1 Introduction

1.1 Main results

The main results of this paper are the following two theorems.

Theorem 1.1

Let \((M,\xi )\) be a closed co-orientable contact manifold. For every \(\varepsilon >0\) there exists a contact form \(\alpha \) on \((M,\xi )\) with volume one such that the topological entropy \(h_{{\text {top}}}(\alpha )\) of its Reeb flow is smaller than \(\varepsilon \).

Given a closed manifold Q let \(h_{{\text {vol}}} (Q)\) be the infimum of the volume entropies of Riemannian metrics on Q that have volume one. This number is equal to \(2 \sqrt{\pi (k-1)}\) for a closed orientable surface of genus \(k \ge 2\), and it is positive for instance if Q admits a Riemannian metric of negative curvature. Given a Finsler metric F on Q we denote by \(h_{{\text {top}}} (F)\) the topological entropy of the time-one map of the geodesic flow of F. Define the dimension constants

$$\begin{aligned} c_n:= \frac{1}{(n!\, \omega _n)^{1/n}} \end{aligned}$$

where \(\omega _n\) is the volume of the Euclidean unit ball in \(\mathbb {R}^n\). For instance \(c_2 = \frac{1}{\sqrt{2\pi }}\), and asymptotically \(c_n \sim \sqrt{\frac{e}{2\pi }} \frac{1}{\sqrt{n}}\).

Theorem 1.2

Let Q be a closed connected n-dimensional manifold. Then for every Finsler metric F on Q of Holmes–Thompson volume one it holds that

$$\begin{aligned} h_{{\text {top}}} (F) \,\ge \, c_n \, h_{{\text {vol}}} (Q), \;\; \end{aligned}$$

and if F is symmetric that

$$\begin{aligned} h_{{\text {top}}} (F) \,\ge \, 2 c_n \, h_{{\text {vol}}} (Q). \end{aligned}$$

In the rest of this introduction, we recall the notions appearing in these theorems, describe in more detail the results proved in this paper, put them into context, and formulate a few open problems they give rise to. We first tell our story for the special case of unit circle bundles over closed orientable surfaces of higher genus. Most ideas are present already for these simple spaces. We keep the presentation informal, referring to the subsequent sections for the precise definitions and arguments.

1.2 The case of unit circle bundles over higher genus surfaces

Let \(Q_k\) be the closed orientable surface of genus \(k \ge 2\). For every Riemannian metric g on \(Q_k\) we consider the geodesic flow \(\phi _g^t\) on the unit circle bundle

$$\begin{aligned} \left\{ (q,v) \in TQ_k \mid g_q(v,v) = 1 \right\} . \end{aligned}$$

A good numerical measure for the complexity of the flow \(\phi _g^t\) is the topological entropy \(h_{{\text {top}}} (g):= h_{{\text {top}}} (\phi _g^1)\). A definition can be found in Appendix A. This is an interesting invariant because it is related to many other complexity measurements of \(\phi ^t_g\), see [60].

For which Riemannian metrics g is \(h_{{\text {top}}} (g)\) minimal? Such a g could then rightly be considered as a best Riemannian metric from a dynamical point of view. Since the topological entropy scales like

$$\begin{aligned} h_{{\text {top}}} (c g) \,=\, \frac{1}{c} \, h_{{\text {top}}} (g), \end{aligned}$$
(1.1)

the problem is meaningful only if one imposes a normalization. We normalize by the Riemannian area and consider the scale invariant quantity

$$\begin{aligned} {\widehat{h}}_{{\text {top}}}(g) \,=\, \sqrt{\textrm{area}_g (Q_k)} \, h_{{\text {top}}} (g). \end{aligned}$$
(1.2)

It is a classical theorem of Dinaburg [42] and Manning [69] that the geodesic flow of any Riemannian metric on \(Q_k\) has positive topological entropy (cf. Appendix A below). Their results do not give a uniform positive lower bound on \({\widehat{h}}_{{\text {top}}}(g)\) nor do they say anything about the minimizers, however. This was achieved in the following remarkable result of Katok [64].

Theorem 1.3

(Katok 1983) For every Riemannian metric g on \(Q_k\) it holds that

$$\begin{aligned} {\widehat{h}}_{{\text {top}}}(g) \,\ge \, 2 \sqrt{\pi (k-1)}. \end{aligned}$$

Moreover, equality holds if and only if g has constant curvature.

Geodesic flows are very special Reeb flows. For our unit circle bundle over \(Q_k\), Reeb flows can be described as follows. We look at the cotangent bundle \(T^*Q_k\) instead of the tangent bundle, endowed with its usual symplectic form \(\omega = d \lambda \), where \(\lambda = \sum _{j=1}^2 p_j dq_j\). Let \(H :T^*Q_k \rightarrow \mathbb {R}\) be a continuous function that is smooth and positive away from the zero-section and fiberwise positively homogenous of degree one: \(H(q, r p) = r H(q,p)\) for all \(r \ge 0\). Then \(S^*(H):= H^{-1} (1)\) is a smooth hypersurface of \(T^*Q_k\) with the property that for each point \(q \in Q_k\) the intersection \(S_q^* (H):= S^*(H) \cap T_q^*Q_k\) with the cotangent plane at q is the smooth boundary of a domain which is starshaped with respect to the origin \(0_q\), see the left drawing in Fig. 1. Denote by \(\phi _H^t\) the restriction of the Hamiltonian flow of H to \(S^*(H)\). The class of these flows are the Reeb flows on our unit circle bundle. This flow is a co-geodesic flow exactly if H restricts on each fiber to the square root of a positive quadratic form. Special shapes of the fibers \(S_q^* (H)\) in \(T_q^*Q_k\) correspond to special Reeb flows:

:

\(\phi _H^t\) is a Riemannian geodesic flow if and only if each \(S_q^* (H)\) is a centrally symmetric ellipse.

:

\(\phi _H^t\) is a reversible Finsler geodesic flow if and only if each \(S_q^* (H)\) is a centrally symmetric closed smooth curve with strictly positive curvature.

(\(\triangle \)):

\(\phi _H^t\) is a Finsler geodesic flow if and only if each \(S_q^* (H)\) is a closed smooth curve with strictly positive curvature.

Here we identified co-Finsler geodesic flows with Finsler geodesic flows via the Legendre transform.

Fig. 1
figure 1

Spheres \(S_q^* (H)\) in \(T_q^*Q_k\) defining (a) a Reeb flow, (b) a Finsler geodesic flow, (c) a reversible Finsler geodesic flow, (d) a Riemannian geodesic flow

Based on [49] it was shown in [68] that the above result of Dinaburg and Manning about Riemannian geodesic flows extends to all Reeb flows:

Theorem 1.4

Every Reeb flow \(\phi _H^t\) on \(S^*(H)\subset T^* Q_k\), \(k\ge 2\), has positive topological entropy.

Does Katok’s theorem also extend to Reeb flows? To make the question meaningful, we again need to normalize. We do this by the symplectic volume of the bounded domain \(D^*(H)\) in \(T^*Q_k\) with boundary \(S^* (H)\), and define the Holmes–Thompson volume of \(Q_k\) associated with H by

$$\begin{aligned} {\text {vol}}_H^{{\text {HT}}} (Q_k) \,=\, \frac{1}{2\pi } \int _{D^*(H)} \omega \wedge \omega . \end{aligned}$$
(1.3)

Then the normalized topological entropy

$$\begin{aligned} {\widehat{h}}_{{\text {top}}}^{{\text {HT}}} (H) :=\, \sqrt{{\text {vol}}_H^{{\text {HT}}}(Q_k)} \; h_{{\text {top}}} (\phi _H^1) \end{aligned}$$

is invariant under scalings of H. In the Riemannian case, this definition agrees with (1.2), since then \({\text {vol}}_H^{{\text {HT}}} (Q_k) = \textrm{area}_g (Q_k)\). The following question was asked in [48, §7.2].

Question 1.5

Is there a positive constant c(k) such that \(\widehat{h}_{{\text {top}}}^{{\text {HT}}} (H) \ge c(k)\) for every Reeb flow on the co-circle bundle over \(Q_k\)?

Let us first try to answer this question in the affirmative for Finsler geodesic flows. Given a Finsler metric F on \(Q_k\), an obvious idea is to find a lower bound for \(\widehat{h}_{{\text {top}}}^{{\text {HT}}}(F)\) by choosing a larger Riemannian metric \(\sqrt{g} \ge F\), cf. (1.1). In general, the topological entropy of geodesic flows is not monotone with respect to the order relation on metrics, however. We therefore pass to a more geometric version of entropy, which is indeed monotone: The volume entropy of F is defined as the exponential growth rate of balls in the universal cover \({\widetilde{Q}}_k\) (which is the plane),

$$\begin{aligned} h_{{\text {vol}}} (F) :=\, \lim _{R \rightarrow \infty } \frac{1}{R} \log {\text {Vol }}(B_{{\tilde{q}}}(R)) \end{aligned}$$
(1.4)

where \({\tilde{q}}\) is any fixed point in \({\widetilde{Q}}_k\), \(B_{\tilde{q}} (R)\) is the ball of radius R about \({\tilde{q}}\) with respect to the lifted Finsler metric, and \({\text {Vol }}\) is the volume with respect to the lift of any smooth area form on \(Q_k\) (see Appendix A for details). It is clear that \(F_1 \ge F_2\) implies

$$\begin{aligned} h_{{\text {vol}}} (F_2) \, \ge \, h_{{\text {vol}}} (F_1). \end{aligned}$$
(1.5)

In the case of a Riemmannian metric g, denoting \(h_{{\text {vol}}}(\sqrt{g})\) simply by \(h_{{\text {vol}}}(g)\), we have that

$$\begin{aligned} h_{{\text {top}}} (g) \, \ge \, h_{{\text {vol}}} (g) \end{aligned}$$
(1.6)

with equality if g has non-positive curvature, as proven by Manning in [69]. His proof of (1.6) readily generalizes to all Finsler metrics, see Appendix A:

$$\begin{aligned} h_{{\text {top}}} (F) \, \ge \, h_{{\text {vol}}} (F). \end{aligned}$$
(1.7)

Let g be a Riemannian metric such that \(\sqrt{g}\ge F\). Using (1.7) and (1.5) we can now estimate

$$\begin{aligned} {\widehat{h}}_{{\text {top}}}^{{\text {HT}}} (F):= & {} \sqrt{{\text {vol}}_F^{{\text {HT}}}(Q_k)} \,h_{{\text {top}}} (F) \\\ge & {} \sqrt{{\text {vol}}_F^{{\text {HT}}}(Q_k)} \, h_{{\text {vol}}} (g) \\= & {} \sqrt{\frac{{\text {vol}}_F^{{\text {HT}}} (Q_k)}{{\text {vol}}_g^{{\text {HT}}} (Q_k)}} \, \, \widehat{h}_{{\text {vol}}} (g). \end{aligned}$$

In [64], Katok actually proved Theorem 1.3 for the normalized volume entropy \({\widehat{h}}_{{\text {vol}}}\) (which by Manning’s theorem implies Theorem 1.3). Hence we obtain

$$\begin{aligned} {\widehat{h}}_{{\text {top}}}^{{\text {HT}}} (F) \, \ge \, \sqrt{\frac{{\text {vol}}_F^{{\text {HT}}} (Q_k)}{{\text {vol}}_g^{{\text {HT}}} (Q_k)}} \, \, 2 \sqrt{\pi (k-1)}. \end{aligned}$$
(1.8)

To get a uniform lower bound for \({\widehat{h}}_{{\text {top}}}^{{\text {HT}}} (F)\) we therefore look for a Riemannian metric g with \(\sqrt{g} \ge F\) that is as close as possible to F in the sense of the Holmes–Thompson volume. We best do this directly in the cotangent bundle \(T^*Q_k\). We thus look at each \(q \in Q_k\) for a centrally symmetric ellipse in \(T_q^* Q_k\) such that, denoting by \(E_q\) the region bounded by it, we have \(E_q \supset D_q^* (F)\) and \(E_q\) is as close to \(D_q^*(F)\) in volume as possible.

If \(D_q^*(F)\) is centrally symmetric, the best choice is Loewner’s outer ellipse. This is the unique centrally symmetric ellipse enclosing \(D_q^*(F)\) which minimizes the value of the area of the region bounded by it, which we denote by \(E(D_q^*(F))\). Here the area \(| \; |\) is taken with respect to any translation invariant measure on the plane \(T_q^*Q_k\). Loewner’s ellipse depends continuously on q, and the largest area ratio

$$\begin{aligned} \frac{| E(D_q^*(F)) | }{ |D_q^*(F)|} \end{aligned}$$

is \(\frac{\pi }{2}\), which is attained exactly when \(D_q^*(F)\) is a parallelogram. If we take the Riemannian metric g on \(Q_k\) that has the sets \(E(D_q^*(F))\) as unit co-disks, we therefore obtain

$$\begin{aligned} {\text {vol}}_F^{{\text {HT}}} (Q_k) \,\ge \, \frac{2}{\pi } {\text {vol}}_g^{{\text {HT}}} (Q_k). \end{aligned}$$

Together with (1.8) this yields

$$\begin{aligned} {\widehat{h}}_{{\text {top}}}^{{\text {HT}}} (F) \,\ge \, \sqrt{\frac{2}{\pi }} \, 2 \sqrt{\pi (k-1)} \,=\, 2 \sqrt{2 (k-1)}. \end{aligned}$$

If \(D_q^* (F)\) is not centrally symmetric, we observe that the convex hull

$$\begin{aligned} \textrm{conv }\bigl ( D_q^*(F) \cup - D_q^*(F) \bigr ) \end{aligned}$$

is centrally symmetric. It is not hard to see that for every convex body \(K \subset \mathbb {R}^2\) that contains the origin,

$$\begin{aligned} \left| \textrm{conv }(K \cup -K) \right| \,\le \, 4 |K| \end{aligned}$$

with equality attained exactly by the triangles with one vertex at the origin. Therefore,

$$\begin{aligned} \frac{|E (\textrm{conv }(K \cup -K))|}{|K|} \,\le \, 4 \cdot \frac{\pi }{2} \ \,=\, 2\pi . \end{aligned}$$

Note that the constant \(2\pi \) is sharp and is attained exactly by the triangles with one vertex at the origin, see Fig. 2 (b).

Fig. 2
figure 2

The symmetrization \(\textrm{conv }(K \cup -K)\)

Since the two maps

$$\begin{aligned} K \,\mapsto \, \textrm{conv }(K \cup -K) \,\mapsto \, E(K \cup -K) \end{aligned}$$

are continuous, we can take as g the Riemannian metric with unit co-disks

$$\begin{aligned} E \bigl ( \textrm{conv }\bigl ( D_q^*(F) \cup - D_q^*(F) \bigr ) \bigr ) \end{aligned}$$

and obtain

$$\begin{aligned} {\widehat{h}}_{{\text {top}}}^{{\text {HT}}} (F) \,\ge \, \frac{1}{\sqrt{2\pi }}\, 2 \sqrt{\pi (k-1)} \,=\, \sqrt{2(k-1)}. \end{aligned}$$

Summarizing, we obtain Theorem 1.2 for orientable surfaces:

$$\begin{aligned} {\widehat{h}}_{{\text {top}}}^{{\text {HT}}} (F) \,\ge \, \frac{1}{\sqrt{2\pi }} \, h_{{\text {vol}}} (Q_k), \quad \text{ and } \quad {\widehat{h}}_{{\text {top}}}^{{\text {HT}}} (F) \ge \sqrt{\frac{2}{\pi }} \, h_{{\text {vol}}} (Q_k) \, \text{ if } F \text{ is } \text{ symmetric. } \end{aligned}$$
(1.9)

How sharp are these lower bounds? It is still unknown whether the constants \(\frac{1}{\sqrt{2\pi }}\) and \(\sqrt{\frac{2}{\pi }}\) can be replaced by 1. In other words, it is unknown whether there exist Finsler metrics F on \(Q_k\) such that \({\widehat{h}}_{{\text {vol}}}^{{\text {HT}}} (F) < h_{{\text {vol}}} (Q_k)\). We shall say more on this “minimal entropy problem" in Sect. 1.7.

Recall that for the closed orientable surfaces \(Q_k\) of genus k one has

$$\begin{aligned} h_{{\text {vol}}} (Q_k) = 2 \sqrt{\pi (k-1)}. \end{aligned}$$

For the non-orientable surface \(P_k\) whose orientation cover is \(Q_k\) this implies

$$\begin{aligned} h_{{\text {vol}}} (P_k) = \sqrt{2 \pi (k-1)}. \end{aligned}$$

For the other four closed surfaces (the sphere, the torus, the real projective plane and the Klein bottle) Theorem 1.2 is not useful since \(h_{{\text {vol}}}\) vanishes, and in fact there exist geodesic flows on these surfaces with vanishing topological entropy.

We now look at general Reeb flows on the co-circle bundle over \(Q_k\). As said earlier, these flows correspond to Hamiltonian flows on \(S^*(H) = H^{-1}(1)\) of Hamiltonian functions \(H :T^*Q_k \rightarrow \mathbb {R}\) that are fiberwise homogeneous of degree one. Looking for a lower bound for \({\widehat{h}}_{{\text {top}}}^{{\text {HT}}} (\phi _H)\), we proceed as in the case of Finsler geodesic flows, but knowing already (1.9) we now compare H with any Finsler metric. Choose a Finsler Hamiltonian \(F :T^*Q_k \rightarrow \mathbb {R}\) such that \(D^*(H) \subset D^*(F)\), i.e., \(F \le H\).

Definition (1.4) can be extended to Reeb flows: Fix a point \(q \in Q_k\), take a lift \({\tilde{q}} \in {\widetilde{Q}}_k\) of q and the lift \({\widetilde{H}} :T^*{\widetilde{Q}}_k \rightarrow \mathbb {R}\) of H, and then define \(h_{{\text {vol}}}(H,q)\) as the exponential growth rate of the volume of the set \(B_{{\tilde{q}}}({\widetilde{H}}, T)\) of those points \(z \in {\widetilde{Q}}_k\) for which the fiber \(S^*_z(\widetilde{H})\) can be reached in time \(\le T\) by a flow line of \(\phi _{{\widetilde{H}}}^t\) that starts at the fiber \(S^*_{\tilde{q}}({\widetilde{H}})\). As we shall show in Appendix A one then still has Manning’s inequality,

$$\begin{aligned} h_{{\text {top}}}(H) \,\ge \, h_{{\text {vol}}}(H,q). \end{aligned}$$

We now wish to show that there is a constant \(c>0\) depending only on H and F such that \(h_{{\text {vol}}}(H,q) \ge c\, h_{{\text {vol}}}(F)\). The existence of such a constant for non-convex H does not follow from geometric considerations, since it is not true in general that \(F \le H\) implies the inclusion of balls \(B_{{\tilde{q}}} (\widetilde{F},T) \subset B_{{\tilde{q}}}({\widetilde{H}}, T)\). However, using Lagrangian Floer homology in \(T^*Q_k\) one can avoid passing through \(h_{{\text {vol}}}(H, q)\) and prove directly that

$$\begin{aligned} \widehat{h}_{{\text {top}}}^{{\text {HT}}} (H) \,\ge \, \frac{1}{\sigma (H;F)} \, {\widehat{h}}_{{\text {vol}}} (F) \end{aligned}$$
(1.10)

where \(\sigma (H;F)\) is the smallest real number such that \(\frac{1}{\sigma (H;F)} D^*(F) \subset D^*(H)\), cf. Fig. 3. This is explained in Sect. 3, using the proof of the above Theorem 1.4 from [68].

Fig. 3
figure 3

The co-disks \(\frac{1}{\sigma (H;F)} D^*_q (F) \subset D^*_q (H) \subset D^*_q (F)\) in \(T_q^*Q_k\)

The number \(\sigma (H):= \inf \{ \sigma (H;F) \mid F \le H \}\) is a measure for how far the fibers of \(D^*(H)\) are from being convex. Inequalities (1.10) and (1.9) and Katok’s inequality imply

Proposition 1.6

For every Reeb flow \(\phi _H\) on the co-circle bundle over \(Q_k\),

$$\begin{aligned} \widehat{h}_{{\text {top}}}^{{\text {HT}}} (H) \,\ge \, \frac{1}{\sigma (H)} \, \sqrt{2 (k-1)}. \end{aligned}$$
(1.11)

The following special case of Theorem 1.1 shows that the lower bound in (1.11) cannot be made uniform, that the answer to Question 1.5 is ‘no’, and that there is no way to extend Katok’s rigidity theorem to Reeb flows.

Theorem 1.7

For every \(\varepsilon >0\) there exists a Reeb flow \(\phi _H^t\) on \(S^*(H)\) with \({\widehat{h}}_{{\text {top}}}^{{\text {HT}}} (H) \le \varepsilon \).

Proposition 1.6 shows that this entropy collapse cannot happen unless at least some of the co-disks \(D_q^*(H) = D^*(H) \cap T_q^*Q_k\) are very far from convex. Writing down explicitely such star fields on \(T^*Q_k\) that lead to small \({\widehat{h}}_{{\text {top}}}^{{\text {HT}}}\) seems difficult, however. In fact, our proof of Theorem 1.7 does not use the special fibration structure of \(S^*(H)\), but uses the existence of open book decompositions valid for all closed 3-manifolds, see the beginning of Sect. 4.2 for an outline and Sect. 4 for the proof.

1.3 Entropy rigidity for Finsler geodesic flows

Proceeding as in the previous section, one readily arrives at Theorem 1.2 for Finsler geodesic flows on closed manifolds Q of arbitrary dimension n,

$$\begin{aligned} {\widehat{h}}_{{\text {top}}}^{{\text {HT}}} (F) \,\ge \, c_n \, h_{{\text {vol}}} (Q), \quad \text{ and } \quad {\widehat{h}}_{{\text {top}}}^{{\text {HT}}} (F) \ge 2 c_n \, h_{{\text {vol}}} (Q) \, \text{ if } F \text{ is } \text{ symmetric. } \end{aligned}$$
(1.12)

Here the normalization \({\widehat{h}}_{{\text {top}}}^{{\text {HT}}} (F) = \left( {\text {vol}}_F^{{\text {HT}}}(Q)\right) ^{1/n} \, h_{{\text {top}}}(F)\) is done in terms of the Holmes–Thompson volume

$$\begin{aligned} {\text {vol}}_F^{{\text {HT}}} (Q) :=\, \frac{1}{n!\, \omega _n} \int _{D^*(F)} \omega ^n, \end{aligned}$$
(1.13)

which extends definition (1.3). The proof of (1.12) uses Loewner’s outer ellipsoids and the Roger–Shephard volume bounds for symmetrized convex bodies. Similar arguments appear in [6], where they are used to derive systolic inequalities for Finsler metrics from the analogous inequalities for Riemannian metrics.

For Finsler metrics there is another natural volume, the Busemann–Hausdorff volume. For reversible Finsler metrics, this volume is at least the Holmes–Thompson volume, see Sect. 2. The second inequality in (1.12) thus also holds true if we normalize by the Busemann–Hausdorff volume. For irreversible Finsler metrics, however, we do not know whether the first inequality in (1.12) holds true for the Busemann–Hausdorff volume.

For manifolds of dimension \(n \ge 3\), it is more difficult to understand the volume entropy \(h_{{\text {vol}}}(Q)\) than for surfaces. The only sharp result is the following extension of Katok’s theorem.

Theorem 1.8

(Besson–Courtois–Gallot [20, 21]) If Q is a closed manifold of dimension at least 3 that admits a locally symmetric Riemannian metric \(g_0\) of negative curvature, then

$$\begin{aligned} {\widehat{h}}_{{\text {vol}}}(g) \,\ge \, {\widehat{h}}_{{\text {vol}}} (g_0) \end{aligned}$$

for every Riemannian metric g on Q, and equality holds if and only if g is also locally symmetric. In particular, \(h_{{\text {vol}}}(Q) = {\widehat{h}}_{{\text {vol}}} (g_0) >0\).

Note that the space of minimizers up to isometry in Katok’s theorem is the \(6k-6\) dimensional Teichmüller space, while the minimizers in Theorem 1.8 are all isometric up to scaling, by Mostow’s theorem.

In the context of Theorem 1.2 we wish to know when \(h_{{\text {vol}}} (Q) >0\). The main tool for proving \(h_{{\text {vol}}} (Q) >0\) is the simplicial volume \(\Vert Q\Vert \). If Q is orientable, it is defined as \(\inf \sum _i |r_i|\) where the infimum is taken over those sums \(\sum _i r_i \sigma _i\) that represent the fundamental class \([Q] \in H_n (Q;\mathbb {R})\) with real coefficients. If Q is not orientable, pass to the orientation double covering \({\widehat{Q}}\) and put \(\Vert Q\Vert = \frac{1}{2} \Vert {\widehat{Q}}\Vert \). Gromov proved in [59] that

$$\begin{aligned} h_{{\text {vol}}} (Q) \,\ge \, C_n^{-1} \, \Vert Q\Vert ^{1/n} \end{aligned}$$

for an explicit dimension constant \(C_n\).

There are many more manifolds Q with positive simplicial volume \(\Vert Q\Vert \) than those in Theorem 1.8. Indeed, \(\Vert Q\Vert >0\) for all manifolds that admit a Riemannian metric of negative curvature, and positivity of the simplicial volume is preserved under taking the product with any other closed manifold of positive simplicial volume and under taking the connected sum with any other closed manifold of the same dimension. We refer to [59] and [66] for more examples and information on simplicial volume.

1.4 Entropy collapse for Reeb flows

Reeb flows are flows naturally associated to contact manifolds. A contact structure \(\xi \) on a (\(2n-1\))-dimensional manifold M is a maximally non-integrable hyperplane field of the tangent bundle TM. We assume throughout that \(\xi \) is co-orientable, i.e., \(\xi = \ker \alpha \) for a 1-form \(\alpha \) on M. In terms of such a form \(\alpha \), called a contact form for \(\xi \), the maximal non-integrability means that \(\alpha \wedge (d\alpha )^{n-1}\) is a volume form on M. For any non-vanishing function f on M the 1-form \(f \alpha \) is also a contact form on \((M,\xi )\). Each contact form \(\alpha \) gives rise to the Reeb flow \(\phi _\alpha ^t\), which is generated by the Reeb vector field \(R_\alpha \) implicitly defined by the two conditions

$$\begin{aligned} d\alpha (R_\alpha , \cdot ) =0, \qquad \alpha (R_\alpha ) = 1. \end{aligned}$$

For every closed manifold Q the so-called spherization \((S^*Q,\xi _{\textrm{can}\,})\) is a contact manifold whose Reeb flows are exactly the flows \(\phi _H^t\) on \(S^*(H)\) described in Sect. 1.2 in the case of closed surfaces \(Q_k\), see Appendix B.1. Every closed 3-manifold admits infinitely many non-isotopic contact structures, and an odd-dimensional closed manifold M admits a contact structure if and only if its stabilized tangent bundle \(TM \oplus \mathbb {R}\) admits a complex structure [22].

Theorem 1.4 has been extended to many contact manifolds: First, for many closed manifolds Q every Reeb flow on \((S^*Q,\xi _{\textrm{can}\,})\) has positive topological entropy, [68]. Second, there are many closed 3-dimensional manifolds M such that for every contact structure \(\xi \) on M every Reeb flow has positive topological entropy, [7,8,9,10, 74]. For a recent result for non-degenerate Reeb flows see [33].

While in these results the underlying manifolds have rich loop space topology, there are also examples where the positivity of topological entropy of all Reeb flows does not come from the topological complexity of the loop space. For instance, it is shown in [11] that the standard smooth sphere of dimension \(2n-1 \ge 5\) admits a contact structure for which every Reeb flow has positive topological entropy.

Nevertheless, for none of these contact manifolds there can be a uniform bound for the normalized topological entropy: The contact volume of the co-oriented contact manifold \((M,\alpha )\) of dimension \(2n-1\) is defined as

$$\begin{aligned} {\text {vol}}_{\alpha }(M) :=\, \frac{1}{n!\,\omega _n} \int _M \alpha \wedge (d\alpha )^{n-1}. \end{aligned}$$

Now define the normalized topological entropy of the Reeb flow \(\phi _\alpha ^t\) by

$$\begin{aligned} {\widehat{h}}_{{\text {top}}}(\alpha ) :=\, ({\text {vol}}_{\alpha }(M))^{1/n} \, h_{{\text {top}}} (\phi _{\alpha }^1). \end{aligned}$$
(1.14)

This normalization extends the normalizations (1.3) and (1.13) to all contact manifolds, see Appendix B.1. The following result implies Theorem 1.1.

Theorem 1.9

Let \((M,\xi )\) be a closed co-orientable contact manifold of dimension at least three. Then for every real number \(c >0\) there exists a contact form \(\alpha \) for \(\xi \) such that \(\widehat{h}_{{\text {top}}}(\alpha ) = c\).

