup_{\epsilon\to 0}\operatorname{VD}_{\nu}^{\epsilon}(X^{\epsilon}) by Definition 4.14 and VDν(X)1\operatorname{VD}_{\nu}(X)\leq 1 by Lemma 4.18, we can fix ϵ0=ϵ0(γ)\epsilon_{0}=\epsilon_{0}(\gamma) such that VDν(X)γVDνϵ0(Xϵ0)\operatorname{VD}_{\nu}(X)\leq\gamma\operatorname{VD}_{\nu}^{\epsilon_{0}}(X^{\epsilon_{0}}). As we can assume that ϵ0(γ)0\epsilon_{0}(\gamma)\to 0 as γ1\gamma\to 1, it is enough to prove that 2VDνϵ0(Xϵ0)t(ϵ0)2\operatorname{VD}_{\nu}^{\epsilon_{0}}(X^{\epsilon_{0}})\leq t{\ell(\epsilon_{0})}, where (ϵ)\ell(\epsilon) is as in (4.4).

By Remark 4.2, for every 0<ϵχmin0<\epsilon\ll\chi_{\min} we have Xϵ=XX^{\epsilon}=X. In particular, Xϵ0=XX^{\epsilon_{0}}=X is compact. Fix α>1\alpha>1. As in the proof of Proposition 4.26, since μ(X)>0\mu(X)>0, for every η>0\eta>0 and 0<κ<r(ϵ)0<\kappa<r(\epsilon) there exists an NN^{\star} (depending on η\eta and κ\kappa) large enough and a cover {Ui}i1𝒰(ϵ,κ,N)\{U_{i}\}_{i\geq 1}\subset\mathcal{U}(\epsilon,\kappa,N^{\star}) of Xϵ0=XX^{\epsilon_{0}}=X satisfying

(5.2) iμ(Ui)αη.\sum_{i}\mu(U_{i})^{\alpha}\leq\eta.

As Xϵ0=XX^{\epsilon_{0}}=X is compact, we can assume that the cover {Ui}\{U_{i}\} is finite. By Lemma 2.13, we have

(5.3) iVol(Ui)tα/2iCtακtkαeαtN(Ui)kϵ(5M+2)m(μ,κeN(Ui)Mϵ)αμ(Ui)αCtακtkαeαtN+kϵ(5M+2)m(μ,κeN+Mϵ)αiμ(Ui)α,\sum_{i}\operatorname{Vol}(U_{i})^{t\alpha/2}\leq\sum_{i}\frac{C^{t\alpha}\kappa^{tk\alpha}e^{\alpha tN(U_{i})k\epsilon(5M+2)}}{m_{-}(\mu,\kappa\,e^{-N(U_{i})M\epsilon})^{\alpha}}\mu(U_{i})^{\alpha}\\ \leq\frac{C^{t\alpha}\kappa^{tk\alpha}e^{\alpha tN^{+}k\epsilon(5M+2)}}{m_{-}(\mu,\kappa\,e^{-N^{+}M\epsilon})^{\alpha}}\sum_{i}\mu(U_{i})^{\alpha},

where m>0m_{-}>0 is as in Lemma 2.12, C<C<\infty is as in Lemma 2.13, and N+N^{+} is the maximum of the N(Ui)N(U_{i}) (we use here that the cover {Ui}\{U_{i}\} is finite). We deduce from (5.2) and (5.3) that

iVol(Ui)tα/2<CtακtkαeαtN+kϵ(5M+2)m(μ,κeN+Mϵ)αη<.\sum_{i}\operatorname{Vol}(U_{i})^{t\alpha/2}<\frac{C^{t\alpha}\kappa^{tk\alpha}e^{\alpha tN^{+}k\epsilon(5M+2)}}{m_{-}(\mu,\kappa\,e^{-N^{+}M\epsilon})^{\alpha}}\eta<\infty.

This implies that VDνϵ0,κ(Xϵ0)tα/2\operatorname{VD}^{\epsilon_{0},\kappa}_{\nu}(X^{\epsilon_{0}})\leq t\alpha/2 for all α>1\alpha>1. Taking α1\alpha\searrow 1, we have VDνϵ0,κ(Xϵ0)t/2\operatorname{VD}^{\epsilon_{0},\kappa}_{\nu}(X^{\epsilon_{0}})\leq t/2 for all 0<κ<r(ϵ)0<\kappa<r(\epsilon). By Lemma 4.12, we have VDνϵ0(Xϵ0)(ϵ0)t/2\operatorname{VD}^{\epsilon_{0}}_{\nu}(X^{\epsilon_{0}})\leq\ell(\epsilon_{0})t/2. The proof is complete. ∎

The following result implies Theorem 1.3 by taking X=J(f)X=J(f) if ff is hyperbolic (since J(f)J(f) has dense orbits).

