1 Introduction
Let M be a smooth closed n-dimensional manifold. We denote by
$\mathcal {M}$
the Fréchet manifold consisting of smooth metrics on M. We denote by
$\mathcal {M}^{k,\alpha }$
the set of metrics with regularity
$C^{k,\alpha }$
,
$k \in \mathbb {N}, \alpha \in (0,1)$
. We fix a smooth metric
$g_0 \in \mathcal {M}$
with Anosov geodesic flow
$\varphi _t^{g_0}$
and define the unit tangent bundle by
${S_{g_0}M := \{(x,v) \in TM ~|~ |v|_{g_0}= 1\}}$
. Recall that being Anosov means that there exists a flow-invariant continuous splitting
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu1.png?pub-status=live)
such that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu2.png?pub-status=live)
where the constants
$C, \unicode{x3bb}> 0$
are uniform and the norm here is the one induced by the Sasaki metric of
$g_0$
. Such a property is satisfied in negative curvature.
1.1 Geodesic stretch and marked length spectrum rigidity
The set of primitive free homotopy classes
$\mathcal {C}$
of M is in one-to-one correspondence with the primitive conjugacy classes of
$\pi _1(M,x_0)$
(where
$x_0 \in M$
is arbitrary). When
$g_0$
is Anosov, there exists a unique closed geodesic
$\gamma _{g_0}(c)$
in each primitive free homotopy class
$c \in \mathcal {C}$
(see [Reference KlingenbergKli74]). This allows us to define the marked length spectrum of the metric
$g_0$
by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu3.png?pub-status=live)
where
$\ell _{g_0}(\gamma )$
denotes the
$g_0$
-length of a curve
$\gamma \subset M$
computed with respect to
$g_0$
. The marked length spectrum can alternatively be defined for the whole set of free homotopy classes, but it is obviously an equivalent definition. Given
$c \in \mathcal {C}$
, we will write
$\delta _{g_0}(c)$
to denote the probability Dirac measure carried by the unique
$g_0$
-geodesic
$\gamma _{g_0}(c) \in c$
.
It was conjectured by Burns and Katok [Reference Burns and KatokBK85] that the marked length spectrum of negatively curved manifolds determines the metric up to isometry in the sense that two negatively curved metrics g and
$g_0$
with the same marked length spectrum (namely
$L_g = L_{g_0}$
) should be isometric. Although the conjecture was proved for surfaces by Croke and Otal [Reference CrokeCro90, Reference OtalOta90]) and in some particular cases in higher dimension (for conformal metrics by Katok [Reference KatokKat88] and when
$(M,g_0)$
is a locally symmetric space by the work of Hamenstädt and of Besson, Courtois and Gallot [Reference Besson, Courtois and GallotBCG95, Reference HamenstädtHam99]), it is still open in dimension higher than or equal to
$3$
and open even in dimension
$2$
in the more general setting of Riemannian metrics with Anosov geodesic flows. The same type of problems can also be posed for billiards, and we mention recent results on this problem by Avila, De Simoi and Kaloshin [Reference Avila, De Simoi and KaloshinADSK16] and De Simoi, Kaloshin and Wei [Reference De Simoi, Kaloshin and WeiDSKW17] for convex domains close to ellipses (although the Anosov case would rather correspond to the case of hyperbolic billiards). Recently, the first and last author obtained the following result on the Burns–Katok conjecture.
Theorem 1.1. (Guillarmou and Lefeuvre [Reference Guillarmou and LefeuvreGL19])
Let
$(M,g_0)$
be a smooth Riemannian manifold with Anosov geodesic flow, and further assume that its curvature is non-positive if
$\dim M \geq 3$
. Then there exists
$k \in \mathbb {N}$
depending only on
$\dim M$
and
$\varepsilon> 0$
small enough depending on
$g_0$
such that the following statement holds: if
$g \in \mathcal {M}$
is such that
$\|g-g_0\|_{C^k} \leq \varepsilon $
and
$L_g = L_{g_0}$
, then g is isometric to
$g_0$
.