We shall in fact prove the flexibility expressed in Theorem 1.9 for a larger growth rate: Given a \(C^1\)-diffeomorphism \(\phi \) of a compact manifold M, we define the two real numbers

$$\begin{aligned} \Gamma _+(\phi )&:= \lim _{n \rightarrow +\infty } \frac{1}{n} \log \Vert d \phi ^n \Vert _{\infty } \,, \\ \Gamma (\phi )&:= \max \left\{ \Gamma _+(\phi ), \Gamma _+(\phi ^{-1}) \right\} . \end{aligned}$$

Here \(\Vert \cdot \Vert _{\infty }\) denotes the supremum norm induced by a Riemannian metric on M, but the above limit, whose existence follows from the subadditivity of the sequence \(\log \Vert d\phi ^n\Vert _{\infty }\), is clearly independent of the choice of the metric.

The quantity \(\Gamma _+\) was used by Yomdin [96] to measure the difference between topological entropy and volume growth, and the study of the growth type of the sequence \(\Vert d \phi ^n \Vert _\infty \) for various classes of diffeomorphisms was proposed in [39, §7.10]. The more symmetric invariant \(\Gamma \) and its polynomial version were investigated, for instance, in [82, 83]. For Hamiltonian flows and Reeb flows, where uniform measurements (like the Hofer metric) turned out to capture symplectic rigidity, it is particularly natural to look at these two growth rates.

The norm growths \(\Gamma _+\) and \(\Gamma \) are related to the topological entropy by

$$\begin{aligned} h_{{\text {top}}} (\phi ) \,\le \, (\dim M) \, \Gamma _+(\phi ) \,\le \, (\dim M) \, \Gamma (\phi ), \end{aligned}$$
(1.15)

see [60, Corollary 3.2.10] for the first inequality. The numbers \(\Gamma _+ (\phi )\) and \(\Gamma (\phi )\) are upper bounds for several other invariants of \(\phi \), and hence the collapsibility of \(\Gamma \) for Reeb flows also implies the collapsibility of these other invariants. For instance, \(\Gamma _+ (\phi )\) is not less than the largest Lyapunov exponent \(\chi _{\max } (p)\) at every point \(p \in M\). With

$$\begin{aligned} \Sigma (p) = \sum _{\chi _i^+ (p)>0} k_i^+(p) \, \chi _i^+ (p) \end{aligned}$$

the sum of the positive Lyapunov exponents \(\chi _i^+(p)\) at p counted with their multiplicities \(k_i^+(p)\), we then also have \(\Sigma (p) \le (\dim M) \, \Gamma _+ (\phi )\). Together with the Margulis–Ruelle inequality (see [60, Theorem S.2.13]) we obtain that the metric entropy \(h_{\mu }(\phi )\) with respect to any invariant Borel probability measure \(\mu \) has the upper bound

$$\begin{aligned} h_\mu (\phi ) \,\le \, \int _M \Sigma (p)\, d\mu (p) \,\le \, (\dim M) \, \Gamma _+ (\phi ). \end{aligned}$$

Applying the variational principle for the topological entropy, we obtain again (1.15).

From now on we focus on \(\Gamma \). For a flow \(\phi ^t\) we set \(\Gamma (\phi ) = \Gamma (\phi ^1)\), and for a Reeb flow \(\phi _\alpha ^t\) we set \(\Gamma (\alpha ) = \Gamma (\phi _\alpha )\). For \(c>0\) we have \(\phi _{c \alpha }^t = \phi _\alpha ^{t/c}\) and hence \(\Gamma (c \alpha ) = \frac{1}{c} \Gamma (\alpha )\). Like for the topological entropy, the invariant

$$\begin{aligned} {\widehat{\Gamma }} (\alpha ) \,=\, {\text {vol}}_{\alpha } (M)^{1/n} \, \Gamma (\alpha ), \end{aligned}$$

where \(\dim M = 2n-1\), is therefore invariant under scaling. In view of (1.15), the following result improves Theorem 1.1.

Theorem 1.10

Let \((M,\xi )\) be a closed co-orientable contact manifold of dimension at least three. Then for every real number \(c>0\) there exists a contact form \(\alpha \) for \(\xi \) such that \(\widehat{\Gamma }(\alpha ) = c\).

We shall prove Theorems 1.9 and 1.10 along the following lines. The main step is to show that for every \(\varepsilon >0\) there exists a contact form \(\alpha _{\varepsilon }\) for \(\xi \) such that \({\widehat{\Gamma }}(\alpha _{\varepsilon }) \le \varepsilon \). We do this with the help of an open book decomposition of M and an inductive construction, in which the induction step \(\dim 2n-1 \leadsto \dim 2n+1\) is carried out by applying the induction hypothesis to the binding of the open book decomposition of M. We can start the induction in dimension 1 at the circle, for which \({\widehat{\Gamma }} (d\theta ) =0\). We nevertheless present the 3-dimensional case separately in Sect. 4 because we believe that after understanding the geometric ideas in this particular situation it is easier to follow the general argument. The induction step is done in Sect. 6. It uses results of Giroux on the correspondence between contact structures and supporting open books, that we recollect in Sect. 5.

Given contact forms \(\alpha _{\varepsilon }\) as above, Theorems 1.9 and 1.10 follow from (1.15) and from a simple modification of \(\alpha _{\varepsilon }\) that increases \({\widehat{h}}_{{\text {top}}}\) and \({\widehat{\Gamma }}\), see Sect. 7.

1.5 Collapsing the growth rate of symplectic invariants

In the works [9, 11, 74] it is shown that the exponential growth rate of certain symplectic topological invariants provides a lower bound for the topological entropy of Reeb flows. These invariants are linearised Legendrian contact homology [9], wrapped Floer homology [11], and Rabinowitz–Floer homology [74]. Combining these results with Theorem 1.1 we obtain that the growth rate of these invariants can be made arbitrarily small. Details are given in Sect. 8.

1.6 Relations to systolic inequalities

Consider a closed co-orientable contact manifold \((M,\xi )\) of dimension \(2n-1\). Given a contact form \(\alpha \) for \(\xi \) that has at least one periodic Reeb orbit, take the smallest period \(T_{\min }(\alpha )\). The so-called systolic ratio

$$\begin{aligned} \rho _{\textrm{sys}} (\alpha ) \,=\, {\text {vol}}_{\alpha } (M)^{- 1/n} \, T_{\min } (\alpha ) \end{aligned}$$

is then invariant under scalings of \(\alpha \).

While for spherizations \(S^*Q\) of many closed manifolds Q there are famous uniform upper bounds on the systolic ratios of Riemannian Reeb flows, in the full class of Reeb flows one has the following flexibility result.

Theorem 1.11

For any closed co-orientable contact manifold \((M,\xi )\) and every positive number c there exists a contact form \(\alpha \) such that \(\rho _{\textrm{sys}} (\alpha ) > c\).

This result was shown for the tight 3-sphere in [1] and for all contact 3-manifolds in [2] by a plug construction in open book decompositions. The idea in this paper to use open book decompositions for proving entropy collapse came from these works. Theorem 1.11 in dimension \(\ge 5\) was proved in [88]. That proof was later on much simplified [89] by using our inductive construction in Sect. 6. Interestingly, our construction in dimension 3 does not yield a proof of Theorem 1.11. This suggests that at least in the smallest interesting dimension, one has more flexibility to collapse the topological entropy of Reeb flows than to increase their systolic ratio.

1.7 Minimal entropy problems for Finsler and Reeb flows

Given a class \({{\mathcal {C}}}\) of maps on a compact manifold M, it is interesting to understand which maps in class \({{\mathcal {C}}}\) minimize the (normalized) topological entropy. Since topological entropy is a measure for the complexity, these maps can then be considered as the simplest, or the best, maps on M in class \({{\mathcal {C}}}\).

For the class of Riemannian geodesic flows on the spherization SQ of a compact manifold Q, the minimal entropy problem consists of three parts.

  1. (P1)

    Compute the minimal entropy

    $$\begin{aligned} h_{{\text {top}}}(Q, {{\mathcal {G}}}) :=\, \inf \left\{ {\widehat{h}}_{{\text {top}}} ( \phi _g) \mid g\text { a Riemannian metric on } Q \right\} . \end{aligned}$$
  2. (P2)

    Decide whether the infimum is attained or not.

  3. (P3)

    If the infimum is attained, describe the minimizers g.

For manifolds admitting a locally symmetric Riemannian metric of negative curvature, Theorems 1.3 and 1.8 completely solve the minimal entropy problem. Among the many further interesting works on the minimal entropy problem are [65, 81].

The minimal entropy problem can also be formulated for the larger classes of Finsler and Reeb flows. Define three more numbers

$$\begin{aligned} h_{{\text {top}}}^{{\text {HT}}} (Q,{{\mathcal {R}}}) \,\le \, h_{{\text {top}}}^{{\text {HT}}} (Q,{{\mathcal {F}}}) \,\le \, h_{{\text {top}}}^{{\text {HT}}} (Q, {{\mathcal {F}}}_{{\text {rev}}}) \,\le \, h_{{\text {top}}} (Q,{{\mathcal {G}}}) \end{aligned}$$

by taking the infimum in the definition of \(h_{{\text {top}}} (Q,{{\mathcal {G}}})\) over all contact forms on \((S^*Q,\xi _{\textrm{can}\,})\) for \(h_{{\text {top}}}^{{\text {HT}}} (Q,{{\mathcal {R}}})\), over all Finsler metrics for \(h_{{\text {top}}}^{{\text {HT}}} (Q,{{\mathcal {F}}})\), and over all reversible Finsler metrics for \(h_{{\text {top}}}^{{\text {HT}}} (Q,{{\mathcal {F}}}_{{\text {rev}}})\), respectively, where as in (1.13) and  (1.14) we normalize by the Holmes–Thompson volume.

Theorem 1.1 shows that \(h_{{\text {top}}}^{{\text {HT}}} (Q,{{\mathcal {R}}}) =0\) for all compact manifolds Q. This settles (P1) for the class \({{\mathcal {R}}}\). Furthermore, for many manifolds, like those with fundamental group of exponential growth, the answer to (P2) is ‘no’ by the general version of Theorem 1.4 from [68].

We now turn to the invariants \(h_{{\text {top}}}^{{\text {HT}}} (Q,{{\mathcal {F}}})\) and \(h_{{\text {top}}}^{{\text {HT}}} (Q,{{\mathcal {F}}}_{{\text {rev}}})\). By Theorems 1.2 and A.2 we have

$$\begin{aligned} \begin{array}{rclcl} c_n\, h_{{\text {vol}}} (Q) &{}\le &{} h_{{\text {top}}}^{{\text {HT}}} (Q,{{\mathcal {F}}}) &{}\le &{} h_{{\text {top}}} (Q, {{\mathcal {G}}}), \\[0.8em] 2 c_n\, h_{{\text {vol}}} (Q) &{}\le &{} h_{{\text {top}}}^{{\text {HT}}} (Q,{{\mathcal {F}}}_{{\text {rev}}}) &{}\le &{} h_{{\text {top}}} (Q, {{\mathcal {G}}}). \end{array} \end{aligned}$$

For manifolds admitting a locally symmetric Riemannian metric of negative curvature (for which \(h_{{\text {top}}} (Q, {{\mathcal {G}}}) = h_{{\text {vol}}}(Q)\)) nothing more seems to be known about the values of \(h_{{\text {top}}}^{{\text {HT}}} (Q,{{\mathcal {F}}})\) and \(h_{{\text {top}}}^{{\text {HT}}} (Q,{{\mathcal {F}}}_{{\text {rev}}})\), so already (P1) in the entropy problem is wide open for the classes \({{\mathcal {F}}}\) and \({{\mathcal {F}}}_{{\text {rev}}}\).

Addressing (P3) we note that in the Finsler setting one cannot expect metrics of minimal normalized topological entropy to be unique, or even to be characterised in terms of curvature-like invariants. Indeed, any exact symplectomorphism of \(T^*Q\) that is \(C^2\)-close to the identity maps the unit cotangent sphere bundle \(S^*(F)\) of the Finsler metric F to the unit cotangent sphere bundle \(S^*(F')\) of some Finsler metric \(F'\) whose geodesic flow is conjugated to the one of F by a smooth time-preserving conjugacy. In particular, the new Finsler metric \(F'\) has the same normalized topological entropy as F, but need not be isometric to it. See Appendix D for a discussion of this.

Higher rank. More can be said in higher rank. The following result is proved in Sect. 2.6 using Verovic’s work [93].

Proposition 1.12

Let (Qg) be a compact locally symmetric space of non-compact type and of rank \(\ge 2\). Then there exists a constant \(c<1\) such that

$$\begin{aligned} h_{{\text {top}}}^{{\text {HT}}} (Q,{{\mathcal {F}}}_{{\text {rev}}}) \,\le \, c\, {\widehat{h}}_{{\text {vol}}} (g). \end{aligned}$$
(1.16)

The constant c only depends on the globally symmetric space \(({\widetilde{Q}}, {\widetilde{g}})\), and it can be computed from its Weyl data. See Proposition 2.7 below for a more precise statement.

Let \(h_{{\text {vol}}}^{{\text {sym}}}(Q)\) be the minimum of \({\widehat{h}}_{{\text {vol}}}(g)\) taken over all locally symmetric Riemannian metrics g on Q. This number is easy to compute, see [35, §2]. Unfortunately it is still not known whether Theorem 1.8 also holds in higher rank, that is, whether \(h_{{\text {vol}}} (Q) = h_{{\text {vol}}}^{{\text {sym}}}(Q)\). However, this is known if (Qg) is locally isometric to a product of negatively curved symmetric spaces of dimension \(\ge 3\), [35], and for quotients of the k-fold product \((\mathbb {H}^2)^k = \mathbb {H}_2 \times \dots \times \mathbb {H}_2\) of the real hyperbolic plane, [75]. For these spaces, (1.16) can thus be written as

$$\begin{aligned} h_{{\text {top}}}^{{\text {HT}}} (Q,{{\mathcal {F}}}_{{\text {rev}}}) \,\le \, c\, h_{{\text {vol}}}(Q). \end{aligned}$$

We shall compute the constant c for quotients of \((\mathbb {H}^2)^k\) in Sect. 2.6. For instance, \(c (\mathbb {H}^2 \times \mathbb {H}^2) = \root 4 \of {2} \approx 0.841\). This should be compared with the constant \(2 c_4 = \sqrt{\frac{2}{\pi }} \approx 0.61\) for the lower bound in Theorem 1.1.

The minimal entropy problem can also be studied for the volume entropies \(h_{{\text {vol}}}\) instead of \(h_{{\text {top}}}\), and by normalizing either entropy by the Busemann–Hausdorff volume. Much of the above discussion applies also to these minimal entropies.

1.8 Topological pressure

In view of Theorem 1.1, there is no minimal entropy program for Reeb flows. Furthermore, the situation cannot be salvaged by looking at subexponential growth rates, since replacing \(\lim _{n \rightarrow \infty } \frac{1}{n} \log \dots \) in the definition of topological entropy by \(\lim _{n \rightarrow \infty } \frac{1}{n^c} \log \dots \) for some \(c \in (0,1)\) yields \(+ \infty \) for all Reeb flows on many contact manifolds by Theorem 1.4.

However, increasing topological entropy in terms of topological pressure leads to a meaningful problem. Given a closed contact manifold \((M,\xi )\) associate with every continuous function \(f \in C^0(M,\mathbb {R})\) and every contact form \(\alpha \) for \(\xi \) the topological pressure \(P(\alpha , f) = P (\phi _{\alpha },f)\), see [95, Chapter 9] for the definition and basic results on topological pressure. We recall that \(P(\alpha ,0) = h_{{\text {top}}}(\alpha )\) and that the variational principle for topological pressure says

$$\begin{aligned} P(\alpha ,f) \,=\, \sup _{\mu \in {{\mathcal {M}}}(\alpha )} \left\{ h_{\mu }(\phi _{\alpha }) + \int _M f \,d\mu \right\} \end{aligned}$$
(1.17)

where \({{\mathcal {M}}}(\alpha )\) denotes the set of \(\phi _{\alpha }^t\)-invariant Borel probability measures on M and \(h_\mu \ge 0\) is the entropy of the measure \(\mu \). Define

$$\begin{aligned} P (M,f) :=\, \inf \left\{ P(\alpha , f) \mid \alpha \text{ a } \text{ normalized } \text{ contact } \text{ form } \text{ on } (M,\xi ) \right\} . \end{aligned}$$

Since \(P(\alpha , f+c) = P(\alpha , f) +c\) for all \(c \in \mathbb {R}\), we can assume that \(\min f = 0\). Together with Theorem 1.1 we then obtain

$$\begin{aligned} 0 \,\le \, P(M,f) \,\le \, \max f. \end{aligned}$$

It would be interesting to see if these bounds can be sharpened for functions f that do not identically vanish. Our proof of Theorem 1.1 does not help with this problem, since the maximal measures in (1.17) (the so-called equilibrium states) may not be related in any way to the open book decomposition in our proof.

2 Volume entropy for Finsler geodesic flows

2.1 Finsler metrics and their volumes

By a Finsler metric on an n-dimensional manifold Q we mean in this paper a continuous function \(F :TQ \rightarrow [0,+\infty )\) which is fiberwise convex, fiberwise positively homogeneous of degree 1, and positive outside of the zero section. The Finsler metric F is said to be reversible if \(F(v)=F(-v)\) for all \(v \in TQ\).

For \(q \in Q\) the unit disk in \(T_q Q\) determined by the Finsler metric F is the set

$$\begin{aligned} D_q(F):= \left\{ v \in T_qQ \mid F(v) \le 1 \right\} . \end{aligned}$$

This is a convex compact neighborhood of the origin in \(T_q Q\). The function \(F|_{T_q Q}\) is precisely the Minkowski gauge of \(D_q(F)\). The unit co-disk in \(T_q^* Q\) is the polar set of \(D_q(F)\):

$$\begin{aligned} D_q^*(F):= \left\{ p \in T_q^* Q \mid \langle p,v \rangle \le 1 \; \forall v \in D_q(F) \right\} , \end{aligned}$$

where \(\langle \cdot ,\cdot \rangle \) denotes the duality pairing between tangent vectors and co-vectors. This is a compact convex neighborhood of the origin in \(T_q^* Q\).

On compact Finsler manifolds there are two different notions of volume that are used in the literature. From the point of view of this paper, the most natural one is the Holmes–Thompson volume, which can be defined as

$$\begin{aligned} {\text {vol}}_F^{{\text {HT}}}(Q):= \frac{1}{n! \, \omega _n} \, \int _{D^*(F)} \omega ^n, \end{aligned}$$

where

$$\begin{aligned} D^*(F):= \bigcup _{q \in Q} D_q^*(F) \,\subset \, T^* Q \end{aligned}$$

is the unit co-disk bundle of Q, where \(\omega ^n\) denotes the standard volume form on \(T^*Q\) induced by integrating the n-fold exterior power of the canonical symplectic form \(\omega = \sum _j dq_j \wedge dp_j\), and where \(\omega _n\) is the volume of the Euclidean unit ball in \(\mathbb {R}^n\), \(n=\dim Q\). The normalization factor \(n! \,\omega _n\) makes \({\text {vol}}_F^{{\text {HT}}}(Q)\) coincide with the Riemannian volume of Q when \(F(v) = \sqrt{g(v,v)}\) is a Riemannian metric on Q.

Alternatively, the Holmes–Thompson volume can be defined as the integral over Q of a suitable volume density \(\rho _F^*\). Here by volume density we mean a norm on the line bundle \(\Lambda ^n(TQ)\), whose fiber at \(q \in Q\) is the top degree component of the exterior algebra of \(T_q Q\), that is, the 1-dimensional space spanned by \(v_1\wedge \dots \wedge v_n\), where \(v_1,\dots ,v_n\) is a basis of \(T_q Q\). When Q is orientable, a volume density is just the absolute value of a nowhere vanishing differential n-form. A volume density can be integrated over any non-empty open subset of Q, producing a positive number. The volume density \(\rho _F^*\) is defined as follows: Given any volume density \(\rho \) on Q set

$$\begin{aligned} \rho _F^*(q):= \frac{|D_q^*(F)|_{\rho }^*}{\omega _n} \,\rho (q), \end{aligned}$$

where \(|\! \cdot \!|^*_{\rho }\) denotes the Lebesgue measure on \(T_q^* Q\) that is normalized to 1 on the n-dimensional parallelogram spanned by the covectors that are dual to basis vectors \(v_1,\dots ,v_n\) in \(T_q Q\) such that \(\rho (q)[v_1 \wedge \dots \wedge v_n]=1\). We then have

$$\begin{aligned} {\text {vol}}_F^{{\text {HT}}}(Q) = \int _Q \rho _F^*. \end{aligned}$$

Another common choice is to consider the Busemann–Hausdorff volume, which is defined as

$$\begin{aligned} {\text {vol}}_F^{{\text {BH}}}(Q):= \int _Q \rho _F, \end{aligned}$$

where the volume density \(\rho _F\) is given by

$$\begin{aligned} \rho _F(q):= \frac{\omega _n}{|D_q(F)|_{\rho }} \, \rho (q). \end{aligned}$$

Here \(\rho \) is again an arbitrary volume density on Q and \(| \! \cdot \! |_{\rho }\) is the Lebesgue measure on \(T_q Q\) normalized to 1 on the parallelogram spanned by vectors \(v_1,\dots ,v_n\) in \(T_q Q\) with \(\rho (q)[v_1 \wedge \dots \wedge v_n]=1\). When F is reversible, the Busemann–Hausdorff volume of Q coincides with the n-dimensional Hausdorff measure of Q with respect to the distance induced by F.

Both volumes reduce to the standard Riemannian volume when the Finsler metric F is Riemannian. If F is reversible, then

$$\begin{aligned} {\text {vol}}_F^{{\text {HT}}}(Q) \le {\text {vol}}_F^{{\text {BH}}}(Q), \end{aligned}$$
(2.1)

with equality holding if and only if F is Riemannian. This follows from the Blaschke–Santaló inequality, see e.g. [43]. In the non-reversible case, the Holmes–Thompson volume can be much larger than the Busemann–Hausdorff volume. Note that both the Holmes–Thompson and the Busemann–Hausdorff volume depend monotonically on the Finsler metric, meaning that

$$\begin{aligned} F_1 \le F_2 \qquad \Rightarrow \qquad {\text {vol}}_{F_1}^{{\text {HT}}}(Q) \le {\text {vol}}_{F_2}^{{\text {HT}}}(Q), \quad {\text {vol}}_{F_1}^{{\text {BH}}}(Q) \le {\text {vol}}_{F_2}^{{\text {BH}}}(Q), \end{aligned}$$
(2.2)

and rescale as

$$\begin{aligned} {\text {vol}}_{cF}^{{\text {HT}}}(Q) = c^n {\text {vol}}_F^{{\text {HT}}}(Q), \qquad {\text {vol}}_{cF}^{{\text {BH}}}(Q) = c^n {\text {vol}}_F^{{\text {BH}}}(Q), \end{aligned}$$
(2.3)

when the Finsler metric F is multiplied by a positive constant c.

2.2 Volume entropy

Let F be a Finsler metric on a compact n-dimensional manifold Q. This Finsler metric lifts to a Finsler metric on the universal cover \(\widetilde{Q}\) of Q, and we denote the lifted metric by the same symbol F. The R-ball centered at \(q \in \widetilde{Q}\) that is induced by F is the following compact subset of \(\widetilde{Q}\):

$$\begin{aligned} B_q(F,R):= \bigl \{ \gamma (R) \mid \gamma :[0,R] \rightarrow \widetilde{Q} \text{ Lipschitz } \text{ curve, } \gamma (0)=q \text{ and } F \circ {\dot{\gamma }} \le 1 \text{ a.e. } \bigr \}. \end{aligned}$$
(2.4)

When F is reversible, \(B_q(F,R)\) is the ball of the distance on \(\widetilde{Q}\) that is induced by F; in general, it is the forward ball of an asymmetric distance.

The volume entropy of F is the non-negative number

$$\begin{aligned} h_{{\text {vol}}}(F):= \lim _{R\rightarrow \infty } \frac{1}{R} \log {\text {Vol }}(B_q(F,R)). \end{aligned}$$
(2.5)

Here \({\text {Vol }}\) denotes the volume of Borel subsets of \(\widetilde{Q}\) with respect to the lift to \(\widetilde{Q}\) of an arbitrary Riemannian metric on Q. A minor modification of Manning’s argument from [69] shows that the above limit exists and is independent of the choice of the point \(q \in \widetilde{Q}\) and of the Riemannian metric on Q, see Proposition A.1. In the case of the Finsler metric \(G=\sqrt{g(\cdot ,\cdot )}\) that is induced by a Riemannian metric g, we use interchangeably the notations

$$\begin{aligned} h_{{\text {vol}}}(g) = h_{{\text {vol}}}(G). \end{aligned}$$

The volume entropy is monotonically decreasing in F, meaning that

$$\begin{aligned} F_1 \le F_2 \qquad \Rightarrow \qquad h_{{\text {vol}}}(F_1) \ge h_{{\text {vol}}}(F_2). \end{aligned}$$
(2.6)

Indeed, if \(F_1 \le F_2\) on Q then the same inequality holds on \(\widetilde{Q}\) and hence (2.4) implies

$$\begin{aligned} B_q(F_1,R) \supset B_q(F_2,R) \qquad \forall \, q \in \widetilde{Q}, \; \forall R \ge 0, \end{aligned}$$

from which (2.6) follows. Let c be a positive number. From the identity

$$\begin{aligned} B_q(cF,R) = B_q(F,c^{-1} R) \end{aligned}$$

we deduce that the volume entropy rescales as

$$\begin{aligned} h_{{\text {vol}}}(cF) = c^{-1} \,h_{{\text {vol}}}(F). \end{aligned}$$
(2.7)

Together with (2.3), this suggests to consider the normalized volume entropies

$$\begin{aligned} \widehat{h}_{{\text {vol}}}^{{\text {HT}}}(F):= {\text {vol}}_F^{{\text {HT}}}(Q)^{1/n}\, h_{{\text {vol}}}(F), \qquad \widehat{h}_{{\text {vol}}}^{{\text {BH}}}(F):= {\text {vol}}_F^{{\text {BH}}}(Q)^{1/n}\, h_{{\text {vol}}}(F). \end{aligned}$$

These quantities are now invariant under scaling:

$$\begin{aligned} \widehat{h}_{{\text {vol}}}^{{\text {HT}}}(cF) = \widehat{h}_{{\text {vol}}}^{{\text {HT}}}(F), \qquad \widehat{h}_{{\text {vol}}}^{{\text {BH}}}(cF) = \widehat{h}_{{\text {vol}}}^{{\text {BH}}}(F). \end{aligned}$$

Since the Holmes–Thompson and the Busemann–Hausdorff volumes coincide when \(F=G=\sqrt{g}\) is Riemannian, there is just one normalized volume entropy in the Riemannian case, and we denote it by

$$\begin{aligned} \widehat{h}_{{\text {vol}}}(g) = \widehat{h}_{{\text {vol}}}(G). \end{aligned}$$

In the next two subsections, we study how the two different normalized volume entropies of an arbitrary Finsler metric can be bounded from below and from above in terms of the normalized volume entropy of suitable Riemannian metrics. Our arguments follow [6], where similar techniques are used in order to derive bounds for the systolic ratio.