Theorem 5.5.

Let ff be an endomorphism of k\mathbb{P}^{k} of algebraic degree d2d\geq 2 and XkX\subseteq\mathbb{P}^{k} a closed invariant uniformly expanding set containing a dense orbit. Then

δX(f)=pX+(f)=2VD(X)\delta_{X}(f)=p^{+}_{X}(f)=2\operatorname{VD}(X)

and there exists a unique invariant probability measure μX\mu_{X} supported on XX and such that VD(μX)=VD(X)\operatorname{VD}(\mu_{X})=\operatorname{VD}(X).

Proof.

It follows from the general theory of thermodynamic formalism for uniformly expanding systems (see for instance [Bow75] and [PU10, Chapters 3 and 6]) that, for every tt\in\mathbb{R}, there exists a unique invariant probability measure μt\mu_{t} on XX maximizing the pressure function PX(t)=supν{hν(f)tLν(f)}P_{X}(t)=\sup_{\nu}\{h_{\nu}(f)-tL_{\nu}(f)\}, where the supremum is taken over all invariant measures ν\nu on XX. We used here the fact that, since XX is uniformly expanding, the function tlog|Jacf|-t\log|\operatorname{Jac}f| is Hölder continuous on XX.

Let μX\mu_{X} be the invariant measure μt0\mu_{t_{0}} associated to t0=pX+(f)t_{0}=p^{+}_{X}(f). As hμX(f)=pX+(f)LμX(f)h_{\mu_{X}}(f)=p^{+}_{X}(f)L_{\mu_{X}}(f), it follows from Theorem 1.1 that we have pX+(f)=2VD(μX)p^{+}_{X}(f)=2\operatorname{VD}(\mu_{X}). Since VD(μX)VDμX(X)\operatorname{VD}(\mu_{X})\leq\operatorname{VD}_{\mu_{X}}(X) by Definition 4.17 and VD(X) . . =supνX+(f)VDν(X)\operatorname{VD}(X)\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\sup_{\nu\in\mathcal{M}_{X}^{+}(f)}\operatorname{VD}_{\nu}(X) by Definition 4.19, we have

VD(μX)VD(X).\operatorname{VD}(\mu_{X})\leq\operatorname{VD}(X).

We deduce from the above and Proposition 5.4 that

pX+(f)=2VD(μX)2VD(X)δX(f).p^{+}_{X}(f)=2\operatorname{VD}(\mu_{X})\leq 2\operatorname{VD}(X)\leq\delta_{X}(f).

To complete the proof, we prove the inequality δX(f)pX+(f)\delta_{X}(f)\leq p^{+}_{X}(f) by constructing a tt^{\star}-volume conformal measure on XX for some tpX+(f)t^{\star}\leq p^{+}_{X}(f). Since one can follow Patterson’s [Pat76] and Sullivan’s [Sul83] constructions of conformal measures, we only sketch the proof and refer to those papers for more details; see also [PU10, Sections 12.1 and 12.3].

Take xXx\in X. For each m0m\geq 0, set

Em . . =f|Xm(x).E_{m}\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=f|_{X}^{-m}(x).

Then EmE_{m} is finite and Em+1=f|X1(Em)E_{m+1}=f|_{X}^{-1}(E_{m}). For all t0t\geq 0, consider the sequence {am(t)}m1\{a_{m}(t)\}_{m\geq 1} given by

am(t) . . =log(xEmeSmϕt(x))a_{m}(t)\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\log\Big{(}\sum_{x\in E_{m}}e^{S_{m}\phi_{t}(x)}\Big{)}

where ϕt(x) . . =tlog|Jacf(x)|\phi_{t}(x)\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=-t\log|\operatorname{Jac}f(x)| and Smϕt(x) . . =j=0m1(ϕtfj)(x)S_{m}\phi_{t}(x)\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\sum_{j=0}^{m-1}(\phi_{t}\circ f^{j})(x). Let c(t)c(t) be defined as

c(t) . . =lim supmam(t)m.c(t)\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\limsup_{m\to\infty}\frac{a_{m}(t)}{m}.