One of the aims of this paper is to further investigate this result from different perspectives: new stability estimates and a refined characterization of the condition under which the isometry may hold. More precisely, we can relax the assumption that the two marked length spectra of g and
$g_0$
exactly coincide to the weaker assumption that they ‘coincide at infinity’ and still obtain the isometry. In what follows, we say that
$L_g/L_{g_0} \rightarrow 1$
when
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqn1.png?pub-status=live)
for any sequence
$(c_j)_{j \in \mathbb {N}}$
of primitive free homotopy classes such that
$\lim _{j\to \infty }L_{g_0}(c_j)=+\infty $
, or equivalently
$\lim _{j\to \infty }L_{g}(c_j)/L_{g_0}(c_j)=1$
, if
$\mathcal {C}=(c_j)_{j\in \mathbb {N}}$
is ordered by the increasing lengths
$L_{g_0}(c_j)$
. We prove in Appendix A that
$L_g/L_{g_0} \rightarrow 1$
is actually equivalent to
$L_g = L_{g_0}$
. As a consequence, by [Reference Guillarmou and LefeuvreGL19], if (1.1) holds and if
$\|g-g_0\|_{C^k}<\varepsilon $
for some small enough
$\varepsilon>0$
, then g is isometric to
$g_0$
. If we restrict ourselves to metrics with the same topological entropy, the knowledge of
$L_{g}(c_j)/L_{g_0}(c_j)$
for a subsequence so that the geodesic
$\gamma _{g_0}(c_j)$
equidistributes is even sufficient; see Theorem 2.9.
We develop a new proof strategy, different from [Reference Guillarmou and LefeuvreGL19], which relies on the introduction of the geodesic stretch between two metrics. This quantity was first introduced by Croke and Fathi [Reference Croke and FathiCF90] and further studied by the second author [Reference KnieperKni95]. If g is close enough to
$g_0$
, then by Anosov structural stability, the geodesic flows
$\varphi ^{g_0}$
and
$\varphi ^g$
are orbit equivalent via a homeomorphism
$\psi _g$
, that is, they are conjugate up to a time reparametrization
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu4.png?pub-status=live)
for some time rescaling
$\kappa _g(z,t)$
. The infinitesimal stretch is the infinitesimal function of time reparametrization
$a_g(z)=\partial _{t}\kappa _g(z,t)|_{t=0}$
: it satisfies
$d\psi _g(z) X_{g_0}(z) = a_g(z) X_{g}(\psi _g(z))$
where
$z \in S_{g_0}M$
and
$X_{g_0}$
(respectively,
$X_g$
) denotes the geodesic vector field of
$g_0$
(respectively, g). The geodesic stretch between g and
$g_0$
with respect to the Liouville (normalized with total mass 1) measure
$\mu ^{\textrm{L}}_{g_0}$
of
$g_0$
is then defined by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu5.png?pub-status=live)
The function
$a_g$
is uniquely defined up to a coboundary [Reference de la Llave, Marco and MoriyóndlLMM86] so that the geodesic stretch is well defined. (Although this is only used in §5.2, we also point out that the existence of the conjugacy
$\psi _g$
and of the reparametrization
$a_g$
is actually global and one need not assume that the two metrics are close. This is a very particular feature of the geodesic structure. We refer to Appendix B for a proof of this fact.)
Since obviously
$\langle \delta _{g_0}(c_j),a_g\rangle =L_g(c_j)/L_{g_0}(c_j)$
, we have
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu6.png?pub-status=live)
if
$(c_j)_{j\in \mathbb {N}}\subset \mathcal {C}$
is a sequence so that the uniform probability measures
$(\delta _{g_0}(c_j))_{j \in \mathbb {N}}$
supported on the closed geodesics of
$g_0$
in the class
$c_j$
converge to
$\mu ^{\textrm{L}}_{g_0}$
in the weak sense of measures. (The existence of the sequence
$c_j$
follows from [Reference SigmundSig72, Theorem 1].) In particular,
$L_g=L_{g_0}$
implies that
$I_{\mu _{g_0}^{\textrm{L}}}(g_0,g)=1$
(alternatively,
$L_g=L_{g_0}$
implies that
$a_g$
is cohomologous to
$1$
by Livsic’s theorem). While being of interest in its own right, it turns out that this method involving the geodesic stretch provides a new estimate which quantifies locally the distance between isometry classes in terms of this geodesic stretch functional (below
$H^{-1/2}(M)$
denotes the
$L^2$
-based Sobolev space of order
$-1/2$
and
$\alpha \in (0,1)$
is any fixed exponent).
Theorem 1.2. Let
$(M,g_0)$
be a smooth Riemannian n-dimensional manifold with Anosov geodesic flow and further assume that its curvature is non-positive if
$n \geq 3$
. There exists
$k \in \mathbb {N}$
large enough depending only on n, some positive constants
$C,\varepsilon $
depending on
$g_0$
and
$C_n>0$
depending on n such that for all
$\alpha \in (0,1)$
, the following statement holds: for each
$g \in \mathcal {M}^{k,\alpha }$
with
$\|g-g_0\|_{C^{k,\alpha }(M)} \leq \varepsilon $
, there exists a
$C^{k+1,\alpha }$
-diffeomorphism
$\psi : M \rightarrow M$
such that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu7.png?pub-status=live)
where
$J^u_{g_0}$
is the unstable Jacobian of
$\varphi ^{g_0}$
,
${\textbf {P}}$
denotes the topological pressure for the
$\varphi ^{g_0}$
flow defined by (2.11),
$a_g$
is the reparametrization coefficient relating
$\varphi ^{g_0}$
and
$\varphi ^g$
defined above, and
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu8.png?pub-status=live)
In particular, if (1.1) holds, then
$g_0$
and g are isometric.