2.3 From reversible Finsler to Riemannian

Let F be a reversible Finsler metric on the compact n-dimensional manifold Q. Denote by \(E_q\) the inner Loewner ellipsoid of the symmetric convex body \(D_q(F)\), i.e. the ellipsoid centered at the origin which is contained in \(D_q(F)\) and has maximal volume among all ellipsoids with this property. Here by volume we mean any translation invariant measure on \(T_q Q\) (which is unique up to multiplication by a positive constant). It is well known that the inner Loewner ellipsoid is unique, and John proved that it satisfies

$$\begin{aligned} E_q \subset D_q(F) \subset \sqrt{n}\, E_q. \end{aligned}$$
(2.8)

See [62], or [12] for a modern proof of these results. Denote by \(G :TQ \rightarrow [0,+\infty )\) the function which in each tangent space \(T_q Q\) is the Minkowski gauge of \(E_q\). The function G is the square root of a Riemannian metric: \(G(v) = \sqrt{g(v,v)}\) for some continuous Riemannian metric g on Q. Indeed, the continuity of G easily follows from the uniqueness of the inner Loewner ellipsoid. From the inclusions (2.8) we deduce the inequalities

$$\begin{aligned} n^{-1/2}\, G \le F \le G, \end{aligned}$$
(2.9)

which thanks to (2.6) and (2.7) imply the bounds

$$\begin{aligned} h_{{\text {vol}}}(G) \le h_{{\text {vol}}}(F) \le \sqrt{n}\, h_{{\text {vol}}}(G). \end{aligned}$$
(2.10)

By (2.2) and the second inequality in (2.9) the Busemann–Hausdorff volume of (QF) has the upper bound

$$\begin{aligned} {\text {vol}}_F^{{\text {BH}}} (Q) \le {\text {vol}}_G (Q). \end{aligned}$$
(2.11)

In order to get a lower bound for the Holmes–Thompson volume of (QF) we can use (2.2), (2.3) and the first inequality in (2.9) and obtain

$$\begin{aligned} {\text {vol}}_G (Q) \le n^{n/2} {\text {vol}}_F^{{\text {HT}}} (Q). \end{aligned}$$
(2.12)

However, we get a better bound by the following argument. The polar set \(E_q^*=D_q^*(G)\) of \(E_q=D_q(G)\) satisfies

$$\begin{aligned} D_q^*(F) \subset D_q^*(G) \end{aligned}$$

and is the outer Loewner ellipsoid of \(D_q^*(F)\), i.e. the centrally symmetric ellipsoid of minimal volume among those containing \(D_q^*(F)\). Then we have

$$\begin{aligned} |D_q^* (G)|_{\rho }^* \,\le \, \frac{n! \, \omega _n}{2^n} \, |D_q^*(F)|_{\rho }^*. \end{aligned}$$

This follows from the fact that the ratio between the volume of the outer Loewner ellipsoid of a symmetric convex body K and the volume of K is maximal for the cross-polytope, a result that Ball deduced from the inverse Brascamp–Lieb inequality of Barthe in [13, Theorem 5]. Therefore, we obtain

$$\begin{aligned} {\text {vol}}_G(Q) \,\le \, \frac{n! \,\omega _n}{2^n} {\text {vol}}_F^{{\text {HT}}}(Q), \end{aligned}$$
(2.13)

which is a better bound than (2.12) for every \(n \ge 2\), and also asymptotically because

$$\begin{aligned} \lim _{n \rightarrow \infty } \left( \frac{n!\, \omega _n}{2^n \, n^{n/2}} \right) ^{1/n} \,=\, \sqrt{\frac{\pi }{2e}} \end{aligned}$$

by the Stirling formula. By putting together (2.1), (2.10), (2.11) and (2.13) we obtain the following result.

Proposition 2.1

Let F be a reversible Finsler metric on the compact n-dimensional manifold Q and let \(G = \sqrt{g}\) be the Riemannian metric on Q whose unit disks are the inner Loewner ellipsoids of the unit disks of F. Then

$$\begin{aligned} \frac{2}{(n!\, \omega _n)^{1/n}} \, \widehat{h}_{{\text {vol}}}(g) \,\le \, \widehat{h}_{{\text {vol}}}^{{\text {HT}}}(F) \,\le \, \widehat{h}_{{\text {vol}}}^{{\text {BH}}}(F) \,\le \, \sqrt{n} \; \widehat{h}_{{\text {vol}}}(g). \end{aligned}$$

2.4 From irreversible to reversible Finsler

Let F be an arbitrary Finsler metric on the compact n-dimensional manifold Q. We symmetrize the metric F by the following procedure: We define \(S :TQ \rightarrow [0,+\infty )\) to be the reversible Finsler metric on Q whose unit co-disk at each \(q \in Q\) is the reflection body of \(D_q^*(F)\), i.e. the centrally symmetric convex body

$$\begin{aligned} D^*_q(S):= \textrm{conv} \bigl ( D_q^*(F) \cup ( - D_q^*(F)) \bigr ). \end{aligned}$$

Note that

$$\begin{aligned} D_q^*(F) \subset D^*_q(S) \subset \theta \, D_q^*(F), \end{aligned}$$
(2.14)

where \(\theta \) is the irreversibility ratio of F, i.e. the number

$$\begin{aligned} \theta := \max _{\begin{array}{c} v\in TQ \\ F(v)=1 \end{array}} F(-v), \end{aligned}$$
(2.15)

which is at least 1, and equal to 1 if and only if F is reversible. Indeed, the second inclusion in (2.14) follows from the fact that \(\theta \) is an upper bound for the norm of minus the identity on \(T_q Q\) with the asymmetric norm F, and hence also for the norm of minus the identity on \(T_q^* Q\) with the asymmetric norm that is dual to F. Moreover, the volume of \(D^*_q(S)\) has the upper bound

$$\begin{aligned} |D^*_q(S)|_{\rho }^* \,\le \, 2^n \, |D_q^*(F)|_{\rho }^*, \end{aligned}$$
(2.16)

as proven by Rogers and Shephard in [86, Theorem 3]. From (2.14) we deduce

$$\begin{aligned} \theta ^{-1} S \le F \le S, \end{aligned}$$
(2.17)

and hence (2.6) and (2.7) imply

$$\begin{aligned} h_{{\text {vol}}}(S) \le h_{{\text {vol}}} (F) \le \theta \, h_{{\text {vol}}}(S). \end{aligned}$$
(2.18)

On the other hand, from (2.16) and the second inequality in (2.17) we obtain the following inequalities for the Holmes–Thompson volume

$$\begin{aligned} 2^{-n} \, {\text {vol}}_S^{{\text {HT}}} (Q) \le {\text {vol}}_F^{{\text {HT}}} (Q) \le {\text {vol}}_S^{{\text {HT}}} (Q). \end{aligned}$$
(2.19)

The bounds (2.18) and (2.19) imply the following result.

Proposition 2.2

Let F be a Finsler metric on the compact n-dimensional manifold Q with irreversibility ratio \(\theta \) and let S be the reversible Finsler metric whose dual disks are the reflection bodies of the dual disks of F:

$$\begin{aligned} D^*_q(S) = \textrm{conv} \bigl ( D_q^*(F) \cup ( - D_q^*(F)) \bigr ) \qquad \forall \, q \in Q. \end{aligned}$$

Then

$$\begin{aligned} \frac{1}{2} \, \widehat{h}^{{\text {HT}}}_{{\text {vol}}}(S) \,\le \, \widehat{h}_{{\text {vol}}}^{{\text {HT}}} (F) \,\le \, \theta \, \widehat{h}_{{\text {vol}}}^{{\text {HT}}}(S). \end{aligned}$$

The lower bounds of Propositions 2.1 and 2.2, together with Stirling’s formula, have the following consequence:

Corollary 2.3

Let Q be a compact n-dimensional manifold and denote by \(h_{{\text {vol}}}(Q)\) the infimum of \(\widehat{h}_{{\text {vol}}}(g)\) over all Riemannian metrics g on Q. Then the Holmes–Thompson normalized volume entropy of an arbitrary Finsler metric F on Q has the lower bound

$$\begin{aligned} \widehat{h}_{{\text {vol}}}^{{\text {HT}}}(F) \,\ge \, c_n \, h_{{\text {vol}}}(Q), \end{aligned}$$

where

$$\begin{aligned} c_n:= \frac{1}{(n!\, \omega _n)^{1/n}} \sim \sqrt{\frac{e}{2\pi }} \frac{1}{\sqrt{n}}. \end{aligned}$$

Moreover, if the Finsler metric F is reversible, we have

$$\begin{aligned} \widehat{h}_{{\text {vol}}}^{{\text {BH}}}(F) \ge \widehat{h}_{{\text {vol}}}^{{\text {HT}}}(F) \ge 2 c_n \, h_{{\text {vol}}}(Q). \end{aligned}$$

Remark 2.4

If we symmetrize \(D_q^*(F)\) by considering the difference body \(D_q^*(F)-D_q^*(F)\) instead of the reflection body, then we get a worse bound, because in this case the factor \(2^n\) in (2.16) must be replaced by the middle binomial coefficient \(\left( {\begin{array}{c}2n\\ n\end{array}}\right) \), in view of the Rogers–Shephard inequality for the volume of the difference body, see [85]. By using the reflection body instead of the difference body, the systolic upper bounds of Theorem 4.13 and Corollary 4.14 in [6] can be improved by replacing the dimension dependent quantity \(\root n \of {(2n)!/(n!)^2}\) by the constant factor 2.

If the volume entropy is normalized by the Busemann–Hausdorff volume, we do not get a lower bound that is independent of the irreversibility ratio. From (2.2), (2.3), (2.17) and (2.18) we obtain

$$\begin{aligned} \frac{1}{\theta }\, \widehat{h}_{{\text {vol}}}^{{\text {BH}}}(S) \,\le \, \widehat{h}_{{\text {vol}}}^{{\text {BH}}}(F) \,\le \, \theta \, \widehat{h}_{{\text {vol}}}^{{\text {BH}}}(S). \end{aligned}$$

We do not have a lower bound that is independent of \(\theta \) because, unlike the volume ratio \(|D_q^*(S)|_{\rho }^*/ |D_q^*(F)|_{\rho }^*\), the ratio \(|D_q(F)|_{\rho }/|D_q(S)|_{\rho }\) can be arbitrarily large.

On the other hand, the upper bound can be made independent of the irreversibility ratio \(\theta \) by symmetrizing, this time, directly in TQ: We consider the reversible Finsler metric T whose unit ball at q is the set

$$\begin{aligned} D_q(T):= \textrm{conv} \bigl ( D_q(F) \cup (- D_q(F)) \bigr ). \end{aligned}$$

For this metric, we have

$$\begin{aligned} T \le F \le \theta \, T, \end{aligned}$$

from which we obtain

$$\begin{aligned} \frac{1}{\theta } \, h_{{\text {vol}}}(T) \le h_{{\text {vol}}}(F)\le h_{{\text {vol}}} (T). \end{aligned}$$

Moreover, the Rogers–Shephard inequality for the reflection body gives

$$\begin{aligned} {\text {vol}}_T^{{\text {BH}}} (Q) \le {\text {vol}}_F^{{\text {BH}}} (Q) \le 2^n {\text {vol}}_T^{{\text {BH}}} (Q), \end{aligned}$$

and we deduce the following result.

Proposition 2.5

Let F be a Finsler metric on the compact n-dimensional manifold Q with irreversibility ratio \(\theta \) and let T be the reversible Finsler metric whose unit disk at each q is the reflection body

$$\begin{aligned} D_q(T) = \textrm{conv} \bigl ( D_q(F) \cup (- D_q(F)) \bigr ) \end{aligned}$$

of the disk of F at q. Then

$$\begin{aligned} \frac{1}{\theta } \, \widehat{h}^{{\text {BH}}}_{{\text {vol}}}(T) \,\le \, \widehat{h}_{{\text {vol}}}^{{\text {BH}}} (F) \,\le \, 2 \, \widehat{h}_{{\text {vol}}}^{{\text {BH}}}(T). \end{aligned}$$

2.5 Lower bounds on the normalized topological entropy

We now assume that the (possibly irreversible) Finsler metric F on Q has better regularity and convexity properties: Outside of the zero section, \(F :TQ \rightarrow \mathbb {R}\) is of class \(C^{2}\) and the fiberwise second differential of \(F^2\) is positive definite. We will refer to such an F as to a regular Finsler metric. Under these assumptions, the geodesic flow of F is well defined. We denote by \(h_{{\text {top}}}(F)\) the topological entropy of this flow, and by

$$\begin{aligned} \widehat{h}_{{\text {top}}}^{{\text {HT}}}(F) = {\text {vol}}_F^{{\text {HT}}}(Q)^{1/n} \, h_{{\text {top}}}(F), \qquad \widehat{h}_{{\text {top}}}^{{\text {BH}}}(F) = {\text {vol}}_F^{{\text {BH}}}(Q)^{1/n} \, h_{{\text {top}}}(F) \end{aligned}$$

the Holmes–Thompson and Busemann–Hausdorff normalizations of this entropy. Manning’s inequality

$$\begin{aligned} h_{{\text {top}}}(F) \ge h_{{\text {vol}}}(F) \end{aligned}$$

from [69] holds also in the Finsler setting, as shown in Theorem A.2. Then Corollary 2.3 has the following immediate consequence.

Corollary 2.6

Let Q be a compact n-dimensional manifold and denote by \(h_{{\text {vol}}}(Q)\) the infimum of \(\widehat{h}_{{\text {vol}}}(g)\) over all Riemannian metrics g on Q. Then the Holmes–Thompson normalized topological entropy of any regular Finsler metric F on Q has the lower bound

$$\begin{aligned} \widehat{h}_{{\text {top}}}^{{\text {HT}}}(F) \,\ge \, c_n \, h_{{\text {vol}}}(Q), \end{aligned}$$

where

$$\begin{aligned} c_n:= \frac{1}{(n! \, \omega _n)^{1/n}} \sim \sqrt{\frac{e}{2\pi }} \frac{1}{\sqrt{n}}. \end{aligned}$$

Moreover, if the Finsler metric F is reversible, we have

$$\begin{aligned} \widehat{h}_{{\text {top}}}^{{\text {BH}}}(F) \ge \widehat{h}_{{\text {top}}}^{{\text {HT}}}(F) \ge 2 c_n \, h_{{\text {vol}}}(Q). \end{aligned}$$

2.6 Finsler metrics with small topological entropy

The following result is more precise than Proposition 1.12.

Proposition 2.7

Let \(({\widetilde{Q}}, {\widetilde{g}})\) be a Riemannian globally symmetric space of non-compact type and of rank \(\ge 2\). Let G be the connected component of the identity of the isometry group of \(({\widetilde{Q}}, {\widetilde{g}})\). Then there exist computable constants \(c^{{\text {HT}}}< c^{{\text {BH}}} < 1\) that depend only on \((\widetilde{Q}, {\widetilde{g}})\) with the following property: For every discrete co-compact subgroup \(\Gamma \) of G that acts without fixed points on \({\widetilde{Q}}\) and for every \(\varepsilon >0\) there exists a smooth reversible G-invariant Finsler metric F on \(Q = {\widetilde{Q}} / \Gamma \) such that

$$\begin{aligned} {\widehat{h}}_{{\text {top}}}^{{\text {HT}}}(F)= & {} {\widehat{h}}_{{\text {vol}}}^{{\text {HT}}}(F) \;\le \; (1+\varepsilon )\; c^{{\text {HT}}} \; {\widehat{h}}_{{\text {vol}}} ({\widetilde{g}}), \end{aligned}$$
(2.20)
$$\begin{aligned} {\widehat{h}}_{{\text {top}}}^{{\text {BH}}}(F)= & {} {\widehat{h}}_{{\text {vol}}}^{{\text {BH}}}(F) \;\le \; (1+\varepsilon )\; c^{{\text {BH}}} \; {\widehat{h}}_{{\text {vol}}} ({\widetilde{g}}). \end{aligned}$$
(2.21)

In particular, \({\widehat{h}}_{{\text {top}}}^{{\text {HT}}}(Q, {{\mathcal {F}}}_{{\text {rev}}}) \le c^{{\text {HT}}} \; {\widehat{h}}_{{\text {vol}}} ({\widetilde{g}})\) and \({\widehat{h}}_{{\text {top}}}^{{\text {BH}}}(Q, {{\mathcal {F}}}_{{\text {rev}}}) \le c^{{\text {BH}}} \; {\widehat{h}}_{{\text {vol}}} ({\widetilde{g}})\).

Proof

Fix a point \(x_0 \in G\), let \(K \subset G\) be the stabilizer of \(x_0\), let \({\mathfrak {g}}\) and \({\mathfrak {k}}\) be the Lie algebras of G and K, and let \({\mathfrak {g}} = {\mathfrak {k}} \oplus {\mathfrak {p}}\) be the Cartan decomposition associated with \(x_0\). (Then \({\mathfrak {p}} \cong T_{x_0} {\widetilde{Q}}\).) Choose a maximal abelian subalgebra \({\mathfrak {a}} \subset {\mathfrak {p}}\), and let \(W_{{\mathfrak {a}}}\) be its Weyl group.

The set of G-invariant Finsler metrics on \({\widetilde{Q}}\) is in bijection with the set

$$\begin{aligned} {{\mathcal {C}}}\,=\, \left\{ C \subset {\mathfrak {a}} \mid C \text{ centrally } \text{ symmetric } W_{{\mathfrak {a}}}\text{-invariant } \text{ convex } \text{ body } \right\} . \end{aligned}$$

In general, the Finsler metric assoiated with \(C \in {{\mathcal {C}}}\) is only continuous, and it is smooth if and only if the boundary of C is smooth. Let \(C_0\) be “the least convex" body in \({{\mathcal {C}}}\) of \({\widetilde{g}}\)-volume one. (For details we refer to [93], but it should become clear from Examples 2.8 below how to construct \(C_0\).) Since \(\dim {\mathfrak {a}} = \mathrm {rank\,}(G/K) \ge 2\), \(C_0\) is not just a segment, and hence not an ellipsoid, i.e., the Finsler metric \(F_0\) associated with \(C_0\) is not Riemannian. In fact, \(F_0\) is not smooth. Verovic shows that \(F_0\) is the unique minimizer of \({\widehat{h}}_{{\text {vol}}} (F)\) among all G-invariant continuous Finsler metrics on \({\widetilde{Q}}\). In particular, the constant \(c^{{\text {BH}}}\) defined by

$$\begin{aligned} {\widehat{h}}_{{\text {vol}}}^{{\text {BH}}}(F_0) \,=\, c^{{\text {BH}}} \, \widehat{h}_{{\text {vol}}}({\widetilde{g}}) \end{aligned}$$
(2.22)

is strictly less than 1. There is a simple formula computing this constant in terms of the Weyl data of \({\mathfrak {a}}\).

Fix \(\varepsilon >0\), choose a smooth body C from \({{\mathcal {C}}}\) such that

$$\begin{aligned} C_0 \,\subset \, C \,\subset \, (1+\varepsilon )\, C_0, \end{aligned}$$

and let F be the associated Finsler metric. Then

$$\begin{aligned} {\widehat{h}}_{{\text {vol}}}^{{\text {BH}}} (F) \,\le \, (1+\varepsilon ) \, \widehat{h}_{{\text {vol}}}^{{\text {BH}}} (F_0), \end{aligned}$$
(2.23)

cf. Sect. 2.2. Since \({\widetilde{Q}}\) is of non-compact type, G is semi-simple, see for instance [99, Proposition 6.38 (d)]. It thus follows from [40, Theorem 6.3 (2)] that F has negative flag curvature. Therefore, the extension of Manning’s equality to reversible Finsler metrics in [44, Theorem 6.1] implies that

$$\begin{aligned} {\widehat{h}}_{{\text {top}}}^{{\text {BH}}} (F) \,=\, {\widehat{h}}_{{\text {vol}}}^{{\text {BH}}} (F). \end{aligned}$$
(2.24)

The line (2.21) follows from (2.24), (2.23), and (2.22).

Define the constant \(c^{{\text {HT}}}\) by

$$\begin{aligned} {\widehat{h}}_{{\text {vol}}}^{{\text {HT}}}(F_0) \,=\, c^{{\text {HT}}} \; \widehat{h}_{{\text {vol}}}({\widetilde{g}}). \end{aligned}$$

By (2.1) and the Santaló inequality, \(c^{{\text {HT}}} < c^{{\text {BH}}}\). Repeating the above arguments we obtain (2.20). \(\square \)

Example 2.8

1. Let Q be a compact quotient of the symmetric space \((\mathbb {H}^2)^k\) of rank k. The maximal abelian subalgebra \(\mathfrak a\) is \(\mathbb {R}^k\), with Weyl chamber \({\mathfrak {a}}^+ = \mathbb {R}_{>0}^k\). The set of positive roots is given by the dual basis \(\varepsilon _1, \dots , \varepsilon _k\) of the standard basis \(e_1, \dots , e_k\) of \(\mathbb {R}^k\).

The standard Riemannian metric g on Q corresponds, up to to scaling, to the closed unit ball B in \({\mathfrak {a}} = \mathbb {R}^k\), and we take \(F_0\) to be the non-smooth Finsler metric corresponding to the cross-polytope \(C_0\) in \({\mathfrak {a}}\) with vertices \(\pm e_1, \dots , \pm e_k\). Now Proposition 2.2 in [93] shows that

$$\begin{aligned} {\widehat{h}}_{{\text {vol}}}^{{\text {BH}}} (F_0)= & {} \left( {\text {vol}}_{F_0}^{{\text {BH}}}(Q) \right) ^{1/(2k)} \displaystyle \max _{v \in C_0 \cap \overline{\mathfrak a^+}} \bigl ( \varepsilon _1(v) + \dots + \varepsilon _k(v) \bigr ) = \left( {\text {vol}}_{F_0}^{{\text {BH}}}(Q) \right) ^{1/(2k)} \\ {\widehat{h}}_{{\text {vol}}}^{{\text {BH}}} (g)= & {} \left( {\text {vol}}_g^{{\text {BH}}}(Q) \right) ^{1/(2k)} \displaystyle \max _{v \in B \cap \overline{\mathfrak a^+}} \bigl ( \varepsilon _1(v) + \dots + \varepsilon _k(v) \bigr ) = \left( {\text {vol}}_g^{{\text {BH}}}(Q) \right) ^{1/(2k)} \sqrt{k}. \end{aligned}$$

Let \(D_{x_0}(F_0)\) resp. \(D_{x_0}(g)\) be the unit ball of \(F_0\) resp. g in \(T_{x_0}{\widetilde{Q}} \sim {\mathfrak {p}}\). By G-invariance of \(F_0\) and g, and in view of the definition of the Busemann–Hausforff volume in Sect. 2.1,

$$\begin{aligned} \frac{{\text {vol}}_{F_0}^{{\text {BH}}}(Q)}{{\text {vol}}_g^{{\text {BH}}}(Q)} \,=\, \frac{{\text {vol}}_{g}(D_{x_0}(g))}{{\text {vol}}_g(D_{x_0}(F_0))}. \end{aligned}$$
(2.25)

Lemma 2.9

The quotient on the right of (2.25) is equal to \(\frac{(2k)!}{2^k k!}\).

Proof

We have \(D_{x_0}(F_0) = {\text {Ad}}(K) (C_0)\) and \(D_{x_0}(g) = {\text {Ad}}(K) (B)\). For \(k=1\), when \(G = \textrm{SL}(2;\mathbb {R})\) and \(K = \textrm{SO}(2;\mathbb {R})\), a computation in the orthogonal basis \({ \bigl ( {\begin{matrix} 1&{}0\\ 0&{}-1 \end{matrix}} \bigr )}\), \({ \bigl ( {\begin{matrix} 0&{}1\\ 1&{}0 \end{matrix}} \bigr )}\) of \({\mathfrak {p}}\) shows that the orbit \({\text {Ad}}(K) p\) of a point \(p \in {\mathfrak {p}}\) is the circle through p. For general \(k \ge 1\), the \({\text {Ad}}(K)\)-orbit of \(p = (p_1, \dots , p_k) \in {\mathfrak {p}}\) is the k-torus made of circles of radius \(|p_i|\). Since the restrictions of \(F_0\) and g to \({\mathfrak {p}}\) are \({\text {Ad}}(K)\)-invariant, it follows that the quotient on the right of (2.25) is equal to the quotient of the two integrals

$$\begin{aligned} \int _{B \cap \overline{{\mathfrak {a}}^+}} \left( x_1 x_2 \cdots x_k \right) dx_1 dx_2 \cdots dx_k, \qquad \int _{C_0 \cap \overline{{\mathfrak {a}}^+}} \left( x_1 x_2 \cdots x_k \right) dx_1 dx_2 \cdots dx_k. \end{aligned}$$

The first integral equals \(\frac{1}{2^k k!}\) and the second equals \(\frac{1}{(2k)!}\) as one finds using Fubini’s theorem and induction. \(\square \)

Together with the lemma we conclude that

$$\begin{aligned} c^{{\text {BH}}}_k :=\, c^{{\text {BH}}} \left( (\mathbb {H}^2)^k \right) \,=\, \frac{{\widehat{h}}_{{\text {vol}}}^{{\text {BH}}} (F_0)}{{\widehat{h}}_{{\text {vol}}}^{{\text {BH}}} (g)} \,=\, \left( \frac{(2k)!}{k!} \right) ^{1/(2k)} \frac{1}{\sqrt{2k}}. \end{aligned}$$

We next compute the Holmes–Thompson volumes \({\text {vol}}_{F_0}^{{\text {HT}}}(Q)\) and \({\text {vol}}_{g}^{{\text {HT}}}(Q)\). Denote by \(C_0^*\) and \(B^*\) the polar sets of \(C_0\) and B in \({\mathfrak {p}}^*\), respectively, and by \(g^*\) the dual Riemannian metric on \({\mathfrak {p}}^*\). By G-invariance of \(F_0\) and g, and in view of the definition of the Holmes–Thompson volume in Sect. 2.1,

$$\begin{aligned} \frac{{\text {vol}}_{F_0}^{{\text {HT}}}(Q)}{{\text {vol}}_g^{{\text {HT}}}(Q)} \,=\, \frac{{\text {vol}}_{g^*}(D_{x_0}^*(F_0))}{{\text {vol}}_{g^*}(D_{x_0}^*(g))}. \end{aligned}$$
(2.26)

Using that \(D_{x_0}^*(F_0) = {\text {Ad}}^* (K) (C_0^*)\) and \(D_{x_0}^*(g) = {\text {Ad}}^* (K) (B^*)\), that the polar set \(C_0^*\) of the cross-polytope \(C_0\) is the unit cube, and the computations

$$\begin{aligned} \int _{B^* \cap \mathbb {R}_{\ge 0}^k} \left( x_1 \cdots x_k \right) \, dx_1 \cdots dx_k \,=\, \frac{1}{2^k\, k!}, \qquad \int _{C_0^* \cap \mathbb {R}_{\ge 0}^k} \left( x_1 \cdots x_k \right) dx_1 \cdots dx_k \,=\, \frac{1}{2^k}, \end{aligned}$$

we find that the right quotient in (2.26) is k!. Therefore,

$$\begin{aligned} c_k^{{\text {HT}}} :=\, c^{{\text {HT}}} \left( (\mathbb {H}^2)^k \right) \,=\, \frac{{\widehat{h}}_{{\text {vol}}}^{{\text {HT}}} (F_0)}{{\widehat{h}}_{{\text {vol}}}^{{\text {HT}}} (g)} \,=\, \left( k! \right) ^{1/(2k)} \frac{1}{\sqrt{k}}. \end{aligned}$$

It is shown in [75] that

$$\begin{aligned} h_{{\text {vol}}}^{{\text {sym}}} (Q) \,=\, h_{{\text {vol}}}(Q). \end{aligned}$$

Together with Theorem 1.2 and Proposition 1.12 we obtain

$$\begin{aligned} 2 c_{2k} \,h_{{\text {vol}}} (Q)\le & {} {\widehat{h}}_{{\text {top}}}^{{\text {HT}}} (Q,{{\mathcal {F}}}_{{\text {rev}}}) \;\le \; c_k^{{\text {HT}}} \,h_{{\text {vol}}} (Q), \\ 2 c_{2k} \,h_{{\text {vol}}} (Q)\le & {} {\widehat{h}}_{{\text {top}}}^{{\text {BH}}} (Q,{{\mathcal {F}}}_{{\text {rev}}}) \;\le \; c_k^{{\text {BH}}} \,h_{{\text {vol}}} (Q). \end{aligned}$$

The sequence \(c_k^{{\text {BH}}}\), \(k \ge 2\), is monotone decreasing to \(\sqrt{\frac{2}{e}} \approx 0.858\), starting with

$$\begin{aligned} c_2^{{\text {BH}}} \approx 0.931 \quad \text{ and } \quad c_3^{{\text {BH}}} \approx 0.907. \end{aligned}$$

The sequence \(c_k^{{\text {HT}}}\) is monotone decreasing to \(\sqrt{\frac{1}{e}} \approx 0.616\), starting with

$$\begin{aligned} c_2^{{\text {HT}}} \approx 0.841 \quad \text{ and } \quad c_3^{{\text {HT}}} \approx 0.778. \end{aligned}$$

In contrast, the sequence \(2 c_{2k} = \frac{2}{\bigl ( (2k)! \, \omega _{2k} \bigr )^{1/(2k)}}\) is monotone decreasing like \(\sqrt{\frac{e}{2\pi }} \frac{1}{\sqrt{2k}}\), starting with

$$\begin{aligned} 2 c_4 \approx 0.606 \quad \text{ and } \quad 2 c_6 \approx 0.508. \end{aligned}$$

The constant \(h_{{\text {vol}}}(Q)\) can be computed as follows. On \((\mathbb {H}^2)^k\) the minimum of the volume entropies among symmetric metrics is attained exactly by multiples of \(g^k:= g \times \dots \times g\), where g is the metric on \(\mathbb {H}\) of constant curvature \(-1\), see [35, §2]. Since \(h_{{\text {vol}}}(g) =1\), we have \(h_{{\text {vol}}}(g^k) = \sqrt{\sum (h_{{\text {vol}}}(g))^2} = \sqrt{k}\). Hence

$$\begin{aligned} h_{{\text {vol}}}(Q) \,=\, h_{{\text {vol}}}^{{\text {sym}}} (Q) \,=\, \left( {\text {vol}}_{g^k}(Q)\right) ^{1/2k} \sqrt{k}. \end{aligned}$$

For instance, if Q is the product of orientable surfaces of genus \(k_j\), then

$$\begin{aligned} {\text {vol}}_{g^k}(Q) \,=\, \prod _{j=1}^k \textrm{area}_g(Q_{k_j}) \,=\, 2^k \prod _{j=1}^k \sqrt{\pi (k_j-1)}. \end{aligned}$$

2. Take the 5-dimensional symmetric space \(\textrm{SL}(3;\mathbb {R}) / \textrm{SO}(3;\mathbb {R})\) of rank 2. The “least convex” body \(C_0\) from \({{\mathcal {C}}}\) is a regular hexagon. We scale this hexagon such that it is the hexagon \(H_{{\textrm{in}}}\) inscribed the unit disc B of \(\mathbb {R}^2 = {\mathfrak {a}}\). Verovic computed in [93, p. 1644] that for the Finsler metric corresponding to \(\sqrt{\frac{2}{3}} \, H_{{\textrm{in}}}\), the volume growth is 2. Hence the volume growth of the Finsler metric corresponding to \(H_{{\textrm{in}}}\) is \(\sqrt{\frac{3}{2}} \,2\). Further, the volume growth of the Riemannian metric corresponding to B is \(2 \sqrt{2}\).