As a consequence of the expansiveness of f|Xf|_{X} one can prove that

(5.4) c(t)PX(t) for all t0;c(t)\leq P_{X}(t)\quad\mbox{ for all }t\geq 0;

see for instance [PU10, Lemmas 12.2.3 and 12.2.4]. Moreover, the function tc(t)t\mapsto c(t) is continuous. Setting

t . . =inf{t0:c(t)0},t^{\star}\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\inf\{t\geq 0:c(t)\leq 0\},

it follows from (5.4) that tpX+(f)<t^{\star}\leq p_{X}^{+}(f)<\infty.

By [PU10, Lemma 12.1.2], there exists a sequence {bm}m1\{b_{m}\}_{m\geq 1} of positive real numbers such that the quantity

Ms . . =m=1bmeammsM_{s}\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\sum_{m=1}^{\infty}b_{m}e^{a_{m}-ms}

satisfies Ms<M_{s}<\infty for s>ts>t^{\star} and Ms=+M_{s}=+\infty for sts\leq t^{\star}. For s>ts>t^{\star} consider the measure νs\nu_{s} defined as

νs . . =1Msm=1xEmbmeammsδx.\nu_{s}\mathrel{\vbox{\hbox{\scriptsize.}\hbox{\scriptsize.}}}=\frac{1}{M_{s}}\sum_{m=1}^{\infty}\sum_{x\in E_{m}}b_{m}e^{a_{m}-ms}\delta_{x}.

One can check that any weak limit of the measures νs\nu_{s} as sts\searrow t^{\star} is tt^{\star}-volume-conformal; see for instance [PU10, Section 12.1]. The assertion follows. ∎