Note that g need not have non-positive curvature in the theorem. We also remark that the curvature condition on
$g_0$
can be replaced by the injectivity of the X-ray transform
$I_2$
on divergence-free symmetric
$2$
-tensors, and similarly for Theorem 1.3 below. From the proof one sees that the exponent k can be taken to be
$k=3n/2+17$
.
Theorem 1.2 is an improvement over the Hölder stability result [Reference Guillarmou and LefeuvreGL19, Theorem 3] as it only involves the asymptotic behavior of
$L_g/L_{g_0}$
or some natural quantity from thermodynamic formalism. We insist on the fact that the new ingredient here is the stability estimate in itself (the rigidity result is not new).
We also emphasize that one of the key facts to prove this theorem still boils down to some elliptic estimate on some variance operator acting on symmetric
$2$
-tensors, denoted by
$\Pi _2^{g_0}$
in [Reference Guillarmou and LefeuvreGL19, Reference GuillarmouGui17]: indeed, we show that the combination of the Hessians of the geodesic stretch at
$g_0$
and of the pressure functional can be expressed in terms of this variance operator, which enjoys uniform lower bounds
$C_{g_0}\|\psi ^*g-g_0\|_{H^{-1/2}}$
for some
$C_{g_0}>0$
, at least once we have factored out the gauge (the diffeomorphism action by pullback on metrics).
We also notice that in Theorem 1.2, although the
$H^{-1/2}(M)$
norm is a weak norm, a straightforward interpolation argument using that
$\|g\|_{C^{k,\alpha }}\leq \|g_0\|_{C^{k,\alpha }}+\varepsilon $
is uniformly bounded shows that an estimate of the form
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu9.png?pub-status=live)
holds for any
$k^{\prime }<k-n/2$
and some explicit
$\delta \in (0,1/2)$
depending on
$k,k^{\prime }$
(
$C>0$
depending only on
$g_0$
).
1.2 Variance and pressure metric
The variance operator appearing in the proof of Theorem 1.2 can be defined for
$h_1,h_2\in C^\infty (M;S^2T^*M)$
satisfying the condition
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqn2.png?pub-status=live)
for
$i=1,2$
(see §2.3 for further details on tensor analysis) by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu10.png?pub-status=live)
where
$z = (x,v) \in SM$
and, given a symmetric
$2$
-tensor
$h \in C^\infty (M;S^2 T^*M)$
, we define the pullback operator
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu11.png?pub-status=live)
The quadratic form
$\langle \Pi _2^{g_0}h,h\rangle $
corresponds to the variance
${\textrm{Var}}_{\mu _L}(\pi _2^*h)$
for
$\varphi _t^{g_0}$
with respect to the Liouville measure of the lift
$\pi _2^*h$
of the tensor h to
$SM$
(see §2.5 and (2.5)). Note that the trace-free condition (1.2) is equivalent to
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu12.png?pub-status=live)
see §2.3. The integral defining
$\Pi _2^{g_0}$
then converges (in the
$L^1$
sense) by the rapid mixing of
$\varphi ^{g_0}$
(proved in [Reference LiveraniLiv04]). The operator
$\Pi ^{g_0}_2$
is a pseudodifferential operator of order
$-1$
that is elliptic on divergence-free tensors (see [Reference Goüezel and LefeuvreGL, Reference Guillarmou and LefeuvreGL19, Reference GuillarmouGui17]). As a consequence, it satisfies elliptic estimates on all Sobolev or Hölder spaces (see Lemma 2.1). More precisely, there is
$C_{g_0}>0$
such that, for all
$h\in H^{-1/2}(M;S^2T^*M)$
which is divergence-free (that is,
${\textrm{Tr}}_{g_0}(\nabla ^{g_0}h)=0$
),
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqn3.png?pub-status=live)
provided
$g_0$
is Anosov with non-positive curvature (or simply Anosov if
$\dim M=2$
). We show in Proposition 4.1 that
$g\mapsto \Pi _2^g$
is continuous with values in
$\Psi ^{-1}(M)$
and this implies that for
$g_0$
a smooth Anosov metric (with non-positive curvature if
$\dim M>2$
), (1.3) holds uniformly if we replace
$g_0$
by any metric g in a small
$C^\infty $
-neighborhood of
$g_0$
. This allows us to obtain a more uniform version of Theorem 1.2.