To compute the volumes, since \(\textrm{SO}(3;\mathbb {R})\) is 3-dimensional we now have to take \(r^3 dx dy\) as density on \({\mathfrak {a}}\). The integral of \(r^3\) over B and over \(H_{{\textrm{in}}}\) are, respectively, \(\frac{2\pi }{5}\) and

$$\begin{aligned} I_{{\textrm{in}}} \,=\, \frac{3 \sqrt{3}}{640} \left( 27 \ln 3 + 68 \right) . \end{aligned}$$

With this we find along the lines of the previous examples that

$$\begin{aligned} c^{{\text {BH}}} (\textrm{SL}(3;\mathbb {R}) / \textrm{SO}(3;\mathbb {R})) \,=\, \left( \frac{2\pi }{5 I_{{\textrm{in}}}} \right) ^{1/5} \frac{\sqrt{3}}{2} \,\approx \, 0.95. \end{aligned}$$

The polar set of \(H_{{\textrm{in}}}\) is a regular hexagon \(H_{{\textrm{out}}}\) circumscribed the unit co-disc. After identifying \({\mathfrak {a}}\) with \({\mathfrak {a}}^*\) by the inner product, we have that \(H_{{\textrm{out}}}\) is obtained from \(H_{{\textrm{in}}}\) by dilation by \(\frac{2}{\sqrt{3}}\) and rotation by \(\frac{\pi }{6}\). Hence the integral of \(r^3\) over \(H_{{\textrm{out}}}\) is

$$\begin{aligned} I_{{\textrm{out}}} \,=\, \left( \frac{2}{\sqrt{3}} \right) ^{5} I_{{\textrm{in}}}. \end{aligned}$$

Therefore,

$$\begin{aligned} c^{{\text {HT}}} (\textrm{SL}(3;\mathbb {R}) / \textrm{SO}(3;\mathbb {R})) \,=\, \left( \frac{5 I_{{\textrm{out}}} }{2\pi } \right) ^{1/5} \frac{\sqrt{3}}{2} \,=\, \frac{\sqrt{3}}{2 \, c^{{\text {BH}}}} \,\approx \, 0.912. \end{aligned}$$

These two constants should be compared with the constant \(2 c_5 \approx 0.551\) for the lower bound in Corollary 2.6.

Fig. 4
figure 4

The co-disks \(\frac{1}{\sigma _+} D^*_q (F) \subset D^*_q (H) \subset \sigma _- \, D^*_q (F)\) in \(T_q^*Q\)

Question 2.10

Recall that the non-smooth Finsler metric \(F_0\) is the unique minimizer of \({\widehat{h}}_{{\text {vol}}}^{{\text {BH}}}(F)\) among G-invariant continuous Finsler metrics on Q. This Finsler metric has a high degree of symmetry: Its restriction to \({\mathfrak {a}}\) is invariant under the Weyl group, and it is G-invariant. Since in Theorems 1.3 and 1.8 the Riemannian minimizers are the locally symmetric metrics, one may expect that \(F_0\) minimizes \({\widehat{h}}_{{\text {vol}}}^{{\text {BH}}}(F)\) and \({\widehat{h}}_{{\text {vol}}}^{{\text {HT}}}(F)\) among all continuous Finsler metrics on Q. Would this imply that there are no smooth minimizers?

3 A lower entropy bound for Reeb flows on spherizations

Recall from Theorem 1.1 that there cannot be a uniform lower bound for the normalized topological entropy of Reeb flows. In this section we show that for many base manifolds Q, one nevertheless has a control on the entropy collapse of Reeb flows on the spherization \(S^*Q\) in terms of the geometry of their defining star fields: Entropy collapse can only happen if some fibers are far from convex. The proof relies on Floer homology.

We consider a closed manifold Q and two Reeb flows on \(S^*Q\), one arbitrary and one Finsler. As in the previous section and as in Appendix B.1 we work in \(T^*Q\). We then have two Hamiltonian functions \(H,F :T^*Q \rightarrow \mathbb {R}\) that are fiberwise positively homogeneous of degree one and smooth and positive away from the zero section. Again we denote by \(\phi _H^t\) the flow of H on \(S^*(H) = H^{-1}(1)\), and similarly for F. Let \(\sigma _-\) and \(\sigma _+\) be the smallest positive numbers such that

$$\begin{aligned} \frac{1}{\sigma _-}\, F \,\le \, H \,\le \, \sigma _+\, F \quad \text{ on } T^*Q. \end{aligned}$$

For the co-disk bundles we then have

$$\begin{aligned} \frac{1}{\sigma _+}\, D^*(F) \,\subset \, D^*(H) \,\subset \, \sigma _- \, D^*(F), \end{aligned}$$

see Fig. 4.

The number

$$\begin{aligned} \sigma (H;F) :=\, \sigma _- \, \sigma _+. \end{aligned}$$

does not change under rescalings of H or F. We have \(\sigma (H;F) \ge 1\) with equality if and only if \(H=cF\) for some positive number c. Moreover, \(\sigma (H;F) \le \sqrt{n}\) if H is a reversible Finsler Hamiltonian and F is chosen to be the Riemannian Hamiltonian associated with the outer Loewner ellipsoids of \(D^*(H)\), see (2.8).

Proposition 3.1

Let F be a (possibly irreversible) \(C^\infty \)-regular Finsler metric on the closed manifold Q. Then for every \(C^\infty \)-smooth Reeb flow \(\phi _H^t\) on \(S^*Q\) we have

$$\begin{aligned} {\widehat{h}}_{{\text {top}}}^{{\text {HT}}}(\phi _H) \,\ge \, \frac{1}{\sigma (H;F)} \, {\widehat{h}}_{{\text {vol}}}(F). \end{aligned}$$

Proof

After scaling F we can assume that \(\sigma _- \!=\!1\). We abbreviate \(\sigma (H;F) \!=\! \sigma _+ \!=: \!\sigma \).

Lemma 3.2

\(h_{{\text {top}}}(\phi _H) \ge h_{{\text {vol}}}(F)\).

Proof

The lemma can be extracted from [68]. We briefly review the proof. Instead of working with \(\phi _H\) and \(\phi _F\), we work with the Hamiltonian flows \(\Phi _H\) and \(\Phi _F\) on \(T^*Q\) of the functions \(H^2\) and \(F^2\). Then \(\Phi _H = \phi _H\) on \(S^*(H)\). Using the variational principle for topological entropy and the homogeneity of \(H^2\) one finds

$$\begin{aligned} h_{{\text {top}}} (\phi _H) \,=\, h_{{\text {top}}} (\Phi _H |_{S^*(H)}) \,=\, h_{{\text {top}}} (\Phi _H |_{D^*(H)}). \end{aligned}$$

Fixing a point \(q \in Q\) we can further estimate, using Yomdin’s theorem from [96] and the \(C^\infty \)-smoothness of \(\Phi _H\),

$$\begin{aligned} h_{{\text {top}}} (\Phi _H |_{D^*(H)}) \,\ge \, \lim _{n \rightarrow \infty } \frac{1}{n} \log \mu _{g^*} \bigl ( \Phi _H^n (D^*_q(H)) \bigr ). \end{aligned}$$

Here \(\mu _{g^*} (S)\) denotes the Riemannian volume of the submanifold \(S \subset T^*Q\) with respect to the restriction to S of the Riemannian metric on \(T^*Q\) induced by a Riemannian metric g on Q. In Theorem 4.6 and Section 5.1 of [68] it is shown by Lagrangian Floer homology that for every \(\varepsilon >0\) there exists \(N(\varepsilon )\) such that

$$\begin{aligned} \mu _{g^*} \bigl ( \Phi _H^n (D^*_q(H)) \bigr ) \,\ge \, {\text {vol}}_g(Q) \, e^{(\gamma (F)-\varepsilon ) n} \quad \text{ for } n \ge N(\varepsilon ) \end{aligned}$$

where \(\gamma (F)\) is the exponential growth rate of the number of elements in the fundamental group of Q that can be represented by a loop of F-length \(\le R\). It is easy to see that \(\gamma (F) = h_{{\text {vol}}} (F)\). (The proof in [60, Prop. 9.6.6] given for a Riemannian F applies without changes to a general Finsler metric.) The lemma follows. \(\square \)

In view of the inclusion \(\frac{1}{\sigma }D^*(F) \subset D^*(H)\) we infer from Lemma 3.2 that

$$\begin{aligned} {\widehat{h}}_{{\text {top}}}^{{\text {HT}}} (\phi _H)= & {} \left( {\text {vol}}_H^{{\text {HT}}}(Q) \right) ^{1/n} \, h_{{\text {top}}} (\phi _H) \\\ge & {} \frac{1}{\sigma } \left( {\text {vol}}_F^{{\text {HT}}}(Q) \right) ^{1/n} \, h_{{\text {vol}}} (F) \\= & {} \frac{1}{\sigma } \, {\widehat{h}}_{{\text {vol}}}(F) \end{aligned}$$

as claimed. \(\square \)

We now define the module of starshapedness of H by

$$\begin{aligned} \sigma (H) :=\, \inf \left\{ \sigma (H;F) \mid F \text{ is } \text{ a } \text{ Finsler } \text{ metric } \right\} . \end{aligned}$$

While an individual \(\sigma (H;F)\) can be large even for a Riemannian Hamiltonian H, the number \(\sigma (H)\) is a measure for the maximal starshapedness, or non-convexity, of the fibers of \(D^*(H)\). For instance, \(\sigma (H) = 1\) if and only if H is Finsler. From Proposition 3.1 and Corollary 2.3 we obtain the following result.

Corollary 3.3

Let Q be a closed manifold. For every \(C^\infty \)-smooth Reeb flow \(\phi _H^t\) on \(S^*Q\) we have

$$\begin{aligned} {\widehat{h}}_{{\text {top}}}^{{\text {HT}}}(\phi _H) \,\ge \, \frac{c_n}{\sigma (H)} \, {\widehat{h}}_{{\text {vol}}}(Q). \end{aligned}$$

Remark 3.4

  1. (1)

    In the special case that \(H=F\) is a \(C^\infty \)-regular reversible Finsler Hamiltonian, Proposition 3.1 applied to Riemannian metrics and the Loewner bound (2.8) yield the uniform lower bound

    $$\begin{aligned} {\widehat{h}}_{{\text {top}}}^{{\text {HT}}}(F) \,\ge \, \tfrac{1}{\sqrt{n}} \, \widehat{h}_{{\text {vol}}} (Q). \end{aligned}$$

    Even in this special case, this lower bound for \(\widehat{h}_{{\text {top}}}^{{\text {HT}}}(F)\) coming from Floer homology and from the Loewner bound is only slightly weaker than the lower bound

    $$\begin{aligned} {\widehat{h}}_{{\text {top}}}^{{\text {HT}}}(F) \,\ge \, 2 c_n \, {\widehat{h}}_{{\text {vol}}} (Q) \end{aligned}$$

    from Corollary 2.6 that comes from Manning’s inequality and the Loewner bound. Indeed, recalling that \(c_n = \frac{1}{ (n! \, \omega _n)^{1/n} }\), the function \(f(n) = 2 c_n \sqrt{n} :\mathbb {N}\rightarrow [1, \infty )\) is strictly monotone increasing, with

    $$\begin{aligned} f(2) = \tfrac{2}{\sqrt{\pi }} \approx 1.13 \quad \text{ and } \quad \lim _{n \rightarrow + \infty } f(n) = \sqrt{\tfrac{2e}{\pi }} \approx 1.315. \end{aligned}$$
  2. (2)

    In the case that H is a Finsler Hamiltonian and F is a Riemannian Hamiltonian, we have obtained the inequality in Lemma 3.2 in Sect. 2.5 by estimating

    $$\begin{aligned} h_{{\text {top}}}(\phi _H) \ge h_{{\text {vol}}}(\phi _H) \ge h_{{\text {vol}}}(F). \end{aligned}$$

    The first inequality, which is Manning’s inequality, also holds for \(C^\infty \)-smooth Reeb flows, see Theorem A.8. The second inequality holds in the Finsler case in view of the inclusion of balls (2.6), which follows from the triangle inequality. But in the Reeb case there is no triangle inequality. Floer homology (or, more precisely: properties of Floer continuation maps that stem from the Floer–Gromov compactness theorem for J-holomorphic strips) makes up for this.

  3. (3)

    Proposition 3.1 and Corollary 3.3 are interesting only if \({\widehat{h}}_{{\text {vol}}}(F)\) and \(h_{{\text {vol}}}(Q)\) are positive, which is possible only if the fundamental group of Q has exponential growth. The results in [68] imply meaningful variations of Proposition 3.1 and Corollary 3.3 for many other manifolds. For instance, assume that Q is a simply connected manifold such that the exponential growth rate \(\gamma (\Omega Q)\) of the dimension of the \(\mathbb {Z}_2\)-homology of degree \(\le k\) of the based loop space \(\Omega Q\) is positive. Then

    $$\begin{aligned} {\widehat{h}}_{{\text {top}}}^{{\text {HT}}} (\phi _H) \,\ge \, \frac{1}{\sigma (H;F)} \; C(F) \, \gamma (\Omega Q) \end{aligned}$$

    with a positive constant C(F) that does not change under rescalings of F.

  4. (4)

    We refer to [38] for a thorough study of continuity properties of topological entropy implied by Floer homological techniques.

4 Entropy collapse for Reeb flows in dimension 3

In this section we prove Theorem 1.1 in dimension 3.

4.1 Recollections on open books

In this paragraph we collect results on open books needed in our proof. For more information and details we refer to [46] and [52, §4.4].

Let M be a closed connected orientable 3-manifold. An open book for M is a triple \((\Sigma ,\psi ,\Psi )\), where \(\Sigma \) is a compact oriented surface with non-empty boundary \(\partial \Sigma \) and \(\psi \) is a diffeomorphism of \(\Sigma \) that is the identity near the boundary such that there is a diffeomorphism \(\Psi \) from

$$\begin{aligned} M(\psi ):= \Sigma (\psi ) \cup _{\textrm{id}} \left( {\overline{\mathbb {D}}} \times \partial \Sigma \right) \end{aligned}$$

to M. Here \(\Sigma (\psi )\) denotes the mapping torus

$$\begin{aligned} \Sigma (\psi ) \,=\, \left( [0,2\pi ] \times \Sigma \right) / \sim \end{aligned}$$

where \((2\pi ,p) \sim (0,\psi (p))\) for each \(p \in \Sigma \), and \({\overline{\mathbb {D}}}\) is the closed unit disk. Viewing \(S^1\) as the interval \([0,2\pi ]\) with endpoints identified, we write \(\partial (\Sigma (\psi ))\) as \(S^1 \times \partial \Sigma \). The manifold M is thus presented as the union of the mapping torus \(\Sigma (\psi )\) and finitely many full tori, one for each boundary component of \(\Sigma \), glued along their boundaries by the identity map

$$\begin{aligned} \partial (\Sigma (\psi )) \,=\, S^1 \times \partial \Sigma \,{\mathop {\longrightarrow }\limits ^{\textrm{id}\,}}\, \partial ( {\overline{\mathbb {D}}} \times \partial \Sigma ). \end{aligned}$$

We remark that the diffeomorphism

$$\begin{aligned} \Psi :M(\psi ) \rightarrow M \end{aligned}$$

is part of the definition of the open book. If \(\psi '\) is another diffeomorphism of \(\Sigma \) that is the identity near the boundary and is isotopic to \(\psi \) via an isotopy that fixes each point of \(\partial \Sigma \), then \(M(\psi ')\) is diffeomorphic to \(M(\psi )\). We also remark that what we call an open book is usually called an abstract open book decomposition in the literature.

For each \(\theta \in S^1\) denote by \(\Sigma _{\theta }^\circ \) the image under the diffeomorphism \(\Psi \) of the union of \(\{\theta \} \times \Sigma \) with the union of half-open annuli

$$\begin{aligned} A_{\theta } \,=\, \bigl \{ (\theta ,r) \in {\overline{\mathbb {D}}} \setminus \{0\} \bigr \} \times \partial \Sigma . \end{aligned}$$

The closure \(\Sigma _{\theta }\) of \(\Sigma _{\theta }^\circ \), called a page, is diffeomorphic to \(\Sigma \), and the common boundary of the pages \(\Sigma _{\theta }\), called the binding of the open book, is the image under \(\Psi \) of \( \{0\} \times \partial \Sigma \subset \mathbb {D}\times \partial \Sigma \). The orientation of \(\Sigma \) induces orientations on the pages and the binding.

There are several different beautiful constructions proving the existence of an open book for every 3-manifold M as above. The first of these constructions was given by J. W. Alexander [5] as early as 1920, who used his findings that every such M is a branched covering of the 3-sphere branching along a link and that every link in \(\mathbb {R}^3\) can be obtained as the closure of a braid, see also [87, p. 340]. Alexander’s construction in fact provides an open book such that \(\Sigma \) has just one boundary component.

Contact structures. Let M be a closed connected oriented 3-manifold and \((\Sigma , \psi ,\Psi )\) be an open book for M.

Definition 4.1

A contact form \(\alpha \) on M is said to be adapted to the open book \((\Sigma ,\psi ,\Psi )\) if

  • \(\alpha \) is positive on the binding,

  • \(d\alpha \) is a positive area form on the interior of every page.

It is not hard to see that a contact form \(\alpha \) is adapted to an open book if and only if

  • the Reeb vector field \(R_\alpha \) is positively transverse to the interior of the pages,

  • the Reeb vector field is tangent to the binding and induces the positive orientation on the binding.

Definition 4.2

A contact 3-manifold \((M,\xi )\) is said to be supported by an open book \((\Sigma ,\psi ,\Psi )\) if there exists a contact form \(\alpha \) on \((M,\xi )\) adapted to this open book.

Remark 4.3

If a contact 3-manifold \((M,\xi )\) is supported by an open book \((\Sigma ,\psi ,\Psi )\) and if \(\psi '\) is another diffeomorphism of \(\Sigma \) that is the identity near \(\partial \Sigma \) and that is isotopic to \(\psi \) via an isotopy that fixes \(\partial \Sigma \) pointwise, then \((M,\xi )\) is also supported by an open book \((\Sigma ,\psi ',\Psi ')\). This follows easily from the fact that for such a \(\psi '\) there exists a diffeomorphism from \(M(\psi )\) to \(M(\psi ')\) that takes pages to pages.

The following result of Giroux shows the central role played by open books in 3-dimensional contact topology.

Theorem 4.4

(Giroux) Given a closed connected oriented contact 3-manifold \((M,\xi )\), there exists an open book for M supporting \((M,\xi )\). Moreover, the open book can be chosen to have connected binding. Two contact structures supported by the same open book are diffeomorphic.

For the proof of the first and the third assertion we refer to [53, Theorem 3 and Proposition 2] and to [46, Theorem 4.6 and Proposition 3.18]. That the binding can be assumed to be connected is shown in [46, Corollary 4.25] and in [34].

4.2 Proof of entropy collapse in dimension 3

We now proceed with the proof of the main result of this section.

Theorem 4.5

Let \((M,\xi )\) be a closed co-orientable contact manifold of dimension 3. Then for every \(\varepsilon > 0\) there exists a contact form \(\alpha \) on \((M,\xi )\) such that \({\text {vol}}_{\alpha } (M) =1\) and \(h_{{\text {top}}}(\alpha ) \le \varepsilon \).

While our proof works verbatim when \(\partial \Sigma \) is not connected, the geometry in our argument is easier to visualize for connected \(\partial \Sigma \), so we assume this property.

The structure of the proof is as follows.

  • Given \((M,\xi )\) as in Theorem 4.5 we use the first statement in Theorem 4.4 to obtain an open book \((\Sigma ,\psi ,\Psi )\) for M that supports \((M,\xi )\).

  • We then apply a classical recipe due to Thurston–Winkelnkemper to construct for each \(\varepsilon >0\) a contact form \({\widetilde{\alpha }}_\varepsilon \) adapted to \((\Sigma ,\psi ,\Psi )\) with \({\text {vol}}_{\widetilde{\alpha }_\varepsilon }(M) = 1\) and \(h_{{\text {top}}}({\widetilde{\alpha }}_\varepsilon ) \le \varepsilon \).

  • By the second statement of Theorem 4.4, \(\ker \Psi _* {\widetilde{\alpha }}_\varepsilon \) is diffeomorphic to \(\xi \) by a diffeomorphism \(\rho _\varepsilon \). Hence \((\rho _\varepsilon \circ \Psi )_* ({\widetilde{\alpha }}_\varepsilon )\) is a contact form on \((M,\xi )\) with the properties asserted in Theorem 4.5.

For the construction of \({\widetilde{\alpha }}_{\varepsilon }\), we first construct on the mapping torus \(\Sigma (\psi )\) for all small \(s>0\) contact forms \(\alpha _s\) with \({\text {vol}}_{\alpha _s}(\Sigma (\psi )) = O(s)\) and \(h_{{\text {top}}}(\alpha _s) = O(1)\). Crucially, near the boundary of \(\Sigma (\psi )\) these contact forms are such that they extend to contact forms (also denoted \(\alpha _s\)) on the full torus \(\overline{\mathbb {D}}\times \partial \Sigma \) in such a way that the Reeb flows are linear on each torus \(S^1(r) \times \partial \Sigma \). Therefore, even though the Reeb vector fields “explode" in the interior of the full torus as \(s \rightarrow 0\) (see Fig. 7), the topological entropy on the full torus vanishes for all s. Since also \({\text {vol}}_{\alpha _s}({\overline{\mathbb {D}}} \times \partial \Sigma ) = O(s)\), we find that \({\text {vol}}_{\alpha _s}(M) = O(s)\) and \(h_{{\text {top}}}(\alpha _s) = O(1)\) for all small \(s >0\). The form \({\widetilde{\alpha }}_\varepsilon \) is now obtained by taking s small and rescaling \(\alpha _s\).

Proof of Theorem 4.5

  Step 1: A family of contact forms \(\alpha _s\) on \(\Sigma (\psi )\). By Theorem 4.4 there exists an open book \((\Sigma ,\psi ,\Psi )\) for M that supports \((M,\xi )\). We first choose a collar neighbourhood \(N \subset \Sigma \) of \(\partial \Sigma \) on which \(\psi \) is the identity. Thus N is diffeomorphic to \([1,1+\delta ] \times \partial \Sigma \), and we have polar coordinates (rx) for N, where x is the angular coordinate for \(\partial \Sigma \), such that the boundary of \(\partial \Sigma \) corresponds to \(r=1\).

Choose an area form \(\omega \) on \(\Sigma \) such that \(\omega = dx \wedge dr\) on N. By Remark 4.3 and by Moser’s isotopy theorem we can assume that the diffeomorphism \(\psi \) is a symplectomorphism of \((\Sigma , \omega )\). Since \(H^2(\Sigma ,\partial \Sigma ;\mathbb {R})\) vanishes, there exists a primitive \(\lambda \) of \(\omega \) that equals \((2-r) dx\) on N.

Fig. 5
figure 5

The neighbourhood N of \(\partial \Sigma \)

We now construct for each sufficiently small \(s>0\) a contact form \(\alpha _s\) on the mapping torus \(\Sigma (\psi )\). For this, let \(\chi :[0,2\pi ] \rightarrow [0,1]\) be a smooth monotone function such that \(\chi (0)=0\), \(\chi (2\pi )=1\), and \(\chi '\) has support in \((0,2\pi )\). On \([0,2\pi ] \times \Sigma \) define the 1-form

$$\begin{aligned} \alpha _s :=\, d\theta + s \bigl ( (1-\chi (\theta )) \lambda + \chi (\theta ) \psi ^*\lambda \bigr ). \end{aligned}$$
(4.1)

By the properties of \(\chi \), each 1-form \(\alpha _s\) descends to a 1-form on \(\Sigma (\psi )\), that we still denote by \(\alpha _s\). Using that \(\psi \) is a symplectomorphism of \((\Sigma , \omega )\) we compute that

$$\begin{aligned} \alpha _s \wedge d \alpha _s \,=\, s d\theta \wedge \omega + O(s^2). \end{aligned}$$
(4.2)

Hence there exists \(s_0>0\) such that \(\alpha _s\) is a contact form for all \(s \in (0,s_0]\).

We now compute the Reeb vector field \(R_{\alpha _s}\). With \(\lambda _{\theta }:= (1-\chi (\theta )) \lambda + \chi (\theta ) \psi ^*\lambda \) one checks that

$$\begin{aligned} R_{\alpha _s} \,=\, \frac{\partial _{\theta } + Y}{1 + s \lambda _{\theta } (Y)}, \end{aligned}$$
(4.3)

where Y is the vector field that is tangent to \(\{\theta \} \times \Sigma \) for all \(\theta \in S^1\) and satisfies

$$\begin{aligned} \iota _Y \omega \,=\, \chi '(\theta ) (\psi ^*\lambda - \lambda ). \end{aligned}$$

The formula (4.3) shows that \(R_{\alpha _s}\) is positively transverse to each surface \(\{\theta \} \times \Sigma \).

The next lemma gives an upper bound for \(h_{{\text {top}}}(\phi _{\alpha _s})\) when s is sufficiently small. Notice that it makes sense to talk about \(h_{{\text {top}}}(\phi _{\alpha _s})\), since \(\Sigma (\psi )\) is compact and \(R_{\alpha _s}\) is tangent to \(\partial \Sigma (\psi )\).

Lemma 4.6

There exists \(s_1 \in (0,s_0)\) and a constant \(E>0\) such that for every \(s \in (0,s_1]\),

$$\begin{aligned} h_{{\text {top}}} (\phi _{\alpha _s}) \le E. \end{aligned}$$
(4.4)

Proof

Choose \(s_1 >0\) such that \(\frac{1}{2} \le \frac{1}{1+ s \lambda _\theta (Y)} \le 2\) on \(\Sigma (\psi )\) for every \(s \in (0,s_1]\). For each such s define the function \(f_s = \frac{1}{1+ s \lambda _\theta (Y)}\) on \(\Sigma (\psi )\). Then for every \(s \in (0,s_1]\) we have \(R_{\alpha _s} = f_s \, (\partial _{\theta } + Y)\) with \(\frac{1}{2} \le f_s \le 2\). By Ohno’s result from [79], for a non-vanishing vector field X and a positive function f on a compact manifold,

$$\begin{aligned} h_{{\text {top}}} (\phi _{fX}) \,\le \, \max f \cdot h_{{\text {top}}} (\phi _{X}). \end{aligned}$$

It follows that \(h_{{\text {top}}} (\phi _{\alpha _s}) \le 2 h_{{\text {top}}} (\phi _{\partial _{t} + Y }) =:E\). \(\square \)

Step 2: A family of contact forms \(\sigma _s\) on \(\overline{\mathbb {D}} \times \partial \Sigma \). We have constructed a family of contact forms \(\alpha _s\) on the mapping torus \(\Sigma (\psi )\). We now wish to extend these forms to contact forms on \(M(\psi )\). For this let V be the collar neighbourhood of \(\partial \Sigma (\psi )\) defined by

$$\begin{aligned} V:= [0,2\pi ]\times N / \sim \end{aligned}$$

where \( (2\pi ,p) \sim (0,p)\) for each \(p \in \Sigma \). On V the contact form \(\alpha _s\) reads

$$\begin{aligned} \alpha _s \,=\, d \theta + s (2-r) dx \end{aligned}$$
(4.5)

where (rx) are the coordinates on N introduced above and \(\theta \in S^1\).