References

  • [BB18] François Berteloot and Fabrizio Bianchi “Perturbations d’exemples de Lattès et dimension de Hausdorff du lieu de bifurcation” In J. Math. Pures Appl. (9) 116, 2018, pp. 161–173
  • [BBD18] François Berteloot, Fabrizio Bianchi and Christophe Dupont “Dynamical stability and Lyapunov exponents for holomorphic endomorphisms of k\mathbb{P}^{k} In Ann. Sci. Éc. Norm. Supér. (4) 51.1, 2018, pp. 215–262
  • [BD01] Jean-Yves Briend and Julien Duval “Deux caractérisations de la mesure d’équilibre d’un endomorphisme de Pk(){\rm P}^{k}(\mathbb{C}) In Publ. Math. Inst. Hautes Études Sci., 2001, pp. 145–159
  • [BD03] Ilia Binder and Laura DeMarco “Dimension of pluriharmonic measure and polynomial endomorphisms of n\mathbb{C}^{n} In Int. Math. Res. Not., 2003, pp. 613–625
  • [BD05] François Berteloot and Christophe Dupont “Une caractérisation des endomorphismes de Lattès par leur mesure de Green” In Comment. Math. Helv. 80.2, 2005, pp. 433–454
  • [BD19] François Berteloot and Christophe Dupont “A distortion theorem for iterated inverse branches of holomorphic endomorphisms of k\mathbb{P}^{k} In J. Lond. Math. Soc. (2) 99.1, 2019, pp. 153–172
  • [BD22] Fabrizio Bianchi and Tien-Cuong Dinh “Equilibrium states of endomorphisms of k\mathbb{P}^{k} II: spectral stability and limit theorems” In arXiv preprint arXiv:2204.02856, 2022
  • [BD23] Fabrizio Bianchi and Tien-Cuong Dinh “Equilibrium states of endomorphisms of k\mathbb{P}^{k} I: existence and properties” In J. Math. Pures Appl. (9) 172, 2023, pp. 164–201
  • [BD99] Jean-Yves Briend and Julien Duval “Exposants de Liapounoff et distribution des points périodiques d’un endomorphisme de Pk\mathbb{C}{\rm P}^{k} In Acta Math. 182.2, 1999, pp. 143–157
  • [BDM07] Araceli Bonifant, Marius Dabija and John Milnor “Elliptic curves as attractors in 2\mathbb{P}^{2}. I. Dynamics” In Experiment. Math. 16.4, 2007, pp. 385–420
  • [BDM08] François Berteloot, Christophe Dupont and Laura Molino “Normalization of bundle holomorphic contractions and applications to dynamics” In Ann. Inst. Fourier (Grenoble) 58.6, 2008, pp. 2137–2168
  • [BDR23] Fabrizio Bianchi, Tien-Cuong Dinh and Karim Rakhimov “Monotonicity of dynamical degrees for horizontal-like and polynomial-like maps” In arXiv preprint arXiv:2307.10665, 2023
  • [Bes45] A.. Besicovitch “A general form of the covering principle and relative differentiation of additive functions” In Proc. Cambridge Philos. Soc. 41, 1945, pp. 103–110
  • [Bia19] Fabrizio Bianchi “Misiurewicz parameters and dynamical stability of polynomial-like maps of large topological degree” In Math. Ann. 373.3-4, 2019, pp. 901–928
  • [Bil65] Patrick Billingsley “Ergodic theory and information.” John Wiley & Sons, Inc., New York-London-Sydney, 1965, pp. xiii+195
  • [BK83] M. Brin and A. Katok “On local entropy” In Geometric dynamics (Rio de Janeiro, 1981) 1007, Lecture Notes in Math. Springer, Berlin, 1983, pp. 30–38
  • [Bow75] Rufus Bowen “Equilibrium states and the ergodic theory of Anosov diffeomorphisms.” Springer-Verlag, Berlin-New York, 1975, pp. i+108
  • [Bow79] Rufus Bowen “Hausdorff dimension of quasicircles” In Inst. Hautes Études Sci. Publ. Math., 1979, pp. 11–25
  • [BPS99] Luis Barreira, Yakov Pesin and Jörg Schmeling “Dimension and product structure of hyperbolic measures” In Ann. of Math. (2) 149.3, 1999, pp. 755–783
  • [BR22] Fabrizio Bianchi and Karim Rakhimov “Strong probabilistic stability in holomorphic families of endomorphisms of k()\mathbb{P}^{k}(\mathbb{C}) and polynomial-like maps” In arXiv preprint arXiv:2206.04165, 2022
  • [BT17] Fabrizio Bianchi and Johan Taflin “Bifurcations in the elementary Desboves family” In Proc. Amer. Math. Soc. 145.10, 2017, pp. 4337–4343
  • [CFS12] Isaac P. Cornfeld, Sergej V. Fomin and Yakov Grigor’evǐc Sinai “Ergodic theory” Springer Science & Business Media, 2012
  • [CPZ19] Vaughn Climenhaga, Yakov Pesin and Agnieszka Zelerowicz “Equilibrium states in dynamical systems via geometric measure theory” In Bull. Amer. Math. Soc. (N.S.) 56.4, 2019, pp. 569–610
  • [DD04] Tien-Cuong Dinh and Christophe Dupont “Dimension de la mesure d’équilibre d’applications méromorphes” In J. Geom. Anal. 14.4, 2004, pp. 613–627
  • [deT08] Henry deThélin “Sur les exposants de Lyapounov des applications méromorphes” In Invent. Math. 172.1, 2008, pp. 89–116
  • [DR20] Christophe Dupont and Axel Rogue “Dimension of ergodic measures and currents on (2)\mathbb{C}\mathbb{P}(2) In Ergodic Theory Dynam. Systems 40.8, 2020, pp. 2131–2155