We proceed to construct for each \(s \in (0,s_1)\) a contact form \(\sigma _s\) on \({\overline{\mathbb {D}}} \times \partial \Sigma \), where \(\overline{\mathbb {D}}\) is again the closed unit disk in \(\mathbb {R}^2\). Consider polar coordinates \((\varvec{\uptheta }, {{\textbf {r}}}) \in S^1 \times (0,1]\) on \(\overline{\mathbb {D}} \setminus \{0\}\) and the coordinate \({{\textbf {x}}}\) on \(\partial \Sigma \). We can then consider coordinates \((\varvec{\uptheta }, {{\textbf {r}}}, {{\textbf {x}}})\) on \(\overline{\mathbb {D}} {\setminus } \{0\} \times \partial \Sigma \). We pick a smooth function \(f :(0,1] \rightarrow \mathbb {R}\) such that

  • \(f'<0\),

  • \(f({{\textbf {r}}}) = 2-{{\textbf {r}}}\) on a neighbourhood of 1,

  • \(f({{\textbf {r}}}) = 2 -{{\textbf {r}}}^4\) on a neighbourhood of 0,

and we pick another smooth function \(g :(0,1]\rightarrow \mathbb {R}\) satisfying

  • \(g'>0\) on (0, 1),

  • \(g(1) =1\) and all derivatives of g vanish at 1,

  • \(g({{\textbf {r}}}) = \frac{{{\textbf {r}}}^2}{2}\) on a neighbourhood of 0 see Fig. 6.

Fig. 6
figure 6

The functions f and g

Define the 1-form

$$\begin{aligned} \sigma _s (\varvec{\uptheta }, {{\textbf {r}}}, {{\textbf {x}}}) \,=\, g({{\textbf {r}}}) d\varvec{\uptheta }+ s f({{\textbf {r}}}) d{{\textbf {x}}}\end{aligned}$$
(4.6)

on \(\overline{\mathbb {D}} \setminus \{0\} \times \partial \Sigma \). Then

$$\begin{aligned} \sigma _s \wedge d \sigma _s \,=\, s h({{\textbf {r}}}) d{{\textbf {r}}}\wedge d \varvec{\uptheta }\wedge d{{\textbf {x}}}\end{aligned}$$
(4.7)

where \(h({{\textbf {r}}}) = (fg'-f'g)({{\textbf {r}}})\). It follows that \(\sigma _s\) is a contact form on \(\overline{\mathbb {D}} {\setminus } \{0\} \times \partial \Sigma \). For \({{\textbf {r}}}\) near 0 we have \(h({{\textbf {r}}}) = {{\textbf {r}}}(2+{{\textbf {r}}}^4)\), whence \(\sigma _s\) extends to a smooth contact form on \(\overline{\mathbb {D}} \times \partial \Sigma \), that we also denote by \(\sigma _s\). The Reeb vector field of \(\sigma _s\) is given by

$$\begin{aligned} R_{\sigma _s} (\varvec{\uptheta }, {{\textbf {r}}}, {{\textbf {x}}}) \,=\, \frac{1}{h({{\textbf {r}}})} \left( - f'({{\textbf {r}}}) \partial _{\varvec{\uptheta }} + \frac{1}{s} g'({{\textbf {r}}}) \partial _{{{\textbf {x}}}} \right) . \end{aligned}$$
(4.8)

It follows that \(R_{\sigma _s}\) is tangent to the tori \(\mathbb {T}_{{{\textbf {r}}}}:= \{ {{\textbf {r}}}= \text {const} \}\) and that for each \({{\textbf {r}}}\in (0,1]\) the flow of \(R_{\sigma _s}\) is linear:

$$\begin{aligned} \phi _{\sigma _s}^t (\varvec{\uptheta }, {{\textbf {r}}}, {{\textbf {x}}}) \,=\, \left( \varvec{\uptheta }- \frac{f'({{\textbf {r}}})}{h({{\textbf {r}}})} t, \, {{\textbf {r}}}, \, {{\textbf {x}}}+ \frac{g'({{\textbf {r}}})}{s h({{\textbf {r}}})} t \right) . \end{aligned}$$
(4.9)

In particular, using our choices of f and g we see that \(R_{\sigma _s} = \partial _{\varvec{\uptheta }}\) on the boundary torus \(\mathbb {T}_1\), and that \(R_{\sigma _s} = \frac{1}{2 s} \partial _{{{\textbf {x}}}}\) is tangent along the core circle \(C= \{ {{\textbf {r}}}=0 \}\) of the full torus, and gives the positive orientation to \(\partial \Sigma \). Furthermore, (4.8) shows that \(R_{\sigma _s}\) is positively transverse to the half-open annuli see Fig. 7

$$\begin{aligned} A_{\varvec{\uptheta }}:= \{\varvec{\uptheta }\} \times (0,1] \times \partial \Sigma \,\subset \, \overline{\mathbb {D}} \setminus \{0\} \times \partial \Sigma . \end{aligned}$$

The Reeb flow \(\phi _{\sigma _s}^t\) on the full torus \({\overline{\mathbb {D}}} \times \partial \Sigma \) is integrable. More precisely, the core circle C of \({\overline{\mathbb {D}}} \times \partial \Sigma \) is the trace of a periodic orbit of \(\phi _{\sigma _s}^t\), and \(({\overline{\mathbb {D}}} \times \partial \Sigma ) {\setminus } C\) is foliated by the flow-invariant tori \(\mathbb {T}_{{\textbf {r}}}\), on which \(\phi _{\sigma _s}^t\) is the linear flow (4.9). The topological entropy of these linear flows of course vanishes. By the variational principle for topological entropy we therefore find that

$$\begin{aligned} h_{{\text {top}}}(\phi _{\sigma _s}) \,=\, \sup _{0 \le r \le 1} h_{{\text {top}}}(\phi _{\sigma _s} |_{\mathbb {T}_{{{\textbf {r}}}}}) \,=\, 0. \end{aligned}$$
(4.10)
Fig. 7
figure 7

Some vectors of \(R_{\sigma _s}\) and the annulus \(A_{\varvec{0}}\)

Step 3: A family of contact forms on M. We first observe that the coordinates \(\theta \) and \(\varvec{\uptheta }\), r and \({{\textbf {r}}}\), x and \({{\textbf {x}}}\) are glued via the identification map used to glue \(\Sigma (\psi )\) and \({\overline{\mathbb {D}}} \times \partial \Sigma \). It follows that they extend to coordinates on

$$\begin{aligned} V \cup _{\textrm{id}} \left( {\overline{\mathbb {D}}} \times \partial \Sigma \right) . \end{aligned}$$

In view of the expressions (4.5) and (4.6) for the contact forms \(\alpha _s\) on V, and \(\sigma _s\) on \({\overline{\mathbb {D}}} \times \partial \Sigma \), these two contact forms are glued to a smooth contact form \(\tau _s\) on \(M(\psi )\).

As mentioned above, the Reeb vector field \(R_{\tau _s}\) is positively transverse to the surfaces \(\{\theta \} \times \Sigma \) in \(\Sigma (\psi )\) and to the annuli \(A_{\varvec{\uptheta }}\) in \({\overline{\mathbb {D}}} \times \partial \Sigma \). Also, \(R_{\tau _s}\) is tangent to the core circle C of \({\overline{\mathbb {D}}} \times \partial \Sigma \), giving the positive orientation. It follows that the Reeb vector field of \(\Psi _* \tau _s\) is positively transverse to the interior of the pages of the open book \((\Sigma , \psi ,\Psi )\), and positively tangent to the binding of the open book. By Theorem 4.4 the contact structure \(\ker \Psi _* \tau _s\) is diffeomorphic to \(\xi \), by some diffeomorphism \(\rho _s :M \rightarrow M\). Summarizing, there are diffeomorphisms \(\Psi \) and \(\rho _s\) such that

$$\begin{aligned} \left( M(\psi ), \tau _s \right) {\mathop {\longrightarrow }\limits ^{\Psi }} \left( M, \Psi _* \tau _s \right) {\mathop {\longrightarrow }\limits ^{\rho _s}} \left( M, (\rho _s \circ \Psi )_*\tau _s \right) \end{aligned}$$
(4.11)

with \(\ker ( (\rho _s \circ \Psi )_* \tau _s) = \xi \).

Step 4: Estimating the volume and the topological entropy of \(\tau _s\). The Reeb flow \(\phi _{\tau _s}^t\) of \(\tau _s\) leaves the compact sets \(\Sigma (\psi )\) and \(\overline{\mathbb {D}} \times \partial \Sigma \) invariant. Since these compact sets cover M and since \(\phi _{\tau _s}^t |_{\Sigma (\psi )} = \phi _{\alpha _s}^t\) and \(\phi _{\tau _s}^t |_{\overline{\mathbb {D}} \times \partial \Sigma } = \phi _{\sigma _s}^t\), it follows from [60, Proposition 3.1.7 (2)] and from Lemma 4.6 and (4.10) that

$$\begin{aligned} h_{{\text {top}}}(\phi _{\tau _s}) = \max \left\{ h_{{\text {top}}}(\phi _{\alpha _s}), h_{{\text {top}}}(\phi _{\sigma _s}) \right\} \,\le \, E. \end{aligned}$$
(4.12)

We decompose the integral of \(\tau _s \wedge d \tau _s\) as

$$\begin{aligned} \int _{M(\psi )} \tau _s \wedge d \tau _s \,=\, \int _{\Sigma (\psi )} \alpha _s \wedge d \alpha _s + \int _{\overline{\mathbb {D}} \times \partial \Sigma } \sigma _s \wedge d \sigma _s. \end{aligned}$$
(4.13)

For the first summand we have by (4.2) that

$$\begin{aligned} \int _{\Sigma (\psi )} \alpha _s \wedge d \alpha _s \,= & {} \, \int _{[0,2\pi ] \times \Sigma } s \,d\theta \wedge \omega + O(s^2) \,\nonumber \\= & {} \, s \, 2\pi \int _\Sigma \omega + O(s^2) \,=\, O(s). \end{aligned}$$
(4.14)

To estimate the second term in (4.13) we use (4.7):

$$\begin{aligned} \int _{\overline{\mathbb {D}} \times \partial \Sigma } \sigma _s \wedge d \sigma _s \,= & {} \, \int _{\overline{\mathbb {D}} \setminus \{0\} \times \partial \Sigma } \sigma _s \wedge d \sigma _s \,\nonumber \\= & {} \, s \int _{\overline{\mathbb {D}} \setminus \{0\} \times \partial \Sigma } h({{\textbf {r}}}) \,d{{\textbf {r}}}\wedge d\varvec{\uptheta }\wedge d{{\textbf {x}}}. \end{aligned}$$
(4.15)

Since the right integral is finite, it follows that also

$$\begin{aligned} \int _{\overline{\mathbb {D}} \times \partial \Sigma } \sigma _s \wedge d \sigma _s = O(s). \end{aligned}$$

Together with (4.13) and (4.14) we conclude that

$$\begin{aligned} \int _{M(\psi )} \tau _s \wedge d \tau _s\,=\, O(s). \end{aligned}$$
(4.16)

Step 5: End of proof. By (4.16) we know that given \(\varepsilon >0\) there exists \(s \in (0,s_1]\) such that

$$\begin{aligned} {\text {vol}}_{\tau _s} \bigl ( M(\psi ) \bigr ) \,=\, \frac{1}{2\pi }\, \int _{M(\psi )} \tau _s \wedge d \tau _s \,\le \, \frac{\varepsilon ^2}{E^2}. \end{aligned}$$
(4.17)

Defining \(\widetilde{\tau }:= \left( {\text {vol}}_{\tau _s} \bigl ( M(\psi ) \bigr ) \right) ^{-\frac{1}{2}} \, \tau _s\) we obtain

$$\begin{aligned} {\text {vol}}_{\widetilde{\tau }} \bigl ( M(\psi ) \bigr ) \,=\, 1, \end{aligned}$$

and by (4.12) and (4.17)

$$\begin{aligned} h_{{\text {top}}}(\phi _{\widetilde{\tau }}) \,=\, \left( {\text {vol}}_{\tau _s} \bigl ( M(\psi ) \bigr ) \right) ^{1/2} \, h_{{\text {top}}} (\phi _{\tau _s}) \,\le \, \varepsilon . \end{aligned}$$

Together with (4.11) and in view of the conjugacy invariance of topological entropy if follows that \((\rho _s \circ \Psi )_* {\widetilde{\tau }}\) is a contact form on \((M,\xi )\) of volume 1 and topological entropy at most \(\varepsilon \). \(\square \)

Question 4.7

It would be interesting to see how in the case of a spherization \(S^*Q_2\) over the closed orientable surface of genus 2 the open book decomposition used in the above proof looks like. Since our construction of the contact form \({\widetilde{\tau }}\) is explicit, one could then maybe understand the star field \(\{ D_q^*(H) \}\) corresponding to \({\widetilde{\tau }}\). In view of (1.11) some of the stars must be very spiky. How many spikes appear in these stars?

5 Generalities on Giroux’s correspondence in higher dimensions

In this section, we summarize those concepts and results on the Giroux correspondence between contact structures and supporting open books in higher dimensions that we shall use in the proof of Theorem 1.1. While we reprove those parts that we use in a somewhat different form, we refer to [53] and [54] for the parts that we can cite and for further results.

5.1 Ideal Liouville domains

Let F be a 2n-dimensional compact manifold with non-empty boundary K, and denote by \(F^\circ \) the interior of F. A symplectic form \(\omega \) on \(F^\circ \) is called an ideal Liouville structure on F (abbreviated ILS) if \(\omega \) admits a primitive \(\lambda \) on \(F^\circ \) such that for some (and then any) smooth function

$$\begin{aligned} u :F \rightarrow [0,+\infty ) \; \text{ for } \text{ which } K=u^{-1}(0) \text{ is } \text{ a } \text{ regular } \text{ level } \text{ set } \end{aligned}$$
(5.1)

the 1-form \(u \lambda \) on \(F^\circ \) extends to a smooth 1-form \(\beta \) on F which is a contact form along K.

If such a 2-form \(\omega \) exists, the pair \((F,\omega )\) is called an ideal Liouville domain (ILD), and any primitive \(\lambda \) with the above property is called an ideal Liouville form (ILF). Given an ILD \((F,\omega )\), the contact structure

$$\begin{aligned} \xi := \ker (\beta |_{TK}) \end{aligned}$$

depends on the 2-form \(\omega \) but neither on \(\lambda \) nor u, see Proposition 2 in [54]. Moreover, once \(\lambda \) is chosen, one can recover every (positive) contact form on \((K,\xi )\) as the restriction to K of the extension of \(u\lambda \) for some function u with property (5.1). This is why the pair \((K,\xi )\) is called the ideal contact boundary of \((F,\omega )\). We note that the orientation of K that is determined by the co-oriented contact structure \(\xi \) coincides with the orientation of K as the boundary of \((F,\omega )\).

A very useful feature of an ILD is that a neighborhood of its boundary admits an explicit parametrization in which any ILF has a very nice form.

Lemma 5.1

Let \((F,\omega )\) be an ILD and \(\lambda \) be an ILF. Let u be a function satisfying (5.1) and let \(\beta \) be the extension of \(u\lambda \). Then for any contact form \(\alpha _0\) on \((K,\xi )\), there exists an embedding

$$\begin{aligned} \imath :[0,+\infty ) \times K \rightarrow F \end{aligned}$$

such that

$$\begin{aligned} \imath ^*\lambda =\frac{1}{r}\alpha _0 \, \text { and } \, \imath (0,q)=q \, \text { for all } \, q\in K, \end{aligned}$$

where \(r \in [0,+\infty )\). In particular,

$$\begin{aligned} \imath ^*\beta = \frac{u \circ \imath }{r}\alpha _0 \, \text { on }\, F^\circ \end{aligned}$$

and for all \(q\in K\),

$$\begin{aligned} (\beta |_{TK})(q) = \left( \frac{\partial (u \circ \imath )}{\partial r}(0,q) \right) \alpha _0. \end{aligned}$$

Proof

The above statement is a reformulation of Proposition 3 in [54]. We give a similar but more explicit proof.

Let \(\dim F=2n\). Since \(\beta \) is by assumption a positive contact form on K, \(\beta \wedge (d\beta )^{n-1}\) is a positive volume form on K. Using \(\omega = d \lambda = d(\beta / u)\) on \(F^\circ \) we compute

$$\begin{aligned} \omega ^n = \left( d(\beta /u) \right) ^n = u^{-n-1} (u \, d\beta + n \beta \wedge du) \wedge (d\beta )^{n-1} = u^{-n-1}\mu \end{aligned}$$
(5.2)

where \(\mu \) is the 2n-form

$$\begin{aligned} \mu := (u \, d\beta + n\beta \wedge du) \wedge (d\beta )^{n-1}. \end{aligned}$$

The above expression shows that \(\mu \) is smooth on F and, together with (5.2) and the fact that 0 is a regular value of u, that it is a positive volume form on F. Define the smooth vector field X on F by

$$\begin{aligned} \iota _X \mu = -n \beta \wedge (d\beta )^{n-1}. \end{aligned}$$
(5.3)

Recall that the Liouville vector field Y of \(\lambda \) is the vector field on \(F^\circ \) defined by \(\iota _Y d\lambda = \lambda \). Using \(\beta = u \lambda \) on \(F^\circ \) we compute

$$\begin{aligned} -n\beta \wedge (d\beta )^{n-1} = -n u^n \lambda \wedge (d\lambda )^{n-1} = -u^n \iota _Y \omega ^n = -u^{-1} \iota _Y \mu . \end{aligned}$$

Comparing with (5.3) we find \(Y = -u X\). Then

$$\begin{aligned} \beta (X) = - \lambda (Y) = - d\lambda (Y,Y) = 0 \end{aligned}$$

on \(F^{\circ }\), and by continuity \(\beta (X)=0\) on F. Hence on \(F^\circ \),

$$\begin{aligned} L_X \beta= & {} \iota _X d\beta = -\frac{1}{u} \iota _Y \left( du \wedge \lambda + u d\lambda \right) \nonumber \\= & {} -\frac{1}{u} \left( du(Y) \lambda + u \lambda \right) = \frac{1}{u} \left( du(X)-1\right) \beta . \end{aligned}$$
(5.4)

This shows that \(du (X) =1\) along K and that the function \(\frac{1}{u} (du(X)-1)\) is smooth on F. In particular, X points inwards on \(K=\partial F\) and hence the flow \(\phi _X^t\) of X on the compact manifold F is well-defined for every \(t\ge 0\). We define the smooth embedding

$$\begin{aligned} \Phi :[0,+\infty ) \times K \rightarrow F, \quad (t,q) \mapsto \phi ^t_{X}(q). \end{aligned}$$

By construction we have \(\Phi ^*X=\partial _t\). Put \(\hat{\beta }:= \Phi ^*\beta \), \(\hat{u}:= \Phi ^*u\), and \(\hat{\lambda }:= \Phi ^*\lambda \). The identities \(\beta (X)=0\) and (5.4) say that on \([0,+\infty ) \times K\),

$$\begin{aligned} \hat{\beta }(\partial _t)=0, \quad \partial _t \hat{\beta } = \frac{\partial _t {\hat{u}} -1}{{\hat{u}}} {\hat{\beta }}. \end{aligned}$$
(5.5)

Here the function \(\frac{\partial _t {\hat{u}} -1}{{\hat{u}}}\) is smooth and bounded on \([0,+\infty ) \times K\) since by (5.4) the function \(\frac{1}{u} (du(X)-\!1)\) is smooth on F. Define the smooth function \(v_0 :[0,+\!\infty ) \!\times \! K \rightarrow \mathbb {R}\) by

$$\begin{aligned} v_0(t,q) \,=\, \int _0^t \frac{\partial _t \hat{u} (\tau ,q)-1}{\hat{u}(\tau ,q)} \,d\tau . \end{aligned}$$

The solution of the problem (5.5) with initial condition \(\beta _0(q) = \beta (0,q)\) is then

$$\begin{aligned} \hat{\beta }(t,q) = \exp \left( v_0(t,q) \right) \beta _0(q), \qquad \forall \, (t,q)\in [0,+\infty ) \times K, \end{aligned}$$

and therefore

$$\begin{aligned} \hat{\lambda }(t,q) \,=\, \frac{1}{\hat{u}(t,q)} \exp \left( v_0(t,q) \right) \beta _0(q), \qquad \forall \, (t,q)\in (0,+\infty ) \times K. \end{aligned}$$

Now let \(\alpha _0\) be a positive contact form on \((K,\xi )\). Then there is a positive function \(\kappa \) on K such that \(\beta _0 = \kappa \alpha _0\). On \((0,+\infty )\times K\) define the function

$$\begin{aligned} \Lambda (t,q) \,=\, \frac{\kappa (q)}{\hat{u}(t,q)} \exp \left( v_0(t,q) \right) . \end{aligned}$$
(5.6)

Then \(\hat{\lambda } = \Lambda \alpha _0\). It is clear that \(\Lambda >0\), and \(\lim _{t \rightarrow 0} \Lambda (t,q) = +\infty \) for all \(q \in K\). We note that

$$\begin{aligned} \frac{\partial \Lambda }{\partial t} =-\frac{\Lambda }{\hat{u}}<0 \end{aligned}$$

and therefore

$$\begin{aligned} \Lambda (t,q) \,=\, \Lambda (1,q) \exp \left( -\int _{1}^t\frac{1}{\hat{u}(\tau ,q)} \,d\tau \right) . \end{aligned}$$
(5.7)

On \([0,+\infty ) \times K\), \(\hat{u}\) is bounded from above since F is compact. Therefore \(\lim _{t \rightarrow +\infty } \Lambda (t,q)=0\) for all \(q \in K\). It follows that \(\Lambda (\cdot ,q)\) is an orientation reversing diffeomorphism from \((0,+\infty )\) onto \((0,+\infty )\) for all q. Hence there exists a positive smooth function f on \((0,+\infty ) \times K\) such that

$$\begin{aligned} \Lambda (f(r,q),q) \,=\, \frac{1}{r} \quad \forall \, (r,q) \in (0,+\infty ) \times K, \end{aligned}$$

and for every \(q\in K\) the function \(f(\cdot ,q)\) is an orientation preserving diffeomorphism from \((0,+\infty )\) onto \((0,+\infty )\). Define the embedding

$$\begin{aligned} \Psi :(0,+\infty ) \times K \rightarrow [0,+\infty ) \times K, \quad (r,q) \mapsto (f(r,q),q). \end{aligned}$$

By construction \(\Psi ^*\hat{\lambda }=\frac{1}{r}\alpha _0\). We claim that \(\Psi \) extends to a smooth embedding

$$\begin{aligned} \Psi :[0,+\infty ) \times K \rightarrow [0,+\infty ) \times K \quad \text{ with } \Psi (0,q)=(0,q). \end{aligned}$$

Postponing the proof of the claim, we note that \(\imath = \Phi \circ \Psi \) is then the desired embedding. The rest of the statement of the lemma follows immediately from the identity \(\imath ^*\lambda = \frac{1}{r} \alpha _0\).

We now show that the extension of \(\Psi \) given by \(\Psi (0,q) = (0,q)\) is smooth. Combining (5.6) and (5.7) we get

$$\begin{aligned} \Lambda (1,q)^{-1}\exp \left( \int _{1}^{f(r,q)}\frac{1}{\hat{u}(\tau ,q)} \,d\tau \right) =r. \end{aligned}$$
(5.8)

We consider the function

$$\begin{aligned} g(t,q) :=\, \Lambda (1,q)^{-1} \exp \left( \int _{1}^{t} \frac{1}{\hat{u}(\tau ,q)} \,d\tau \right) \end{aligned}$$

on \((0,+\infty ) \times K\) and we define

$$\begin{aligned} \widetilde{\Psi } :(0,+\infty ) \times K \rightarrow (0,+\infty ) \times K, \quad \widetilde{\Psi }(t,q)=(g(t,q),q). \end{aligned}$$

Then we have

$$\begin{aligned} \widetilde{\Psi } \circ \Psi (r,q) \,=\, \widetilde{\Psi } (f(r,q),q) \,=\, \bigl ( g(f(r,q),q),q \bigr ) \,=\, (r,q) \end{aligned}$$

on \((0,+\infty )\times K\). We claim that \(\widetilde{\Psi }\) extends smoothly to \([0,+\infty )\times K\) by \(\widetilde{\Psi }(0,q)=q\) for all \(q\in K\). To see this we define the function \(v_1 :[0,+\infty ) \times K \rightarrow \mathbb {R}\) by

$$\begin{aligned} v_1(t,q) :=\, \int _1^t \frac{\partial _t {\hat{u}}(\tau ,q)-1}{{\hat{u}}(\tau ,q)} \,d\tau . \end{aligned}$$

We recall that the above integrand is smooth and bounded on \([0,+\infty ) \times K\) and so is the function \(v_1\). For every \(t \in (0,+\infty )\) and \(q \in K\),

$$\begin{aligned} e^{v_1} \,=\, \exp \left( \int _1^t \frac{\partial _t {\hat{u}}-1}{{\hat{u}}} \, d\tau \right)= & {} \exp \left( \log {\hat{u}}(t,q)-\log {\hat{u}}(1,q) -\int _1^t \frac{1}{{\hat{u}}} \, d\tau \right) \\= & {} \frac{{\hat{u}}(t,q)}{{\hat{u}}(1,q)} \exp \left( -\int _1^t \frac{1}{{\hat{u}}} \, d\tau \right) \end{aligned}$$

and so

$$\begin{aligned} g(t,q) \,=\, \Lambda (1,q)^{-1} \, e^{-v_1(t,q)} \, \frac{{\hat{u}}(t,q)}{{\hat{u}}(1,q)}. \end{aligned}$$

Note that \(\Lambda (1,q)\ne 0\). The above expression says that g extends smoothly to \([0,+\infty ) \times K\) by \(g(0,q)=0\) for \(q \in K\). For \(t>0\) we compute

$$\begin{aligned} \partial _t g (t,q)= & {} \Lambda (1,q)^{-1} \,e^{-v_1(t,q)} \left[ - \partial _t v_1 (t,q) \, \frac{{\hat{u}}(t,q)}{{\hat{u}}(1,q)} + \frac{\partial _t {\hat{u}} (t,q)}{{\hat{u}}(1,q)} \right] \\= & {} \Lambda (1,q)^{-1} \, e^{-v_1(t,q)} \left[ -\frac{(\partial _t {\hat{u}}(t,q)-1)}{{\hat{u}}(t,q)} \frac{{\hat{u}}(t,q)}{{\hat{u}}(1,q)} + \frac{\partial _t {\hat{u}}(t,q)}{{\hat{u}}(1,q)} \right] \\= & {} \Lambda (1,q)^{-1} \, e^{-v_1(t,q)} \,\frac{1}{{\hat{u}}(1,q)}. \end{aligned}$$

By the smoothness of g, this expression also holds true for \(t=0\). In particular, \(\partial _t g(t,q) >0\) for all \((t,q) \in [0,+\infty ) \times K\). It follows that \(D \widetilde{\Psi }(t,q)\) is invertible for all \((t,q) \in [0,+\infty ) \times K\). By the inverse function theorem, the extension of \(\Psi \) over \([0,+\infty )\times K\) is \(C^1\) and in fact \(C^\infty \)-smooth since \(\widetilde{\Psi }\) is smooth. \(\square \)

5.2 Ideal Liouville domains and contact structures

Ideal Liouville domains are particularly useful for clarifying the existence and uniqueness of contact structures supported by open books in higher dimensions. We first recollect some facts about open books.

An open book in a closed manifold M is a pair \((K,\Theta )\) where

  1. (ob1)

    \(K\subset M\) is a closed submanifold of co-dimension two with trivial normal bundle;

  2. (ob2)

    \(\Theta :M \setminus K \rightarrow S^1= \mathbb {R}/2 \pi \mathbb {Z}\) is a locally trivial smooth fibration that on a deleted neighbourhood \((\mathbb {D}\setminus \{0\}) \times K\) of K reads \(\Theta (re^{i\theta },q) = \theta \).

The submanifold K is called the binding of the open book, and the closures of the fibres of \(\Theta \) are called the pages. The pages are compact submanifolds with common boundary K. The canonical orientation of \(S^1\) induces co-orientations of the pages. Hence if M is oriented, then so are the pages, and then also the binding as the boundary of a page.

Another way of defining an open book is as follows. Let \(h :M \rightarrow \mathbb {C}\) be a smooth function such that 0 is a regular value. Set \(K:= h^{-1}(0)\), and assume that \(\Theta :=h/|h| :M {\setminus } K \rightarrow S^1\) has no critical points. Then the pair \((K,\Theta )\) is an open book in M. Moreover, any open book in M can be recovered via a defining function h as above, and such a defining function is unique up to multiplication by a positive function on M.

Given an open book \((K,\Theta )\) in a closed manifold M, one finds a vector field X on M, called a spinning vector field, such that

  1. (m1)

    X lifts to a smooth vector field on the manifold with boundary obtained from M by real oriented blow-up along K, in which each disk \(\mathbb {D}\times \{q\}\) of the neighbourhood \(\mathbb {D}\times K\) is replaced by the annulus \(S^1 \times [0,1] \times \{q\}\);

  2. (m2)

    \(X=0\) on K and \(d\Theta (X) = 1\) on \(M {\setminus } K\).

Then the time-\(2\pi \) map of the flow of X is a diffeomorphism

$$\begin{aligned} \phi :F\rightarrow F \end{aligned}$$

of the 0-page \(F:= \Theta ^{-1}(0)\cup K\), which fixes K pointwise. The isotopy class \([\phi ]\) of \(\phi \) among the diffeomorphisms of F that fix K pointwise is called the monodromy of the open book. It turns out that the open book is characterized by the pair \((F,[\phi ])\). Namely, given the pair \((F,\phi )\), one defines the mapping torus

$$\begin{aligned} {\text {MT}}(F,\phi ):= ([0,2\pi ] \times F ) \big /\sim \quad \text{ where } (2\pi ,p) \sim (0,\phi (p)). \end{aligned}$$

This is a manifold with boundary. One has the natural fibration

$$\begin{aligned} \widehat{\Theta } :{\text {MT}}(F,\phi )\rightarrow S^1 \end{aligned}$$

with fibres diffeomorphic to F, and there is a natural parametrization of the fibre \(\widehat{\Theta }^{-1}(0)\) via the restriction of the above quotient map to \(\{0\} \times F\). For every \(\phi '\in [\phi ]\) there is a diffeomorphism between \({\text {MT}}(F,\phi )\) and \({\text {MT}}(F,\phi ')\) that respects the fibrations over \(S^1\) and the natural parametrizations of the 0-fibres. Now, given \({\text {MT}}(F,\phi )\) one collapses its boundary, which is diffeomorphic to \(S^1 \times K\), to K and obtains the so-called abstract open book \({\text {OB}}(F,\phi )\). In fact, the closed manifold \({\text {OB}}(F,\phi )\) admits an open book given by the pair \((K,\Theta )\), where \(\Theta \) is induced by \(\widehat{\Theta }\). Moreover, for \(\phi ' \in [\phi ]\), the diffeomorphism between \({\text {MT}}(F,\phi )\) and \({\text {MT}}(F,\phi ')\) descends to a diffeomorphism between the corresponding abstract open books. In particular, M and \({\text {OB}}(F,\phi )\) may be identified together with their open book structures. We note that one can choose the spinning vector field X smooth on M and such that its flow is \(2\pi \)-periodic near K. However, not every representative of the monodromy class can be obtained via a smooth spinning vector field, see Remark 12 in [54]. To obtain all representatives of the monodromy class, one needs to use the whole affine space of spinning vector fields.

Open books meet with contact topology via the following definition. Let M be a compact manifold with a co-oriented contact structure \(\xi \). We say that \(\xi \) is supported by an open book \((K,\Theta )\) on M or that the open book \((K,\Theta )\) supports \(\xi \) if there exists a contact form \(\alpha \) on \((M,\xi )\), that is \(\xi =\ker \alpha \), such that

  • \(\alpha \) restricts to a positive contact form on K;

  • \(d\alpha \) restricts to a positive symplectic form on each fibre of \(\Theta \).

It turns out that given a closed manifold M, the isotopy classes of co-oriented contact structures on M are in one-to-one correspondence with equivalence classes of supporting open books. This statement is a very rough summary of what is called the Giroux correspondence. We will recall certain pieces of this celebrated statement in detail.

Theorem 5.2

(Theorem 10 in [53]) Any contact structure on a compact manifold is supported by an open book with Weinstein pages.

This result is the core of the correspondence between supporting open books and contact structures. The existence statement of the opposite direction of the correspondence is relatively easy to achieve, especially in dimension three. Namely, given an open book in a 3-dimensional compact manifold, it is not hard to construct a contact form on the corresponding abstract open book whose kernel is supported. In higher dimensions, however, one needs that the pages are exact symplectic and that the monodromy is symplectic in order to construct a contact form on an abstract open book whose kernel is supported, see Proposition 9 in [53] and Proposition 17 in [54]. We will carry out such a construction in Sects. 6.1 and 6.2. Concerning the uniqueness features of the Giroux correspondence, we are mainly interested in one side, namely the “uniqueness” of supported contact structures. This result is again more involved in higher dimensions. Heuristically, given an open book, the symplectic geometry of the pages determines the supported contact structures. In dimension three, any two symplectic structures on a page are isotopic since they are just area forms on a given surface, but in higher dimensions this is not the case.

In [54] Giroux introduced the notion of a Liouville open book, which clears out the technicalities to which we pointed above.

A Liouville open book (LOB) in a closed manifold M is a triple \((K,\Theta ,(\omega _\theta )_{\theta \in S^1})\) where

  1. (lob1)

    \((K,\Theta )\) is an open book on M with pages \(F_\theta =\Theta ^{-1}(\theta )\cup K\), \(\theta \in S^1\);

  2. (lob2)

    \((F_\theta ,\omega _\theta )\) is an ILD for all \(\theta \in S^1\) and the following holds: there is a defining function \(h :M \rightarrow \mathbb {C}\) for \((K,\Theta )\) and a 1-form \(\beta \) on M such that the restriction of \(d(\beta /|h|)\) to each page is an ILF. More precisely,

    $$\begin{aligned} \omega _\theta =d(\beta /|h|) |_{TF_\theta ^\circ } \end{aligned}$$

    for all \(\theta \in S^1\).

The 1-form \(\beta \) is called a binding 1-form associated to h. If \(h'\) is another defining function for \((K,\Theta )\), then \(h'=\kappa h\) for a positive function \(\kappa \) on M, and \(\beta ':= \kappa \beta \) is a binding 1-form associated to \(h'\). We also note that for a fixed defining function, the set of associated binding 1-forms is an affine space.

Similar to the case of classical open books, LOBs are characterized by their monodromy, which now has to be symplectic: One considers symplectically spinning vector fields, namely vector fields X satisfying (m1) and (m2) and generating the kernel of a closed 2-form on \(M \setminus K\) which restricts to \(\omega _\theta \) for all \(\theta \in S^1\). Given such a vector field, the time-\(2\pi \) map of its flow, say \(\phi \), is a diffeomorphism of \(F:= F_0\) which fixes K and preserves \(\omega := \omega _0\). The isotopy class \([\phi ]\), among the symplectic diffeomorphisms that fix K, is called the symplectic monodromy and characterizes the given LOB. For the construction of a LOB in the abstract open book \({\text {OB}}(F,\phi )\), where \(\phi ^*\omega =\omega \), we refer to Proposition 17 in [54] and to our construction at the end of the next section.

Again, symplectically spinning vector fields form an affine space, and all representatives of the symplectic monodromy can be obtained by sweeping out this affine space. The obvious choice of a symplectically spinning vector field X is smooth, and by suitably modifying a given binding 1-form without affecting its restriction to the kernel of \(d\Theta \) one can arrange that the flow of X is \(2\pi \)-periodic near the binding.

Lemma 5.3

(Lemma 15 in [54]) Let \((K, \Theta ,(\omega _\theta )_{\theta \in S^1})\) be a LOB in a closed manifold M, and let \(h :M \rightarrow \mathbb {C}\) be a defining function for \((K,\Theta )\). Then for every binding 1-form \(\beta \), the vector field X on \(M {\setminus } K\) spanning the kernel of \(d(\beta /|h|)\) and satisfying \(d \Theta (X) = 1\) extends to a smooth vector field on M which is zero along K. Furthermore, \(\beta \) can be chosen such that the flow of X is \(2\pi \)-periodic near K.

Natural sources of LOBs are contact manifolds:

Proposition 5.4

(Proposition 18 in [54]) Let \((M, \xi )\) be a closed contact manifold, and let \((K, \Theta )\) be a supporting open book with defining function \(h :M \rightarrow \mathbb {C}\). Then the contact forms \(\alpha \) on \((M,\xi )\) such that \(d(\alpha /|h|)\) induces an ideal Liouville structure on each page form a non-empty convex cone.

Let \((K,\Theta ,(\omega _\theta )_{\theta \in S^1})\) be a LOB in a closed manifold M with a defining function h. A co-oriented contact structure \(\xi \) on M is said to be symplectically supported  by \((K,\Theta ,(\omega _\theta )_{\theta \in S^1})\) if there exists a contact form \(\alpha \) on \((M,\xi )\) such that \(\alpha \) is a binding 1-form of the LOB associated to h.

By our remark following the definition of a binding 1-form, the property of being symplectically supported is independent of the given defining function. But the crucial fact is that once a defining function is fixed, a contact binding 1-form is unique whenever it exists, see Remark 20 in [54]. Hence, once a defining function h is fixed, there is a one-to-one correspondence between contact structures supported by \((K,\Theta ,(\omega _\theta )_{\theta \in S^1})\) and contact binding 1-forms associated to h.

Given two contact structures \(\xi _0\) and \(\xi _1\) supported by \((K,\Theta ,(\omega _\theta )_{\theta \in S^1})\), after fixing h we therefore have unique contact binding 1-forms \(\alpha _0\) and \(\alpha _1\), respectively. Since the set of binding 1-forms associated to h is affine, there is a path \((\beta _t)_{t\in [0,1]}\) of binding 1-forms such that \(\beta _0 = \alpha _0\) and \(\beta _1 = \alpha _1\). Now it is not hard to explicitely deform the forms \(\beta _t\) without affecting their restrictions to \(\ker d\Theta \) in such a way that

  • for all \(s \ge 0\) and \(t \in [0,1]\), \(\beta ^s_t\) is a binding 1-form for \((K,\Theta ,(\omega _\theta )_{\theta \in S^1})\) associated to h (since the deformation of \(\beta _t\) leaves unchanged the restriction to the pages);

  • \(\beta _t^s\) is a contact form for s large enough, uniformly in \(t \in [0,1]\);

  • if \(\beta _t\) is already a contact form, then \(\beta _t^s\) is a contact form for all \(s \ge 0\).

By the first property of these deformations and by the uniqueness discussed above, whenever \(\beta ^s_t\) is a contact form then \(\ker \beta ^s_t\) is symplectically supported by \(\left( K,\Theta ,(\omega _\theta )_{\theta \in S^1}\right) \) and \(\beta _t^s\) is the unique contact binding 1-form associated to h. Together with the other two properties we see that there exists \(c>0\) such that the concatenation of the paths \((\ker \beta _0^s)_{s\in [0,c]}\), \((\ker \beta ^c_t)_{t\in [0,1]}\) and \((\ker \beta ^{c-s}_1)_{s\in [0,c]}\) gives an isotopy between \(\xi _0\) and \(\xi _1\) along contact structures that are symplectically supported by \((K,\Theta ,(\omega _\theta )_{\theta \in S^1})\) (cf. Fig. 8).

Fig. 8
figure 8

Constructing an isotopy between \(\xi _0\) and \(\xi _1\)

In fact the following more general statement holds.

Proposition 5.5

(Proposition 21 in [54]) On a closed manifold, contact structures supported by a given Liouville open book form a non-empty and weakly contractible subset in the space of all contact structures.

6 Entropy collapse for Reeb flows in dimension \(\ge 3\)

This section is devoted to the proof of the following main result.

Theorem 6.1

Let \((M,\xi )\) be a closed co-orientable contact manifold. Then for every real number \(\varepsilon >0\) there exists a contact form \(\alpha \) for \(\xi \) such that \({\widehat{\Gamma }}(\alpha ) \le \varepsilon \).

We prove the statement by induction on n, where \(\dim M=2n+1\). The initial case \(n=0\) is clear: Then M is a circle, and the Reeb flow generated by the vector field \(\partial _\theta \) has vanishing norm growth. It may be interesting to read the subsequent proof for \(n=0\) and to compare the line of argument with the one of Sect. 4.

We now assume by induction that Theorem 6.1 holds for \(n-1 \ge 0\) and fix a contact manifold \((M,\xi )\) of dimension \(2n+1\). By Theorem 5.2 there exists an open book \((K,\Theta )\) in M supporting \(\xi \). Let \(F_\theta := \Theta ^{-1}(\theta ) \cup K\), \(\theta \in S^1=\mathbb {R}/2\pi \mathbb {Z}\), denote the pages of the open book and let \(h :M \rightarrow \mathbb {C}\) be a defining function for \((K,\Theta )\). We wish to construct a contact form on the abstract open book defined via the 0-page

$$\begin{aligned} F:= F_0 = \Theta ^{-1}(0) \cup K. \end{aligned}$$
(6.1)

By Proposition 5.4, there exists a contact form \(\alpha \) on \((M,\xi )\) such that the triple \((K,\Theta , d(\alpha /|h|)|_{TF_\theta ^\circ })\) is a LOB which supports \(\xi \) symplectically. By Lemma 5.3, we can modify the contact binding 1-form \(\alpha \) without affecting its restriction to the kernel of \(d\Theta \), to obtain a binding 1-form \({\widehat{\alpha }}\), not necessarily contact, such that the flow of the associated symplectically spinning vector field X is \(2\pi \)-periodic near K. Hence the time-\(2\pi \) map of the flow of X gives us a diffeomorphism \(\psi :F \rightarrow F\) such that

$$\begin{aligned} \psi ^*(d\lambda ) = d\lambda \end{aligned}$$
(6.2)

where \(\lambda \in \Omega ^1(F^\circ )\) is the ILF

$$\begin{aligned} \lambda := ({\widehat{\alpha }} /|h|) |_{TF^\circ }=(\alpha /|h|) |_{TF^\circ } \end{aligned}$$
(6.3)

and \(\psi =\textrm{id}\) on some neighbourhood of K in F. Now our aim is to recover M as the abstract open book induced by the pair \((F,\psi )\) and to define a contact form on the abstract open book with small norm growth. We first consider the mapping torus

$$\begin{aligned} {\text {MT}}(F,\psi ):= \bigl ( [0,2\pi ] \times F \bigr ) \big / \bigl ( (2\pi ,p)\sim (0,\psi (p)) \bigr ). \end{aligned}$$

Since \(\psi =\textrm{id}\) on some neighbourhood of K, the boundary \(\partial {\text {MT}}(F,\psi )\) has an open neighbourhood given as a product of K with an annulus, in which we collapse the boundary and get the abstract open book \({\text {OB}}(F,\psi )\). We postpone the precise collapsing procedure since it will involve choices of coordinates, but note that the abstract open book is independent of these choices, and that we can make the identifications

$$\begin{aligned} {\text {MT}}(F^\circ ,\psi ) = {\text {MT}}(F,\psi ) \setminus \partial {\text {MT}}(F,\psi ) = {\text {OB}}(F,\psi )\setminus K. \end{aligned}$$

6.1 A family of contact forms away from the binding

On \([0,2 \pi ]\times F^\circ \) with \(\theta \) the coordinate on \([0,2\pi ]\), we define the family of 1-forms

$$\begin{aligned} \alpha _s = d\theta +s \left( \lambda + \chi (\theta ) \lambda _\psi \right) , \qquad s > 0 \end{aligned}$$
(6.4)

where \(\lambda _\psi := \psi ^*\lambda -\lambda \) and \(\chi :[0,2\pi ] \rightarrow [0,1]\) is a smooth function such that \(\chi (0)=0\), \(\chi (2\pi )=1\) and \(\chi '\) has support in \((0,2\pi )\). By the choice of \(\chi \), each 1-form \(\alpha _s\) descends to a 1-form on \({\text {MT}}(F^\circ ,\psi )\), that we still denote by \(\alpha _s\).

Lemma 6.2

There exists \(s_0>0\), depending on \(\psi , \lambda , \chi \), such that \(\alpha _s\) is a contact form on \({\text {MT}}(F^\circ ,\psi )\) for all \(s \in (0,s_0]\).

Proof

Since \(d\lambda _\psi =0\), we get \(d\alpha _s = s \left( \chi ' d\theta \wedge \lambda _\psi +d\lambda \right) \) and

$$\begin{aligned} \alpha _s \wedge (d\alpha _s)^{n}= & {} \bigl ( d\theta +s \left( \lambda + \chi \lambda _\psi \right) \bigr ) \wedge s^n \Bigl ( n \chi ' d\theta \wedge \lambda _\psi \wedge (d\lambda )^{n-1}+(d\lambda )^n \Bigr ) \\= & {} s^n d\theta \wedge \Bigl ( (d\lambda )^n - n s \chi ' \lambda \wedge \lambda _\psi \wedge (d\lambda )^{n-1} \Bigr ). \end{aligned}$$

Since \(d\theta \wedge (d\lambda )^n\) is a volume form and since with \(\lambda _\psi \) also \(\lambda \wedge d\theta \wedge \lambda _\psi \wedge (d\lambda )^{n-1}\) is compactly supported in \({\text {MT}}(F^\circ ,\psi )\), there exists \(s_0>0\) such that \(\alpha _s \wedge (d\alpha _s)^{n}\) is a positive volume form for all \(s \in (0,s_0]\). \(\square \)

Next, we study the Reeb vector field \(R_{\alpha _s}\) of \(\alpha _s\) on \({\text {MT}}(F^\circ ,\psi )\). Let Y be the vector field on \({\text {MT}}(F^\circ ,\psi )\) that is tangent to \(\{\theta \}\times F^\circ \) for each \(\theta \) and along each \(\{\theta \} \times F^\circ \) satisfies

$$\begin{aligned} \imath _Yd\lambda = -\chi '\lambda _\psi . \end{aligned}$$

Since \(\chi '=0\) near 0 and \(\pi \), Y is well defined, and since \(\psi \) is compactly supported in \(F^\circ \), Y is compactly supported in \({\text {MT}}(F^\circ ,\psi )\). We compute

$$\begin{aligned} \imath _{(\partial _\theta +Y)} d \alpha _s= & {} s \bigl ( \imath _{(\partial _\theta +Y)} \chi ' d\theta \wedge \lambda _\psi + \imath _{(\partial _\theta +Y)} d \lambda \bigr ) \\= & {} s \, \chi ' \bigl ( \lambda _\psi - \lambda _\psi (Y) \, d\theta - \lambda _\psi \bigr ) \\= & {} s \, d\lambda (Y,Y)\, d\theta \,=\, 0. \end{aligned}$$

Hence on \({\text {MT}}(F^\circ ,\psi )\) the Reeb vector field of \(\alpha _s\) is

$$\begin{aligned} R_{\alpha _s} =\frac{\partial _\theta +Y}{\alpha _s(\partial _\theta +Y)}. \end{aligned}$$
(6.5)

Note that \(R_{\alpha _s} = \partial _\theta \) near K. Since the \(\partial _\theta \) component of \(R_{\alpha _s}\) never vanishes and since Y is tangent to the pages, \(R_{\alpha _s}\) is transverse to \(F^\circ \times \{\theta \}\) for all \(\theta \). Hence \(F^\circ \) is a global hypersurface of section for \(R_{\alpha _s}\) on \({\text {MT}}(F^\circ ,\psi )\). We have the first return time map

$$\begin{aligned} T_s :F^\circ \rightarrow \mathbb {R}, \quad T_s(p) = \inf \left\{ t>0 \mid \phi _{R_{\alpha _s}}^t (0,p) \in \{0\} \times F^\circ \right\} \end{aligned}$$
(6.6)

and the first return map

$$\begin{aligned} \Upsilon :F^\circ \rightarrow F^\circ , \quad (0,\Upsilon (p)) = \phi _{R_{\alpha _s}}^{T_s(p)}(0,p) \quad \forall \; p \in F^\circ . \end{aligned}$$
(6.7)

Remark 6.3

Since \(R_{\alpha _s}\) is a multiple of the vector field \(\partial _\theta +Y\) that does not depend on s, the return map \(\Upsilon \) is independent of s. This justifies the absence of the subscript in (6.7).

We note that for all \(s \in (0,s_0]\),

$$\begin{aligned} T_s \equiv 2\pi \, \text{ and } \, \Upsilon = \textrm{id} \; \text{ on } F^\circ \setminus \textrm{supp}\,\psi . \end{aligned}$$
(6.8)

Recall that we write \(\Gamma (\alpha _s)\) for the norm growth \(\Gamma (\phi _{\alpha _s})\) of the Reeb flow \(\phi _{\alpha _s}^t\).

Lemma 6.4

There exists \(s_1 \in (0,s_0)\) such that for every \(s \in (0,s_1]\),

$$\begin{aligned} \pi \le T_s \le 4\pi \quad \text{ on } F^\circ \end{aligned}$$
(6.9)

and such that

$$\begin{aligned} \Gamma (\alpha _s) \le E \end{aligned}$$
(6.10)

for some constant \(E >0\) that depends only on \(\psi , \lambda , \chi \).

Proof

We compute

$$\begin{aligned} d\theta \left( R_{\alpha _s}\right) = \frac{1}{\alpha _s(\partial _\theta +Y)} = \frac{1}{1 + s \bigl (\lambda (Y) + \chi \lambda _\psi (Y) \bigr )}. \end{aligned}$$

Since Y is compactly supported, we find \(s_1 \in (0,s_0)\) such that for every \(s \in (0,s_1]\),

$$\begin{aligned} \frac{1}{2} \le \frac{1}{\alpha _s(\partial _\theta +Y)} \le 2 \quad \text{ on } {\text {MT}}(F^\circ ,\psi ). \end{aligned}$$
(6.11)

The inequality (6.9) follows. For the second claim, we apply Proposition C.2 to the vector field \(\partial _\theta + Y\) and the positive function \(\frac{1}{\alpha _s(\partial _\theta +Y)}\), and in view of (6.5) and (6.11) find that \(\Gamma (\alpha _s) \le 2 \, \Gamma (\phi _{\partial _\theta +Y}) =:E\). \(\square \)

6.2 A family of contact forms near the binding

Let \(E>0\) be the constant from Lemma 6.4. By our inductive hypothesis, for any \(\varepsilon >0\) there exists a contact form \(\sigma _\varepsilon \) on \((K, \xi |_K)\) such that

$$\begin{aligned} {\text {vol}}_{\sigma _\varepsilon } (K) = \varepsilon \quad \text{ and } \quad \Gamma (\sigma _\varepsilon ) \le E. \end{aligned}$$
(6.12)

Indeed, there is a contact form \(\alpha _0\) on \((K, \xi |_K)\) such that

$$\begin{aligned} {\text {vol}}_{\alpha _0} (K) = 1 \quad \text{ and } \quad \Gamma (\alpha _0) \le \varepsilon ^{1/n} E. \end{aligned}$$

We can thus take \(\sigma _\varepsilon := \varepsilon ^{1/n} \, \alpha _0\).

Applying Lemma 5.1 to \(\sigma _\varepsilon \) we obtain an embedding

$$\begin{aligned} \imath _\varepsilon :[0,+\infty ) \times K \hookrightarrow F \, \text{ such } \text{ that } \, \imath _\varepsilon ^* \lambda = \frac{1}{r} \sigma _\varepsilon , \end{aligned}$$
(6.13)

and \(\imath _\varepsilon (0,q)=q\) for every \(q \in K\). This embedding induces the smooth coordinate \(r \in [0,+\infty )\) on a neighborhood of \(K=\partial F\) in F. There exists \(r_{\varepsilon } > 0\) that depends only on \(\psi \) and \(\sigma _\varepsilon \) such that

$$\begin{aligned} \imath _\varepsilon \left( [0,r_\varepsilon ] \times K \right) \cap \, \mathrm {supp\,}\psi = \emptyset . \end{aligned}$$
(6.14)

We define

$$\begin{aligned} F_\varepsilon := F \setminus \bigl ( \imath _\varepsilon ([0,r_\varepsilon ) \times K) \bigr ) \end{aligned}$$
(6.15)

and note that near the boundary of \({\text {MT}}(F_\varepsilon ,\psi )\) the expression (6.4) for \(\alpha _s\) reads

$$\begin{aligned} \alpha _s = d\theta +\frac{s}{r} \, \sigma _\varepsilon . \end{aligned}$$
(6.16)

Lemma 6.5

For every \(\varepsilon >0\) and \(s \in (0,r_\varepsilon /2)\) there exist smooth functions

$$\begin{aligned} f,g :[0,r_\varepsilon ] \rightarrow \mathbb {R}\end{aligned}$$

with the following properties.

  1. (f1)

    \(f(r)=s/r\) near \(r=r_\varepsilon \) and \(f(r)=1\) near \(r=0\).

  2. (f2)

    \(f(r_\varepsilon /2)=1/2\).

  3. (f3)

    \(-2/r_\varepsilon \le f' \le 0 \) on \([0,r_\varepsilon ]\) and \(f'<0\) on \([r_\varepsilon /2, r_\varepsilon ]\).

  4. (g1)

    \(g=1\) on \([r_\varepsilon /2,r_\varepsilon ]\) and \(g(r)=r^2/2\) near \(r=0\).

  5. (g2)

    \(0\le g'\le 4/r_\varepsilon \) on \([0,r_\varepsilon ]\) and \(0 < g'\) on \((0,r_\varepsilon /2]\).

Fig. 9
figure 9

The functions f and g

See Fig. 9. The easy proof is left to the reader. For later use we note that the function.

$$\begin{aligned} h:= fg' - f'g :[0,r_\varepsilon ] \rightarrow \mathbb {R}\end{aligned}$$

is positive on \((0,r_\varepsilon ]\), satisfies \(h(r)=r\) near \(r=0\), and

$$\begin{aligned} h \,\le \, \frac{4}{r_\varepsilon } + \frac{2}{r_\varepsilon } = \frac{6}{r_\varepsilon } \; \text{ on } [0,r_\varepsilon ]. \end{aligned}$$
(6.17)

Furthermore,

$$\begin{aligned} \frac{g'}{h} \le 2 \; \text{ on } [0,r_\varepsilon ]. \end{aligned}$$
(6.18)

Indeed, for \(r \in [r_\varepsilon / 2, r_\varepsilon ]\) we have \(( \frac{g'}{h} )(r) =0\) by (g1). In 0 we have \(\frac{g'}{h} =0\). Further, for \(r \in (0, \frac{r_\varepsilon }{2}]\) property (g2) shows that \(g'(r) >0\), and hence \(-f' \frac{g}{g'} \ge 0\). Since \(f(r) \ge \frac{1}{2}\) by (f2) and (f3), we conclude that

$$\begin{aligned} \frac{h}{g'} \,=\, f - f' \frac{g}{g'} \,\ge \, \frac{1}{2}, \end{aligned}$$

as claimed.

Given \(\varepsilon >0\) and \(s \in (0,r_\varepsilon /2)\), we define the 1-form

$$\begin{aligned} \alpha _{s,\varepsilon } (\theta ,r,q) :=\, g(r) d\theta + f(r) \sigma _\varepsilon (q) \end{aligned}$$
(6.19)

on \(S^1 \times [0,r_\varepsilon ] \times K\). We note that by (f1) and (g1),

$$\begin{aligned} \alpha _{s,\varepsilon } (\theta ,r,q) \,=\, \frac{r^2}{2} \, d\theta + \sigma _\varepsilon (q) \quad \text{ near } r=0. \end{aligned}$$

By considering \(\theta \) and r as angular and radial coordinates on the disk \(r_\varepsilon \mathbb {D}\), we thus see that the 1-form \(\alpha _{s,\varepsilon }\) is smooth on \( r_\varepsilon \mathbb {D}\times K\).

Lemma 6.6

For \(\varepsilon >0\) and \(s \in (0,r_\varepsilon /2)\), \(\alpha _{s,\varepsilon }\) is a contact form on \(r_\varepsilon \mathbb {D}\times K\).

Proof

We compute

$$\begin{aligned} \alpha _{s,\varepsilon } \wedge (d\alpha _{s,\varepsilon })^n= & {} (g \, d\theta + f \sigma _\varepsilon ) \wedge (g' dr \wedge d\theta + f' dr \wedge \sigma _\varepsilon + f \, d\sigma _\varepsilon )^n \nonumber \\= & {} n \,h\, f^{n-1} \big ( dr \wedge d\theta \wedge \sigma _\varepsilon \wedge (d \sigma _\varepsilon )^{n-1} \big ) \end{aligned}$$
(6.20)

where \(h =fg'-f'g\). Since \(f^{n-1}h>0\) on \((0,r_\varepsilon ]\), it follows that \(\alpha _{s,\varepsilon }\) is a contact form away from K. Near K we have \(h(r)=r\), so that there \(\alpha _{s,\varepsilon } \wedge (d\alpha _{s,\varepsilon })^{n}\) reads

$$\begin{aligned} n \bigl ( r dr \wedge d\theta \wedge \sigma _\varepsilon \wedge (d \sigma _\varepsilon )^{n-1} \bigr ), \end{aligned}$$

which is a positive volume form at any point on K. \(\square \)

Away from K the Reeb vector field is

$$\begin{aligned} R_{\alpha _{s,\varepsilon }}(\theta ,r,q) = -\frac{f'(r)}{h(r)} \, \partial _\theta + \frac{g'(r)}{h(r)} \, R_{\sigma _\varepsilon }(q) \end{aligned}$$

and has the flow

$$\begin{aligned} \phi ^t_{\alpha _{s,\varepsilon }} (\theta ,r,q) = \left( \theta - \frac{f'(r)}{h(r)}\,t,\, r,\, \phi _{\sigma _\varepsilon }^{\frac{g'(r)}{h(r)} \,t}(q)\right) , \end{aligned}$$
(6.21)

where \(\phi _{\sigma _\varepsilon }^t\) is the flow of \(R_{\sigma _\varepsilon }\).