  • [DS03] Tien-Cuong Dinh and Nessim Sibony “Dynamique des applications d’allure polynomiale” In J. Math. Pures Appl. (9) 82.4, 2003, pp. 367–423
  • [DS10] Tien-Cuong Dinh and Nessim Sibony “Dynamics in several complex variables: endomorphisms of projective spaces and polynomial-like mappings” In Holomorphic dynamical systems Springer, 2010, pp. 165–294
  • [DU91] Manfred Denker and Mariusz Urbański “On Sullivan’s conformal measures for rational maps of the Riemann sphere” In Nonlinearity 4.2, 1991, pp. 365–384
  • [DU91a] Manfred Denker and Mariusz Urbański “On the existence of conformal measures” In Trans. Amer. Math. Soc. 328.2, 1991, pp. 563–587
  • [Dup11] Christophe Dupont “On the dimension of invariant measures of endomorphisms of k{\mathbb{C}}{\mathbb{P}}^{k} In Math. Ann. 349.3, 2011, pp. 509–528
  • [Dup12] Christophe Dupont “Large entropy measures for endomorphisms of k\mathbb{C}\mathbb{P}^{k} In Israel J. Math. 192.2, 2012, pp. 505–533
  • [dV15] Henry deThélin and Gabriel Vigny “On the measures of large entropy on a positive closed current” In Math. Z. 280.3-4, 2015, pp. 919–944
  • [Gro03] Mikhaïl Gromov “On the entropy of holomorphic maps” In Enseign. Math. (2) 49.3-4, 2003, pp. 217–235
  • [HR92] Franz Hofbauer and Peter Raith “The Hausdorff dimension of an ergodic invariant measure for a piecewise monotonic map of the interval” In Canad. Math. Bull. 35.1, 1992, pp. 84–98
  • [Jon99] Mattias Jonsson “Dynamics of polynomial skew products on 2\mathbb{C}^{2} In Math. Ann. 314.3, 1999, pp. 403–447
  • [Led81] François Ledrappier “Some relations between dimension and Lyapounov exponents” In Comm. Math. Phys. 81.2, 1981, pp. 229–238
  • [LY85] François Ledrappier and Lai-Sang Young “The metric entropy of diffeomorphisms. II. Relations between entropy, exponents and dimension” In Ann. of Math. (2) 122.3, 1985, pp. 540–574
  • [Lyu83] M.. Lyubich “Entropy properties of rational endomorphisms of the Riemann sphere” In Ergodic Theory Dynam. Systems 3.3, 1983, pp. 351–385
  • [Mañ81] Ricardo Mañé “A proof of Pesin’s formula” In Ergodic Theory Dynam. Systems 1.1, 1981, pp. 95–102
  • [Man84] Anthony Manning “The dimension of the maximal measure for a polynomial map” In Ann. of Math. (2), 1984, pp. 425–430
  • [Mañ88] Ricardo Mañé “The Hausdorff dimension of invariant probabilities of rational maps” In Dynamical systems, Valparaiso 1986 1331, Lecture Notes in Math. Springer, Berlin, 1988, pp. 86–117
  • [McM00] Curtis T. McMullen “Hausdorff dimension and conformal dynamics. II. Geometrically finite rational maps” In Comment. Math. Helv. 75.4, 2000, pp. 535–593
  • [Ose68] Valery I. Oseledets “A multiplicative ergodic theorem. Characteristic Ljapunov, exponents of dynamical systems” In Trudy Moskov. Mat. Obšč. 19, 1968, pp. 179–210
  • [Par69] William Parry “Entropy and generators in ergodic theory” W. A. Benjamin, Inc., New York-Amsterdam, 1969, pp. xii+124
  • [Pat76] S.. Patterson “The limit set of a Fuchsian group” In Acta Math. 136.3-4, 1976, pp. 241–273
  • [Pes77] Jakov B. Pesin “Characteristic Ljapunov exponents, and smooth ergodic theory” In Uspehi Mat. Nauk 32.4 (196), 1977, pp. 55–112\bibrangessep287
  • [Prz93] Feliks Przytycki “Lyapunov characteristic exponents are nonnegative” In Proc. Amer. Math. Soc. 119.1, 1993, pp. 309–317
  • [PU10] Feliks Przytycki and Mariusz Urbański “Conformal fractals: ergodic theory methods” 371, London Mathematical Society Lecture Note Series Cambridge University Press, Cambridge, 2010, pp. x+354
  • [Rue78] David Ruelle “An inequality for the entropy of differentiable maps” In Bol. Soc. Brasil. Mat. 9.1, 1978, pp. 83–87
  • [Rue82] David Ruelle “Repellers for real analytic maps” In Ergodic Theory Dynam. Systems 2.1, 1982, pp. 99–107
  • [Sul83] Dennis Sullivan “Conformal dynamical systems” In Geometric dynamics (Rio de Janeiro, 1981) 1007, Lecture Notes in Math. Springer, Berlin, 1983, pp. 725–752
  • [SUZ14] Michał Szostakiewicz, Mariusz Urbański and Anna Zdunik “Stochastics and thermodynamics for equilibrium measures of holomorphic endomorphisms on complex projective spaces” In Monatsh. Math. 174.1, 2014, pp. 141–162
  • [Taf10] Johan Taflin “Invariant elliptic curves as attractors in the projective plane” In J. Geom. Anal. 20.1, 2010, pp. 219–225
  • [UZ13] Mariusz Urbański and Anna Zdunik “Equilibrium measures for holomorphic endomorphisms of complex projective spaces” In Fund. Math. 220.1, 2013, pp. 23–69
  • [Wal00] Peter Walters “An introduction to ergodic theory” Springer Science & Business Media, 2000
  • [You82] Lai-Sang Young “Dimension, entropy and Lyapunov exponents” In Ergodic Theory Dynam. Systems 2.1, 1982, pp. 109–124