Lemma 6.7

For \(s \in (0,r_{\varepsilon }/2)\) we have \(\Gamma (\alpha _{s,\varepsilon } |_{r_\varepsilon \mathbb {D}\times K}) \le 2 E\).

Proof

By continuity it suffices to estimate the differentials of \(\phi _{\alpha _{s,\varepsilon }}^{\pm n}\) away from K. We choose a basis \(\partial _\theta , \partial _r, \partial _{q_1}, \dots , \partial _{q_{2n-1}}\) of the tangent space at \((\theta ,r,q)\). In view of (6.21), the images of these vectors under \(d \phi ^{\pm n}_{\alpha _{s,\varepsilon }}(\theta ,r,q)\) are

$$\begin{aligned} \partial _\theta , \qquad \mp \biggl ( \frac{f'}{h}\biggr )' \!(r) \,n\, \partial _\theta + \partial _r \pm \biggl ( \frac{g'}{h} \biggr )' \!(r) \,n\, R_{\sigma _\varepsilon }, \qquad d \phi _{\sigma _\varepsilon }^{\pm \bigl ( \frac{g'}{h} \bigr )(r) \, n} (\theta ,r,q) \, \partial _{q_j}. \end{aligned}$$

The size of the functions \(\bigl ( \frac{f'}{h} \bigr )'\) and \(\bigl ( \frac{g'}{h} \bigr )'\) plays no role when we apply \(\displaystyle \lim _{n \rightarrow \infty } \frac{1}{n} \log \), and together with (6.18) we find that

$$\begin{aligned} \Gamma (\alpha _{s,\varepsilon } |_{r_\varepsilon \mathbb {D}\times K}) \,=\, \max _r \left| \biggl ( \frac{g'}{h} \biggr ) (r) \right| \, \Gamma (\sigma _\varepsilon ) \,\le \, 2 \Gamma (\sigma _\varepsilon ). \end{aligned}$$

The lemma follows together with assumption (6.12). \(\square \)

6.3 A family of contact forms on \({\text {OB}}(F,\psi )\)

For every \(\varepsilon >0\) and \(s \in (0,r_\varepsilon /2)\), we define

$$\begin{aligned} \alpha _{s,\varepsilon } = \left\{ \begin{array}{ll} \alpha _{s} &{} \text { on } {\text {MT}}(F_{\varepsilon },\psi ) \\[.2em] \alpha _{s,\varepsilon } = g(r)\, d\theta + f(r) \,\sigma _\varepsilon &{}\text { on } r_\varepsilon \mathbb {D}\times K \end{array} \right. \end{aligned}$$
(6.22)

on the abstract open book

$$\begin{aligned} {\text {OB}}(F,\psi ) = {\text {MT}}(F_{\varepsilon },\psi ) \cup \left( r_\varepsilon \mathbb {D}\times K\right) \end{aligned}$$

where \(\alpha _s\) on \({\text {MT}}(F_{\varepsilon },\psi )\) is defined by (6.4) and f and g are given in Lemma 6.5. By (6.16) and the properties (f1) and (g1), each \(\alpha _{s,\varepsilon }\) is a well-defined contact form on \({\text {OB}}(F,\psi )\).

We first estimate the volume.

$$\begin{aligned} \int _{{\text {OB}}(F,\psi )}\alpha _{s,\varepsilon }\wedge (d\alpha _{s,\varepsilon })^n \,=\, \int _{{\text {MT}}(F_{\varepsilon },\psi )} \alpha _{s} \wedge (d\alpha _{s})^n + \int _{r_\varepsilon \mathbb {D}\times K} \alpha _{s,\varepsilon } \wedge (d\alpha _{s,\varepsilon })^n. \end{aligned}$$

For \(s \in (0,s_1]\), where the positive number \(s_1\) is given by Lemma 6.4, we have

$$\begin{aligned} \int _{{\text {MT}}(F_\varepsilon ,\psi )} \alpha _{s} \wedge (d \alpha _{s})^n \,=\, \int _{F_{\varepsilon }} T_s\, (d\alpha _s|_{\{0\} \times F_\varepsilon })^n \,=\, \int _{F_{\varepsilon }} T_s\, s^n (d\lambda )^n \,\le \, 2 s^n\int _{F_{\varepsilon }}(d\lambda )^n, \end{aligned}$$

where we used (6.9) in the last inequality. For the second term we use (6.20), the assumption (6.12), and \(f \le 1\) and (6.17) to estimate

$$\begin{aligned} \int _{r_\varepsilon \mathbb {D}\times K}\alpha _{s,\varepsilon }\wedge (d\alpha _{s,\varepsilon })^n= & {} \int _{r_\varepsilon \mathbb {D}\times K} n \, h \, f^{n-1} \bigl ( dr \wedge d\theta \wedge \sigma _\varepsilon \wedge (d \sigma _\varepsilon )^{n-1} \bigr ) \\= & {} 2\pi n {\text {vol}}_{\sigma _\varepsilon } (K) \int _0^{r_\varepsilon } h f^{n-1} dr \\\le & {} 12 \pi n \varepsilon . \end{aligned}$$

Together we get

$$\begin{aligned} \int _{{\text {OB}}(F,\psi )} \alpha _{s,\varepsilon } \wedge (d\alpha _{s,\varepsilon })^n \,\le \, 2 s^n\int _{F_{\varepsilon }} (d\lambda )^n + 12 \pi n \varepsilon . \end{aligned}$$
(6.23)

Next we estimate the norm growth \(\Gamma \). The bound (6.10) also applies to the subset \({\text {MT}}(F_\varepsilon ,\psi )\) of \({\text {MT}}(F^\circ ,\psi )\) since \(\Gamma \) is monotone with respect to inclusion of compact invariant subsets. Since \({\text {MT}}(F_\varepsilon ,\psi )\) and \(r_\varepsilon \mathbb {D}\times K\) are invariant under the flow, we conclude together with assertion (3) of Proposition C.1 and Lemma 6.7 that

$$\begin{aligned} \Gamma (\alpha _{s,\varepsilon }) = \max \left\{ \Gamma (\alpha _s |_{{\text {MT}}(F_\varepsilon ,\psi )}), \Gamma ( \alpha _{s,\varepsilon } |_{r_\varepsilon \mathbb {D}\times K}) \right\} \le 2E. \end{aligned}$$
(6.24)

Now given any \(\varepsilon _0 >0\), we choose \(\varepsilon >0\) such that \(12 \pi n \varepsilon \le \varepsilon _0 /2\). Once \(\varepsilon \) is fixed, so are \(r_\varepsilon \) and \(\int _{F_{\varepsilon }}(d\lambda )^n\). If we choose \(s>0\) such that

$$\begin{aligned} s \le \min \left\{ s_1, \frac{r_\varepsilon }{2}, \left( \frac{\varepsilon _0}{4\int _{F_{\varepsilon }}(d\lambda )^n} \right) ^{\frac{1}{n}}\right\} , \end{aligned}$$

then the right-hand side of (6.23) is \(\le \varepsilon _0\). Since \(n! \omega _n \ge 1\) we thus get a contact form \(\alpha _{s,\varepsilon }\) on \({\text {OB}}(F,\psi )\) such that

$$\begin{aligned} \Gamma (\alpha _{s,\varepsilon }) \le 2 E \quad \text{ and } \quad {\text {vol}}_{\alpha _{s,\varepsilon }} \bigl ( {\text {OB}}(F,\psi ) \bigr ) \le \varepsilon _0. \end{aligned}$$
(6.25)

The last step of the proof consists of pushing the contact form \(\alpha _{s,\varepsilon }\) to M in such a way that the contact structure \(\ker \alpha _{s,\varepsilon }\) is mapped to \(\xi \). This is possible thanks to the following lemma.

Lemma 6.8

There exists a diffeomorphism \(\rho \) of M such that \(\rho _* (\xi ) = \ker \alpha _{s,\varepsilon }\).

Postponing the proof, we use the lemma to complete the proof of Theorem 6.1. Thanks to the lemma, the 1-form

$$\begin{aligned} \tau :=\, \left( {\text {vol}}_{\alpha _{s,\varepsilon }} \bigl ( {\text {OB}}(F,\psi ) \bigr ) \right) ^{-1/(n+1)} \, \rho ^* \alpha _{s,\varepsilon } \end{aligned}$$
(6.26)

is a contact form on \((M,\xi )\), and \({\text {vol}}_\tau (M) = 1\). By the elementary properties (1) and (4) in Proposition C.1 and by (6.25),

$$\begin{aligned} {\widehat{\Gamma }} (\tau ) \,=\, \Gamma (\tau ) \,=\, \left( {\text {vol}}_{\alpha _{s,\varepsilon }} \bigl ( {\text {OB}}(F,\psi ) \bigr ) \right) ^{1/ (n+1) } \, \Gamma (\alpha _{s,\varepsilon }) \,\le \, (2E) \, (\varepsilon _0)^{1/ (n+1) }. \end{aligned}$$

Since the constant E from Lemma 6.4 depends only on \(\psi , \lambda , \chi \), which are fixed data associated with \((M,\xi )\), and since \(\varepsilon _0 >0\) is arbitrarily small, we obtain \(\tau \) with \({\widehat{\Gamma }} (\tau )\) as small as we like.

6.4 Proof of Lemma 6.8

We first show that the obvious open book structure on \({\text {OB}}(F,\psi )\) is a Liouville open book with contact binding form \(\alpha _{s,\varepsilon }\). Let

$$\begin{aligned} {\widetilde{\Theta }} :{\text {OB}}(F,\psi )\setminus K \rightarrow S^1 \end{aligned}$$

be the fibration induced by the projection \({\text {MT}}(F,\psi ) \rightarrow S^1\). We construct a defining function \({\tilde{h}}\) as follows. As before, we consider the variable \(r \in [0,+\infty )\) on a neighborhood of \(K=\partial F\) in F that is induced by the embedding (6.13). Let

$$\begin{aligned} {\tilde{u}} :F \rightarrow [0,\infty ) \end{aligned}$$

be a smooth function and \(d>0\) and \(\delta >0\) be constants such that

  1. (df1)

    \({\tilde{u}} (r,q)=r\) for \((r,q) \in [0,r_\varepsilon ] \times K\),

  2. (df2)

    \({\tilde{u}} \equiv d\) on \(([0,r_\varepsilon +\delta ) \times K)^c\)  and  \(\mathrm {supp\,}\psi \subset ([0,r_\varepsilon +\delta ] \times K)^c\),

  3. (df3)

    \(\tilde{u}\) depends only on r and \(\partial _r {\tilde{u}} \ge 0\) on \([0,r_\varepsilon +\delta ] \times K\).

Fig. 10
figure 10

The function \({\tilde{u}}\), schematically

See Fig. 10. Since \({\tilde{u}}\) is constant on \(\mathrm {supp\,}\psi \), the \(S^1\)-invariant extension of \({\tilde{u}}\) is a well-defined smooth function on \({\text {MT}}(F,\psi )\), which constitutes the function \(|{\tilde{h}}|\). Pairing \(|{\tilde{h}}|\) with \({\widetilde{\Theta }}\) leads to a well-defined defining function \({\tilde{h}}\) for the open book \((K,{\widetilde{\Theta }})\) on \({\text {OB}}(F,\psi )\). Note that on \(r_\varepsilon \mathbb {D}\times K\), \({\tilde{h}}\) is simply the projection to the disk \(r_\varepsilon \mathbb {D}\), which is smooth.

Claim 1

\(d \bigl ( \alpha _{s,\varepsilon }/|\tilde{h}| \bigr )\) induces an ideal Liouville structure on each fibre of \(\widetilde{\Theta }\).

Proof

For each \(\theta \in S^1\) we abbreviate

$$\begin{aligned} {\tilde{\lambda }}_\theta := \bigl ( \alpha _{s,\varepsilon }/|\tilde{h}| \bigr ) |_{T(\{\theta \} \times F^\circ )}. \end{aligned}$$
(6.27)

We need to check that \(d{{\tilde{\lambda }}}_\theta \) is a positive symplectic form on \(\{\theta \} \times F^\circ = \widetilde{\Theta }^{-1}(\theta )\). We do this on the three regions separately:

\(\underline{\hbox {On}\, \{\theta \} \times (0,r_\varepsilon ] \times K}\): By (6.22) and (df1) we have

$$\begin{aligned} {\tilde{\lambda }}_\theta = \frac{f(r)}{r} \, \sigma _{\varepsilon } \end{aligned}$$
(6.28)

and so

$$\begin{aligned} (d{\tilde{\lambda }}_\theta )^n = n f^{n-1} \frac{f'r-f}{r^{n+1}}\,dr \wedge \sigma _{\varepsilon } \wedge (d\sigma _{\varepsilon })^{n-1}. \end{aligned}$$

In view of the parametrization (6.13), \(d (\frac{1}{r} \sigma _{\varepsilon })\) is a positive symplectic form on \(\{\theta \} \times F^\circ \), whence \((d(\frac{1}{r} \sigma _{\varepsilon }))^n = - \frac{n}{r^{n+1}} \, dr \wedge \sigma _{\varepsilon } \wedge (d\sigma _{\varepsilon })^{n-1}\) is a positive volume form. Since \(f'r-f<0\) by (f1) and (f3), it follows that \((d {\tilde{\lambda }}_\theta )^n\) is a positive volume form, i.e., \(d {\tilde{\lambda }}_\theta \) is a positive symplectic form.

\(\underline{\hbox {On}\, \{\theta \} \times [r_\varepsilon ,r_\varepsilon +\delta ]\times K}\): By (df2) we have \(\psi = \textrm{id}\) on this set. Hence, by (6.16),

$$\begin{aligned} {\tilde{\lambda }}_\theta = \frac{s}{r {\tilde{u}}} \, \sigma _{\varepsilon }. \end{aligned}$$
(6.29)

By (df3), \({\tilde{u}}\) depends only on r. We thus obtain

$$\begin{aligned} (d{\tilde{\lambda }}_\theta )^n = - s n \frac{\tilde{u} + r \partial _r \tilde{u}}{r^{n+1} {\tilde{u}}^{n+1}} \, dr \wedge \sigma _{\varepsilon } \wedge (d\sigma _{\varepsilon })^{n-1}. \end{aligned}$$

Also by (df3), \(\tilde{u} + r \partial _r \tilde{u}>0\), and the claim follows as in the previous case.

\(\underline{\hbox {On}\, \{\theta \} \times ([0,r_\varepsilon +\delta )\times K)^c}\): By (df2), \({\tilde{u}} \equiv d\) and

$$\begin{aligned} {\tilde{\lambda }}_\theta = \frac{s}{d} \left( \lambda + \chi (\theta ) \lambda _\psi \right) . \end{aligned}$$
(6.30)

Hence \(d {\tilde{\lambda }}_\theta =\frac{s}{d} d\lambda \), which is a positive symplectic form. \(\square \)

Now we are in the following situation. On \({\text {OB}}(F,\psi )\) we have the Liouville open book

$$\begin{aligned} \left( K,\widetilde{\Theta }, d(\alpha /|h|)|_{T(\{\theta \} \times F^\circ )} \right) , \end{aligned}$$
(6.31)

which symplectically supports the contact structure \(\xi = \ker \alpha \). Here \(\alpha \), \(\xi \), and h stand for the objects induced by the correspondence between M and \({\text {OB}}(F,\psi )\) given by the symplectically spinning vector field X on M. Moreover, by Claim 1 we have a second Liouville open book

$$\begin{aligned} \left( K, \widetilde{\Theta }, d \bigl ( \alpha _{s,\varepsilon }/|\tilde{h}| \bigr )|_{T(\{\theta \} \times F^\circ )} \right) , \end{aligned}$$
(6.32)

which symplectically supports the contact structure \(\ker \alpha _{s,\varepsilon }\). By the identities (6.28), (6.29), and (6.30), the ideal Liouville structures \(\bigl ( d \bigl ( \alpha _{s,\varepsilon }/|\tilde{h}| \bigr ) |_{T(\{\theta \} \times F^\circ )} \bigr )_{\theta \in S^1}\) are invariant under the flow of the vector field \(\partial _\theta \). Although \({{\tilde{\lambda }}}_\theta \) in (6.30) is not invariant under this flow, the symplectic form \(d {{\tilde{\lambda }}}_\theta = \frac{s}{d} \, d\lambda \) is. The vector field \(\partial _\theta \) is therefore a symplectically spinning vector field on the LOB (6.32). Note that the symplectically spinning vector field X on M also reads \(\partial _\theta \) on the LOB (6.31).

Claim 2

There exists a diffeomorphism

$$\begin{aligned} \Phi :{\text {OB}}(F,\psi ) \rightarrow {\text {OB}}(F,\psi ) \end{aligned}$$
(6.33)

such that \(\Phi \circ {\widetilde{\Theta }} = {\widetilde{\Theta }} \circ \Phi \) and the restriction of \(\Phi \) to each fibre is symplectic, that is,

$$\begin{aligned} \Phi ^*d \bigl ( \alpha _{s,\varepsilon }/|\tilde{h}| \bigr )|_{T(\{\theta \} \times F^\circ )} = d \bigl ( \alpha /|h| \bigr )_{T(\{\theta \} \times F^\circ )}, \quad \forall \,\theta \in S^1. \end{aligned}$$

If such a diffeomorphism exists, then \(\ker \Phi ^*\alpha _{s,\varepsilon }\) and \(\ker \alpha \) are two contact structures on \({\text {OB}}(F,\psi )\) which symplectically support the Liouville open book (6.31). Hence they are isotopic by Proposition 5.5. By Gray’s stability theorem we then find a diffeomorphism \({\hat{\rho }}\) of M such that \({\hat{\rho }}_* (\ker \alpha ) = \ker \Phi ^* \alpha _{s,\varepsilon }\). Set \(\rho = \Phi \circ {\hat{\rho }}\). Since \(\ker \Phi ^* \alpha _{s,\varepsilon } = \Phi _*^{-1} (\ker \alpha _{s,\varepsilon })\), we conclude that \(\rho _* (\ker \alpha ) = \ker \alpha _{s,\varepsilon }\), as claimed in Lemma 6.8.

Proof of Claim 2

We have the following ideal Liouville structures on the 0-page:

$$\begin{aligned} {\widetilde{\omega }}&:= d \bigl ( \alpha _{s,\varepsilon }/|\tilde{h}| \bigr )|_{T(\{0\} \times F^\circ )}, \end{aligned}$$
(6.34)
$$\begin{aligned} \omega&:= d \bigl ( \alpha /|h| \bigr )|_{T(\{0\}\times F^\circ )}=d\lambda . \end{aligned}$$
(6.35)

We first show that

$$\begin{aligned} \omega _t:= (1-t) \omega + t {\widetilde{\omega }} \end{aligned}$$

is symplectic on \(F^\circ \) for all \(t\in [0,1]\). In fact, we claim that

$$\begin{aligned} \lambda _t = (1-t) \lambda + t {\tilde{\lambda }} \end{aligned}$$
(6.36)

is a Liouville form on \(F^\circ \) for all \(t \in [0,1]\), where \(\lambda \) is the primitive of \(\omega \) given by (6.3) and \({{\tilde{\lambda }}}\) is the primitive of \({\widetilde{\omega }}\) given by (6.27). Again, we compute \(d\lambda _t\) on different subsets of \(F^\circ \):

\(\underline{\hbox {On}\, \{\theta \} \times (0,r_\varepsilon ]\times K}\): By (6.28) we have

$$\begin{aligned} \lambda _t = (1-t) \,\frac{1}{r}\, \sigma _\varepsilon + t \,\frac{f(r)}{r}\, \sigma _\varepsilon = \frac{\kappa (r)}{r} \,\sigma _\varepsilon \end{aligned}$$

where \(\kappa =(1-t)+tf\). We have \(\kappa >0\) and \(\kappa '<0\), so that \(\kappa 'r-\kappa <0\). Hence the claim follows as in the first case of Claim 1.

\(\underline{\hbox {On}\, \{\theta \}\times [r_\varepsilon ,r_\varepsilon +\delta ]\times K}\): By (6.29) we have

$$\begin{aligned} \lambda _t = (1-t) \,\frac{1}{r}\, \sigma _\varepsilon + t \,\frac{s}{r {\tilde{u}}}\, \sigma _\varepsilon = \frac{\kappa (r)}{r} \, \sigma _\varepsilon \end{aligned}$$

where \(\kappa = (1-t) + t s/ {\tilde{u}}\). We have \(\kappa >0\) and \(\kappa ' \le 0\), so that \(\kappa 'r-\kappa <0\). The claim follows as above.

\(\underline{\hbox {On}\, \{\theta \} \times ([0,r_\varepsilon +\delta )\times K)^c}\): By (6.30) we have

$$\begin{aligned} d \lambda _t = (1-t) d\lambda + t \,\frac{s}{d}\, d\lambda = \left( (1-t) + t \frac{s}{d} \right) \, d \lambda . \end{aligned}$$

Hence \(\omega _t = d\lambda _t\) is symplectic on \(F^\circ \) for all \(t \in [0,1]\).

Recall that by (6.13) and (f1),

$$\begin{aligned} \omega = \widetilde{\omega } = d \left( \frac{1}{r} \,\sigma _\varepsilon \right) \quad \text{ on } \text{ a } \text{ deleted } \text{ neighbourhood } \text{ of } K. \end{aligned}$$

Hence this identity holds on the same deleted neighbourhood of K for all the symplectic forms \(\omega _t\), \(t \in [0,1]\). Applying the standard Moser argument to the path \(\omega _t\), we obtain a smooth isotopy \((\psi _t)_{t\in [0,1]}\) of F such that

(\(\Psi \)1):

\(\psi _0 = \textrm{id}\);

(\(\Psi \)2):

\(\psi _t = \textrm{id}\) near K for all \(t \in [0,1]\);

(\(\Psi \)3):

\(\psi _t^* \, \omega _t = \omega _0 = \omega \) for all \(t \in [0,1]\).

Now we define \(\Phi :[0,2\pi ] \times F \rightarrow [0,2\pi ] \times F\) by

$$\begin{aligned} \Phi (\theta ,p):= \left( \theta ,\psi _{1} \circ \psi ^{-1}_{\frac{\theta }{2\pi }} \circ \psi ^{-1} \circ \psi _{\frac{\theta }{2\pi }} (p) \right) \end{aligned}$$
(6.37)

where \(\psi \) is the monodromy that we fixed at the outset of the proof. We note that

$$\begin{aligned} \Phi (2\pi ,p) = \bigl (2\pi , \psi ^{-1} \circ \psi _{1}(p) \bigr ), \end{aligned}$$

and by (\(\Psi \)1),

$$\begin{aligned} \Phi (0,\psi (p)) = \left( 0, \psi _{1}(p) \right) = \bigl ( 0, \psi (\psi ^{-1} \circ \psi _{1}(p)) \bigr ). \end{aligned}$$

Hence \(\Phi \) descends to a diffeomorphism on \({\text {MT}}(F,\psi )\). Since \(\psi = \textrm{id}\) near K and \(\psi _t = \textrm{id}\) near K for each t by (\(\Psi \)2), we have that \(\Phi = \textrm{id}\) on a neighbourhood of \(\partial {\text {MT}}(F,\psi )\). Hence \(\Phi \) descends to a diffeomorphism on \({\text {OB}}(F,\psi )\). By definition, \(\Phi \) commutes with \(\widetilde{\Theta }\).

Now recall that \(\partial _{\theta }\) is a symplectically spinning vector field for both LOBs (6.31) and (6.32). In view of (6.34) and (6.35) and identifying \(\{\theta \} \times F^\circ \) with \(\{0\} \times F^\circ \) via the flow of \(\partial _\theta \), we can therefore identify

$$\begin{aligned} d \bigl ( \alpha / |h| \bigr ) |_{T (\{\theta \} \times F^\circ )}&\text{ with }&\omega |_{T (\{\theta \} \times F^\circ )}:= \omega , \\ d \bigl ( \alpha _{s,\varepsilon } / |{\tilde{h}}| \bigr ) |_{T (\{\theta \} \times F^\circ )}&\text{ with }&{\widetilde{\omega }} |_{T (\{\theta \} \times F^\circ )}:= {\widetilde{\omega }}. \end{aligned}$$

Also recall that \(\psi ^* \omega = \omega \). Since \(\partial _\theta \) generates the monodromy \(\psi \) and preserves \({\widetilde{\omega }}\), we also have \(\psi ^* {\widetilde{\omega }} = {\widetilde{\omega }}\). Therefore, \(\psi ^*\omega _t = \omega _t\) for all \(t \in [0,1]\). Inserting (6.37) and using (\(\Psi \)3) we obtain, with the abbreviation \(F_\theta ^\circ = T (\{\theta \} \times F^\circ )\),

$$\begin{aligned} \Phi ^* d \bigl ( \alpha _{s,\varepsilon }/|\tilde{h}| \bigr ) |_{F_\theta ^\circ }= & {} \Phi ^* {\widetilde{\omega }} |_{F_\theta ^\circ } \\= & {} \big (\psi _{1}\circ \psi ^{-1}_{\frac{\theta }{2\pi }} \circ \psi ^{-1}\circ \psi _{\frac{\theta }{2\pi }} \big )^* {\widetilde{\omega }} |_{F_\theta ^\circ } \\= & {} \psi _{\frac{\theta }{2\pi }}^* \, (\psi ^{-1})^* \, (\psi ^{-1}_{\frac{\theta }{2\pi }})^* \, \psi _{1}^* \omega _{1} |_{F_\theta ^\circ } \\= & {} \psi _{\frac{\theta }{2\pi }}^* \, (\psi ^{-1})^* \, (\psi ^{-1}_{\frac{\theta }{2\pi }} )^* \, \omega _0 |_{F_\theta ^\circ } \\= & {} \psi _{\frac{\theta }{2\pi }}^* \, (\psi ^{-1})^* \omega _{\frac{\theta }{2\pi }} |_{F_\theta ^\circ } \\= & {} \psi _{\frac{\theta }{2\pi }}^* \, \omega _{\frac{\theta }{2\pi }} |_{F_\theta ^\circ } \\= & {} \omega _0 |_{F_\theta ^\circ } = \omega |_{F_\theta ^\circ } = d(\alpha /|h|) |_{F_\theta ^\circ }. \end{aligned}$$

This concludes the proof of Lemma 6.8, and hence of Theorem 6.1.

7 Full entropy spectrum

By Theorem 6.1 every closed contact manifold \((M,\xi )\) admits normalized contact forms with arbitrarily small topological entropy. On the other hand, one can always find normalized contact forms \(\alpha \) on \((M,\xi )\) with arbitrarily large topological entropy. We first sketch a direct proof of this fact. A proof of a stronger statement relying on our previous construction is given thereon.

Existence of contact forms with large entropy. Choose a transverse knot \(\gamma \), that is, a simple closed curve \(\gamma :S^1 \rightarrow M\) such that \({\dot{\gamma }}\) is everywhere transverse to \(\xi \). Let \(B^{2n}\) be the closed ball in \(\mathbb {R}^{2n}\) of radius 1. By the normal form theorem for transverse knots [52, Example 2.5.16], we find a full torus \({{\mathcal {T}}}= B^{2n} \times S^1\) around \(\gamma \) with coordinates \((\theta _1, \dots , \theta _n, r_1, \dots , r_n, q)\) such that \(\gamma \) is parametrized by \(r_j = 0\) and q and such that on \({{\mathcal {T}}}\) the contact structure \(\xi \) is the kernel of \(\tau := d q + \sum _j r_j d \theta _j\).

Next, take a second such full torus \(({{\mathcal {T}}}_2, \tau _2)\) that is disjoint from \({{\mathcal {T}}}\), perturb \(\tau _2\) in the interior of \({{\mathcal {T}}}_2\) to a contact form \(\tau _2'\) with positive topological entropy, and take a contact form \(\alpha \) on \((M,\xi )\) that agrees with \(\tau \) on \({{\mathcal {T}}}\) and with \(\tau _2'\) on \({{\mathcal {T}}}_2\). Now for \(\delta >0\) let \(F :M \rightarrow \mathbb {R}\) be a positive smooth function that is equal to \(\delta \) outside \({{\mathcal {T}}}\), on \({{\mathcal {T}}}\) depends only on the coordinates \(r_1, \dots , r_n\), and is such that \({\text {vol}}_{F\tau } ({{\mathcal {T}}}) = 1\). Take the smooth contact form \(\alpha _{\delta , F}\) on M that is equal to \(F\tau \) on \({{\mathcal {T}}}\) and to \(\delta \alpha \) on \(M {\setminus } {{\mathcal {T}}}\). Then \(h_{{\text {top}}} (\alpha _{\delta , F}|_{{\mathcal {T}}}) = 0\) whence

$$\begin{aligned} h_{{\text {top}}}(\alpha _{\delta , F}) \,=\, h_{{\text {top}}}(\alpha _{\delta , F} |_{M\setminus {{\mathcal {T}}}}) \,=\, h_{{\text {top}}}(\delta \alpha |_{M\setminus {{\mathcal {T}}}}) \,=\, \delta ^{-1} \, h_{{\text {top}}}(\alpha ). \end{aligned}$$

With this one then readily finds

$$\begin{aligned} \left( {\widehat{h}}_{{\text {top}}}(\alpha _{\delta , F}) \right) ^{n+1} \,=\, \left( {\widehat{h}}_{{\text {top}}}(\alpha |_{M \setminus {{\mathcal {T}}}}) \right) ^{n+1} + \left( \frac{h_{{\text {top}}}(\alpha )}{\delta } \right) ^{n+1}. \end{aligned}$$

Varying \(\delta \) on \((0,\infty )\) we obtain a normalized contact form with topological entropy c for every \(c > {\widehat{h}}_{{\text {top}}}(\alpha |_{M {\setminus } {{\mathcal {T}}}})\). \(\square \)

Remark 7.1

For spherizations \(S^*Q_k\) of closed orientable surfaces of genus \(k \ge 2\) this result has been obtained in [45] inside the much smaller class of geodesic flows of negatively curved Riemannian metrics: For every \(c \ge 2\pi \sqrt{2(k-1)}\) there exists a negatively curved Riemannian metric g on \(Q_k\) such that \({\widehat{h}}_{{\text {top}}}(g) =c\). In the class of all Riemannian metrics, geodesic flows with arbitrarily large \({\widehat{h}}_{{\text {top}}}\) were constructed on all closed manifolds of dimension at least two already in [70].

We shall now combine the above argument with the construction in the proof of Theorem 6.1 to prove the following more precise result.

Proposition 7.2

Let \((M,\xi )\) be a closed co-orientable contact manifold of dimension \(2n+1 \ge 3\). Then for every \(c>0\) there exist normalized contact forms \(\alpha \) and \(\alpha '\) on \((M,\xi )\) such that \(h_{{\text {top}}}(\alpha ) = c\) and \(\Gamma (\alpha ') =c\).

Proof

We give the proof for the topological entropy. The proof for the norm growth \(\Gamma \) is similar. Fix \(c>0\) as in the proposition. As in (6.26) let

$$\begin{aligned} \tau \,=\, v^{-1/(n+1)} \, \rho ^* \alpha _{s,\varepsilon } \end{aligned}$$

be a contact form with \(h_{{\text {top}}}(\tau ) \le c\) and \({\text {vol}}_\tau (M) = 1\). Here we abbreviated \(v:= {\text {vol}}_{\alpha _{s,\varepsilon }} \bigl ( {\text {OB}}(F,\psi ) \bigr )\). In view of (6.25) we can assume that

$$\begin{aligned} v \le 1. \end{aligned}$$
(7.1)

Recall that \(\alpha _{s,\varepsilon }\) on \(r_\varepsilon \mathbb {D}\times K\) was constructed recursively, starting from the circle \(S^1\). For \(j=1, \dots , n\) let \(f_j,g_j,h_j\) be the functions from Lemma 6.5 that we used in (6.19) to construct \(\alpha _{s,\varepsilon _j}\) on \(r_{\varepsilon _j} \mathbb {D}\times K_j\). Set \(r = \min \{ r_{\varepsilon _1}, \dots , r_{\varepsilon _n} \} >0\). Let \(B^{2n}(r)\) be the closed r-ball in \(\mathbb {R}^{2n}\) with polar coordinates \((\varvec{\uptheta }, {{\textbf {r}}}) \equiv (\theta _1, \dots , \theta _n, r_1, \dots , r_n)\), and abbreviate \({{\mathcal {T}}}_r = B^{2n}(r) \times S^1\).

For \(\delta >0\) choose a positive smooth function \(F :{\text {OB}}(F,\psi ) \rightarrow \mathbb {R}\) with the following properties:

  1. (F1)

    \(F = \delta \) on \({{\mathcal {T}}}_r^c:= {\text {OB}}(F,\psi ) {\setminus } {{\mathcal {T}}}_r\).

  2. (F2)

    F only depends on \({{\textbf {r}}}\) on \({{\mathcal {T}}}_r\).

  3. (F3)

    \({\text {vol}}_{F \alpha _{s,\varepsilon }} ({{\mathcal {T}}}_r) = 1\).

Lemma 7.3

\(h_{{\text {top}}} ( F \alpha _{s,\varepsilon } |_{{{\mathcal {T}}}_r}) =0\).

Proof

The full torus \({{\mathcal {T}}}_r\) is foliated by the tori

$$\begin{aligned} \mathbb {T}_{{{\textbf {r}}}} :=\, (S^1)^k \times \left\{ {{\textbf {r}}}= (r_1, \dots , r_n) \right\} \times S^1 \end{aligned}$$

with \(r_1, \dots , r_n \ge 0\) constant, of dimension \(k+1 \le n+1\). By (f1) and (g1),

$$\begin{aligned} \frac{f_j'(r_j)}{h_j(r_j)} = -2r_j \quad \text{ and } \quad \frac{g_j'(r_j)}{h_j(r_j)} = 1 \quad \text{ for } r_j \text{ near } 0. \end{aligned}$$
(7.2)

Writing out (6.21) recursively and using (7.2) we see that the Reeb flow of \(\alpha _{s,\varepsilon }\) leaves the tori \(\mathbb {T}_{{{\textbf {r}}}}\) invariant and on each \(\mathbb {T}_{{{\textbf {r}}}}\) is a Kronecker flow.

Applying now (6.19) recursively we see that on \({{\mathcal {T}}}_r\),

$$\begin{aligned} \alpha _{s,\varepsilon } (\varvec{\uptheta }, {{\textbf {r}}}, q) \,=\, \rho _1({{\textbf {r}}}) d\theta _1 + \dots + \rho _n({{\textbf {r}}}) d\theta _n +\rho ({{\textbf {r}}}) dq \end{aligned}$$

with smooth functions \(\rho _j, \rho \). The Reeb flow of any 1-form of this form leaves the tori \(\mathbb {T}_{{\textbf {r}}}\) invariant and there restricts to a Kronecker flow. This is clear at \({{\textbf {r}}}\) if the Jacobian determinant of \(\left( \frac{\partial \rho _i}{\partial r_j} ({{\textbf {r}}}) \right) \) does not vanish, and in general follows by approximation. Therefore, for each \({{\textbf {r}}}\) the Reeb flow of \(F \alpha _{s,\varepsilon }\) is a Kronecker flow on \(\mathbb {T}_{{\textbf {r}}}\), and hence \(h_{{\text {top}}} (F \alpha _{s,\varepsilon } |_{\mathbb {T}_{{{\textbf {r}}}}}) =0\). The variational principle for topological entropy now implies that

$$\begin{aligned} h_{{\text {top}}} (F \alpha _{s,\varepsilon }) \,=\, \sup _{{{\textbf {r}}}} h_{{\text {top}}} (F \alpha _{s,\varepsilon } |_{\mathbb {T}_{{{\textbf {r}}}}}) = 0, \end{aligned}$$

as claimed. \(\square \)

By Lemma 7.3 the topological entropy of \(\alpha _{s,\varepsilon }\) and \(F \alpha _{s,\varepsilon }\) on \({{\mathcal {T}}}_r\) vanishes. Together with (F1) we obtain

$$\begin{aligned} h_{{\text {top}}}(F \alpha _{s,\varepsilon }) \,=\, h_{{\text {top}}}(F \alpha _{s,\varepsilon } |_{{{\mathcal {T}}}_r^c}) \,=\, \tfrac{1}{\delta } \, h_{{\text {top}}}(\alpha _{s,\varepsilon } |_{{{\mathcal {T}}}_r^c}) \,=\, \tfrac{1}{\delta } \, h_{{\text {top}}}(\alpha _{s,\varepsilon }). \end{aligned}$$
(7.3)

Now consider the contact form

$$\begin{aligned} (F \circ \rho ) \tau \,=\, v^{-1/(n+1)} \, \rho ^* (F \alpha _{s,\varepsilon }) \end{aligned}$$
(7.4)

on \((M,\xi )\). By (7.4) and (7.3),

$$\begin{aligned} {\widehat{h}}_{{\text {top}}} \bigl ( (F \circ \rho ) \tau \bigr )^{n+1}= & {} {\text {vol}}_{(F \circ \rho ) \tau } (M) \; h_{{\text {top}}} \bigl ( (F \circ \rho ) \tau \bigr )^{n+1} \\= & {} v^{-1} {\text {vol}}_{F \alpha _{s,\varepsilon }} \bigl ( {\text {OB}}(F,\psi ) \bigr ) \; v \, h_{{\text {top}}} (F \alpha _{s,\varepsilon })^{n+1} \\= & {} {\text {vol}}_{F \alpha _{s,\varepsilon }} \bigl ( {\text {OB}}(F,\psi ) \bigr ) \, \delta ^{-(n+1)} \, h_{{\text {top}}} (\alpha _{s,\varepsilon })^{n+1}. \end{aligned}$$

By (7.1), \({\underline{v}}:= {\text {vol}}_{\alpha _{s,\varepsilon }} ({{\mathcal {T}}}_r^c) \in (0,1)\), and by (F1) and (F3),

$$\begin{aligned} {\text {vol}}_{F\alpha _{s,\varepsilon }} \bigl ( {\text {OB}}(F,\psi ) \bigr ) \,=\, 1+\delta ^{n+1} {\underline{v}}, \end{aligned}$$

whence

$$\begin{aligned} {\widehat{h}}_{{\text {top}}} \bigl ( (F \circ \rho ) \tau \bigr )^{n+1} \,=\, \left( {\underline{v}} + \delta ^{-(n+1)} \right) h_{{\text {top}}} (\alpha _{s,\varepsilon })^{n+1} \,=:\, (f(\delta ))^{n+1}. \end{aligned}$$

Assume first that \(h_{{\text {top}}} (\alpha _{s,\varepsilon }) >0\). Then the range of the function \(f :(0,\infty ) \rightarrow \mathbb {R}\) is \(\left( {\underline{v}}^{1/(n+1)} \, h_{{\text {top}}} (\alpha _{s,\varepsilon }), \infty \right) \). Since \({\underline{v}}<1\) and \(h_{{\text {top}}} (\alpha _{s,\varepsilon }) \le c\), we in particular find \(\delta \) such that \({\widehat{h}}_{{\text {top}}} \bigl ( (F \circ \rho ) \tau \bigr ) =c\). If \(h_{{\text {top}}} (\alpha _{s,\varepsilon }) =0\), Proposition 7.2 follows from the following result.

Lemma 7.4

We can assume that \(h_{{\text {top}}}(\alpha _{s,\varepsilon } |_{{\text {MT}}(F_\varepsilon , \psi )}) >0\).

Proof

Assume that \(h_{{\text {top}}}(\alpha _{s,\varepsilon } |_{ {\text {MT}}(F_\varepsilon , \psi )}) =0\). By Theorem 6.2 in [76] there exists a contact form \(\alpha _{s,\varepsilon }'\) on \({\text {MT}}(F_\varepsilon ,\psi )\) that is \(C^1\)-close to \(\alpha _{s,\varepsilon }\) and equal to \(\alpha _{s,\varepsilon }\) near the boundary, and whose Reeb flow has a generic 1-elliptic periodic orbit. Hence this flow contains a hyperbolic basic set, and therefore \(h_{{\text {top}}} (\alpha _{s,\varepsilon }') >0\). Further, the \(C^1\)-closeness of \(\alpha _{s,\varepsilon }\) and \(\alpha _{s,\varepsilon }'\) implies that all the 1-forms

$$\begin{aligned} (1-s) \alpha _{s,\varepsilon } + s \alpha _{s,\varepsilon }', \quad s \in [0,1], \end{aligned}$$

are contact forms. Gray’s stability theorem therefore shows that there exists a diffeomorphism \(\zeta \) of \({\text {MT}}(F_\varepsilon , \psi )\) that is the identity near the boundary such that the kernel of \(\zeta ^* \alpha _{s,\varepsilon }'\) is \(\xi \). The contact form on \((M,\xi )\) that agrees with \(\alpha _{s,\varepsilon }\) on \(M {\setminus } {\text {MT}}(F_\varepsilon , \psi )\) and with \(\zeta ^* \alpha _{s,\varepsilon }'\) on \({\text {MT}}(F_\varepsilon , \psi )\) is the contact form we were looking for.

Newhouse’s full Theorem 6.2 starts with the \(C^1\)-closing lemma, and holds in all dimensions. We do not need to appeal to the closing lemma in our situation, and we only need the easier 3-dimensional result: Assume that \(\dim M = 3\). By (6.8) we can choose a flow-invariant neighbourhood \(U \subset {\text {MT}}(F_\varepsilon , \psi )\) of the boundary of \({\text {MT}}(F_\varepsilon , \psi )\) such that all orbits in U are closed. Let \(\gamma \) be one of these orbits that is not on the boundary of \({\text {MT}}(F_\varepsilon ,\psi )\), and choose a flow-invariant open neighbourhood \(N(\gamma )\) whose closure is also disjoint from the boundary of \({\text {MT}}(F_\varepsilon ,\psi )\). Since \(N(\gamma )\) is foliated by closed orbits, \(\gamma \) is elliptic. Using the Birkhoff normal form theorem and the KAM theorem, one can \(C^1\)-perturb \(\alpha _{s,\varepsilon }\) to a contact form \(\alpha _{s,\varepsilon }'\) that agrees with \(\alpha _{s,\varepsilon }\) outside \(N(\gamma )\) and whose Reeb flow has a transverse homoclinic connection near \(\gamma \), see [97]. This Reeb flow therefore contains a horse-shoe and thus has positive topological entropy. As above we can isotope \(\alpha _{s,\varepsilon }'\) without changing it outside \(N(\gamma )\) to a contact form \(\alpha _{s,\varepsilon }''\) for \((M^3,\xi )\). Proceeding from this contact form with positive topological entropy, our inductive construction in Sect. 6.2 shows that \(\sigma _\varepsilon \) on \(M^{2n+1}\) also has positive topological entropy. Hence this also holds true for \(\alpha _{s,\varepsilon }\) on the boundary of \({\text {MT}}(F_\varepsilon , \psi )\) and therefore, by the monotonicity of topological entropy, also on \({\text {MT}}(F_\varepsilon , \psi )\). \(\square \)

8 Collapsing the growth rate of symplectic invariants

Theorem 1.1 on the collapse of topological entropy of Reeb flows implies the collapse of the growth rate of two symplectic invariants: symplectic homology and wrapped Floer homology.

8.1 Liouville domains and fillings

Recall that a Liouville domain is a compact exact symplectic manifold \({\varvec{W}}=(W,\varvec{\omega },\varvec{\lambda })\) with boundary \(\Sigma = \partial W\) and a primitive \(\varvec{\lambda }\) of \(\varvec{\omega }\) such that \(\varvec{\alpha }_W = \varvec{\lambda }|_{\Sigma }\) is a contact form on \(\Sigma \). The Liouville 1-form \(\varvec{\lambda }\) also induces the contact structure \(\xi _W = \ker \varvec{\alpha }_W\) on \(\Sigma \). The Liouville domain \({\varvec{W}}\) is called an exact symplectic filling for the contact form \(\varvec{\alpha }_W\), and we say that the contact form \(\varvec{\alpha }_W\) is exactly filled by \({\varvec{W}}\).

A standard construction (see for example [11, Section 2.2.1]) shows that if a contact form \(\alpha \) on a contact manifold \((\Sigma ,\xi )\) is exactly filled by a Liouville domain \({\varvec{W}}_{\!\!\, \alpha }\), then we can construct for any other contact form \(\alpha '\) on \((\Sigma ,\xi )\) a Liouville domain \({\varvec{W}}_{\!\!\, \alpha '}\) that fills \(\alpha '\). It therefore makes sense to say that a contact manifold is fillable by Liouville domains.

8.2 Symplectic homology and collapse of its exponential growth

Let \({\varvec{W}}_{\!\!\, \alpha }\) be a Liouville domain filling the contact form \(\alpha \) on the contact manifold \((\Sigma ,\xi )\). The symplectic homology \({\text {SH}}({\varvec{W}}_{\!\!\, \alpha })\) is a homology theory associated to \({\varvec{W}}_{\!\!\, \alpha }\). While there are various versions of symplectic homology, we here consider the one originally developed by Viterbo [94].

Geometrically, one can think of the chain complex associated to \({\text {SH}}({\varvec{W}}_{\!\!\, \alpha })\) as the \(\mathbb {Z}_2\)-vector space generated by the periodic orbits of the Reeb flow of \(\alpha \) and by the critical points of a \(C^2\)-small non-positive Morse–Smale function \(f :{\varvec{W}}_{\!\!\, \alpha } \rightarrow \mathbb {R}\) such that \(f^{-1}(0) = \Sigma \) is a regular energy level. The differential of \({\text {SH}}({\varvec{W}}_{\!\!\, \alpha })\) counts Floer cylinders connecting generators. We refer the reader to [24, 78] for details.

There is a filtration of \({\text {SH}}({\varvec{W}}_{\!\!\, \alpha })\) by the action of its generators. For the Reeb orbits the action equals the period. For each real number \(a>0\) let \({\text {SH}}^{<a}({\varvec{W}}_{\!\!\, \alpha })\) be the homology of the subcomplex generated by the Reeb orbits of action \(<a\) and the critical points of f. The inclusion of this subcomplex induces the homomorphism of \(\mathbb {Z}_2\)-vector spaces

$$\begin{aligned} \Psi ^a_{\alpha } :{\text {SH}}^{<a} ({\varvec{W}}_{\!\!\, \alpha }) \rightarrow {\text {SH}}({\varvec{W}}_{\!\!\, \alpha }). \end{aligned}$$

We define the exponential growth rate of \({\text {SH}}({\varvec{W}}_{\!\!\, \alpha })\) by

$$\begin{aligned} \Gamma ( {\text {SH}}({\varvec{W}}_{\!\!\, \alpha })):= \limsup _{a \rightarrow +\infty } \frac{\log \left( \mathrm {rank\,}(\Psi ^a_{\alpha }) \right) }{a}. \end{aligned}$$

The following remarkable result is due to Meiwes [74].

Theorem

If \({\varvec{W}}_{\!\!\, \alpha }\) is a Liouville domain filling a contact form \(\alpha \) on \((\Sigma ,\xi )\), then

$$\begin{aligned} h_{{\text {top}}}(\phi _\alpha ) \,\ge \, \Gamma ( {\text {SH}}({\varvec{W}}_{\!\!\, \alpha })). \end{aligned}$$

Together with Theorem 1.1 we obtain the following result.

Corollary 8.1

Let \((\Sigma ,\xi )\) be a contact manifold fillable by Liouville domains. Then for every \(\varepsilon >0 \) there exists a contact form \(\alpha \) with \({\text {vol}}_\alpha (\Sigma ) = 1\) such that

$$\begin{aligned} \Gamma ( {\text {SH}}( {\varvec{W}}_{\!\!\, \alpha })) \le \varepsilon , \end{aligned}$$

for any Liouville filling \({\varvec{W}}_{\!\!\, \alpha }\) of \(\alpha \).

It follows that one cannot, in general, recover the volume of a contact form \(\alpha \) from the exponential growth rate of \({\text {SH}}({\varvec{W}}_{\!\!\, \alpha })\) of a Liouville filling \({\varvec{W}}_{\!\!\, \alpha }\). To obtain a better geometric formulation of the corollary, we notice that if \({\varvec{W}}_{\!\!\, \alpha }=(W^{2n}_\alpha , \varvec{\omega }_{\alpha }, \varvec{\lambda }_\alpha )\) is a Liouville filling of \(\alpha \), then the symplectic volume \( \int _{W_{\!\!\, \alpha }} (\varvec{\omega }_\alpha )^n \) equals the contact volume of \({\text {vol}}_\alpha (\Sigma )\). Corollary 8.1 thus says that every Liouville fillable contact manifold admits fillings by Liouville domains of symplectic volume 1 and arbitrarily small growth of symplectic homology.

In the opposite direction, one can ask if for a fixed contact manifold \((\Sigma ,\xi )\) there exists a constant \(\textrm{K}_{\Sigma ,\xi }\) such that

$$\begin{aligned} \Gamma ( {\text {SH}}({\varvec{W}})) \,\le \, \textrm{K}_{\Sigma ,\xi } \end{aligned}$$

for every Liouville domain \({\varvec{W}}\) that fills some normalized contact form on \((\Sigma ,\xi )\). A partial negative answer to this question is given by the following result. Recall that the spherization of a closed manifold Q is the contact manifold \((S^*Q,\xi _{\textrm{can}\,})\) whose Reeb flows comprise the co-geodesic flows of Riemannian metrics on Q.

Lemma 8.2

Let \(Q_k\) be the closed orientable surface of genus \(k \ge 2\). Then for every real number \(c \ge 2 \pi \sqrt{2(k-1)}\) there exists a contact form \(\alpha \) of volume 1 on \((S^*Q_k, \xi _{\textrm{can}\,})\) and a Liouville domain \({\varvec{W}}_{\!\!\, \alpha }\) filling \(\alpha \) such that

$$\begin{aligned} \Gamma ( {\text {SH}}({\varvec{W}}_{\!\!\, \alpha })) = c. \end{aligned}$$

Proof

It follows from [45, Theorem A] that for any real number \(c \ge 2 \pi \sqrt{2(k-1)}\) there exists a negatively curved Riemannian metric g with area \(1/(2\pi )\) such that the topological entropy of the geodesic flow \(\phi _g\) is equal to c. Let \(\alpha \) be the contact form on \((S^*Q_k,\xi _{\textrm{can}\,})\) whose Reeb flow is the co-geodesic flow of g. Then \({\text {vol}}_\alpha (S^*Q_k) = 1\) and \(h_{{\text {top}}} (\phi _\alpha ) = h_{{\text {top}}}(\phi _g) = c\).

Let \({\varvec{W}}_{\!\!\, \alpha } = D^*(g) \subset (T^* Q_k, \lambda _{\textrm{can}\,})\) be the unit co-disk bundle associated to the Riemannian metric g, where \(\lambda _{\textrm{can}\,}\) is the Liouville form on the cotangent bundle \(T^*Q_k\). Then \({\varvec{W}}_{\!\!\, \alpha }\) is a Liouville domain filling \(\alpha \).

Since the Riemannian metric g is negatively curved, a theorem of Margulis [71] shows that

$$\begin{aligned} h_{{\text {top}}}(\phi _\alpha ) = \lim _{t \rightarrow +\infty } \frac{\log \left( P^t(\phi _\alpha ) \right) }{t}, \end{aligned}$$

where \(P^t(\phi _\alpha )\) denotes the number of periodic orbits of the flow \(\phi _\alpha \) of length \(<t\), see also [25]. Since all periodic Reeb orbits have Morse index zero and are non-contractible, there are no Floer cylinders starting or ending at these orbits. Hence there is a bijection between the Reeb orbits of action \(<a\) and the generators of \({\text {SH}}^a ({\varvec{W}}_{\!\!\, \alpha })\), up to a finite error coming from the finitely many critical points of the function f. It follows that

$$\begin{aligned} \Gamma ( {\text {SH}}({\varvec{W}}_{\alpha })) \,=\, \lim _{t \rightarrow +\infty } \frac{\log \left( P^t(\phi _\alpha ) \right) }{t} \end{aligned}$$

see [7, 67] for details. Combining these two equalities we get

$$\begin{aligned} \Gamma ( {\text {SH}}({\varvec{W}}_{\alpha })) \,=\, h_{{\text {top}}}(\phi _\alpha ) \,=\, c, \end{aligned}$$

while as noted above \({\text {vol}}_\alpha (S^*Q_k) = 1\). \(\square \)

8.3 Wrapped Floer homology and collapse of its exponential growth

In a similar way we obtain a collapse result for the exponential growth of another symplectic invariant called wrapped Floer homology. The wrapped Floer homo- logy \({\text {WH}}({\varvec{W}},L)\) is an invariant associated to a Liouville domain \({\varvec{W}}\) and an asymptotically conical exact Lagrangian submanifold L of \({\varvec{W}}\). One of several references giving the precise definition of \({\text {WH}}({\varvec{W}},L)\) is [11].

With \(\alpha \) the contact form on the boundary of a Liouville domain \({\varvec{W}}_{\!\!\, \alpha }\), one can think of the chain complex associated to \({\text {WH}}({\varvec{W}}_{\!\!\, \alpha },L)\) as the \(\mathbb {Z}_2\)-vector space generated by the Reeb chords of \(\alpha \) that start and end on \(\partial L\), and by the intersection points of L and a \(C^2\)-small perturbation of L. The differential of \({\text {WH}}({\varvec{W}}_{\!\!\, \alpha },L)\) counts Floer strips connecting generators.

As in the case of symplectic homology there is a filtration of \({\text {WH}}({\varvec{W}}_{\!\!\, \alpha },L)\) by the action of the generators, and again the action of Reeb chords is equal to their time. For each real number \(a>0\) let \({\text {WH}}^{<a}({\varvec{W}}_{\!\!\, \alpha },L)\) be the homology of the subcomplex generated by Reeb chords and intersection points of action \(<a\). Again there are natural homomorphisms

$$\begin{aligned} \Psi ^a_{\alpha } :{\text {WH}}^{<a}({\varvec{W}}_{\!\!\, \alpha },L) \rightarrow {\text {WH}}({\varvec{W}}_{\!\!\, \alpha },L), \end{aligned}$$

and we define the exponential growth rate of \({\text {WH}}({\varvec{W}}_{\!\!\, \alpha },L)\) by

$$\begin{aligned} \Gamma ({\text {WH}}({\varvec{W}}_{\!\!\, \alpha },L)):= \limsup _{a \rightarrow +\infty } \frac{\log \left( \mathrm {rank\,}(\Psi ^a_{\alpha }) \right) }{a}. \end{aligned}$$

The following result was obtained in [11].

Theorem

Let \({\varvec{W}}_{\!\!\, \alpha }\) be a Liouville domain filling a contact form \(\alpha \) on \((\Sigma ,\xi )\), and let L be an asymptotically conical exact Lagrangian submanifold of \({\varvec{W}}_{\!\!\, \alpha }\) whose intersection with \(\partial {\varvec{W}}_{\!\!\, \alpha }\) is a sphere.Footnote 1 Then

$$\begin{aligned} h_{{\text {top}}}(\phi _\alpha ) \,\ge \, \Gamma ( {\text {WH}}({\varvec{W}}_{\!\!\, \alpha }), L). \end{aligned}$$

Together with Theorem 1.1 we obtain

Corollary 8.3

Let \((\Sigma ,\xi )\) be a contact manifold fillable by Liouville domains. Then for every \(\varepsilon >0\) there exists a contact form \(\alpha \) with \({\text {vol}}_\alpha (\Sigma ) = 1\) such that

$$\begin{aligned} \Gamma ( {\text {WH}}({\varvec{W}}_{\!\!\, \alpha },L)) \,\le \, \varepsilon , \end{aligned}$$

for any Liouville filling \({\varvec{W}}_{\alpha }\) of \(\alpha \) and any asymptotically conical Lagrangian submanifold L of \({\varvec{W}}_{\alpha }\) whose intersection with \(\partial {\varvec{W}}_{\alpha }\) is a sphere.

Classical examples of pairs \(({\varvec{W}}_{\!\!\, \alpha }, L)\) are the unit co-disk bundles \(D^*(g)\) over a closed Riemannian manifold Q, with L a co-disk \(D_q^*(g)\) over a point \(q \in Q\). Here is the analogue of Lemma 8.2.

Lemma 8.4

Let \(Q_k\) be the closed orientable surface of genus \(k \ge 2\). Then for every real number \(c \ge 2 \pi \sqrt{2(k-1)}\) there exists a Riemannian metric g on \(Q_k\) of area \(1/(2\pi )\) such that

$$\begin{aligned} \Gamma ( {\text {WH}}(D^*(g), D_q^*(g)) \,=\, c. \end{aligned}$$

Proof

As in the proof of Lemma 8.2 we appeal to [45] and take a negatively curved Riemannian metric g on \(Q_k\) of area \(1/(2\pi )\) whose geodesic flow has topological entropy

$$\begin{aligned} h_{{\text {top}}}(g) \,=\, c. \end{aligned}$$

Since g is negatively curved, Manning’s inequality in Theorem A.2 is an equality,

$$\begin{aligned} h_{{\text {top}}} (g) \,=\, h_{{\text {vol}}}(g). \end{aligned}$$

On the other hand, it is clear that

$$\begin{aligned} h_{{\text {vol}}} (g) \,=\, \lim _{t \rightarrow \infty } \frac{\log C^t(\phi _g,q)}{t} \quad \forall \, q \in Q_k, \end{aligned}$$

where \(C^t(\phi _g,q)\) denotes the number of geodesics from q to q of length \(<t\). Furthermore, since the Morse indices of all geodesics vanish, we have as for the symplectic homology that

$$\begin{aligned} \Gamma ( {\text {WH}}(D^*(g), D_q^*(g)) \,=\, \lim _{t \rightarrow \infty } \frac{\log \left( C^t(\phi _g,q) \right) }{t}. \end{aligned}$$

These four identities prove the lemma. \(\square \)