Theorem 1.3. Let
$(M,g_0)$
be a smooth Riemannian n-dimensional manifold with Anosov geodesic flow and further assume that its curvature is non-positive if
$n \geq 3$
. Then there exist
$k \in \mathbb {N}$
,
$\varepsilon> 0$
and
$C_{g_0}$
depending on
$g_0$
such that for all
$g_1,g_2 \in \mathcal {M}$
such that
$\|g_1-g_0\|_{C^k} \leq \varepsilon $
,
$\|g_2-g_0\|_{C^k} \leq \varepsilon $
, there is a
$C^k$
- diffeomorphism
$\psi :M\to M$
such that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu13.png?pub-status=live)
with
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu14.png?pub-status=live)
In particular, if
$L_{g_1}/L_{g_2}\to 1$
, then
$g_2$
is isometric to
$g_1$
.
This result suggests defining a distance on isometry classes of metrics (here we mean isometries homotopic to the identity) from the marked length spectrum by setting, for two
$C^{k,\alpha }$
metrics
$g_1,g_2$
,
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu15.png?pub-status=live)
We have the following corollary of Theorem 1.3.
Corollary 1.4. The map
$d_L$
descends to the space of isometry classes of Anosov non-positively curved metrics and defines a distance near the diagonal.
We also define the Thurston asymmetric distance by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu16.png?pub-status=live)
and show that this is a distance on isometry classes of metrics with topological entropy equal to
$1$
; see Proposition 5.4. This distance was introduced in Teichmüller theory by Thurston in [Reference ThurstonThu98].
The elliptic estimate (1.3) also allows us to define a pressure metric on the open set consisting of isometry classes of Anosov non-positively curved metrics (contained in
$\mathcal {M}/\mathcal {D}_0$
if
$\mathcal {D}_0$
is the group of smooth diffeomorphisms isotopic to the identity) by setting, for
$h_1,h_2\in T_{g_0}(\mathcal {M}/\mathcal {D}_0)\subset C^\infty (M;S^2T^*M)$
,
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu17.png?pub-status=live)
We show in §3.3.1 that this metric is well defined and restricts to (a multiple of) the Weil–Petersson metric on Teichmüller space if
$\dim M=2$
: it is related to the construction of Bridgeman et al [Reference Bridgeman, Canary, Labourie and SambarinoBCLS15, Reference Bridgeman, Canary and SambarinoBCS18] and McMullen [Reference McMullenMM08], but with the difference that we work here in the setting of variable negative curvature and the space of metrics considered here is infinite-dimensional. In a related but different context with infinite dimension, we note that the variance is used to define a metric on the space of Hölder potentials by Giulietti et al [Reference Giulietti, Kloeckner, Lopes and MarconGKLM18] and its curvature is studied by Lopes and Ruggiero [Reference Lopes and RuggieroLR18].
We finally notice that, in the study of Katok entropy conjecture near locally symmetric spaces, the variance was an important tool in the work of Pollicott and Flaminio [Reference FlaminioFla95, Reference PollicottPol94]. In that case, one can use representation theory to analyze this operator.
2 Preliminaries
2.1 Notation
If
$H=C^k,H^s,C^{-\infty }$
etc. is a regularity scale and
$E\to M$
a smooth bundle over a smooth compact manifold M, we will use the notation
$H(M;E)$
for sections of E with regularity H, while if N is a smooth manifold, we use the notation
$H(M,N)$
for the space of maps from M to N with regularity H.
2.2 Microlocal calculus
On a closed manifold M, we will denote by
$\Psi ^{m}(M;V)$
the space of classical pseudo-differential operators of order
$m\in \mathbb {R}$
acting on a vector bundle V over M (see [Reference Grigis and SjöstrandGS94]; the operators could map sections of two distinct vector bundles, but this will not be needed here). We recall that for fixed
$m \in \mathbb {R}$
, this is a Fréchet space: indeed, using a fixed smooth cutoff function
$\theta $
supported in a small neighborhood of the diagonal, a fixed system of charts, each
$A\in \Psi ^{m}(M;V)$
has Schwartz kernel
$\kappa _A$
that can be decomposed as
$\theta \kappa _A+(1-\theta )\kappa _A$
. For the
$(1-\theta )\kappa _A$
part we can use the
$C^\infty (M\times M; V\otimes V^*)$
topology, while for
$\chi \kappa _A$
we can use the semi-norms of the full symbols of
$\chi \kappa _A$
using the local charts and the left quantization in the charts. We also denote by
$H^{s}(M)$
the
$L^2$
-based Sobolev space of order
$s\in \mathbb {R}$
, with norm given by fixing an arbitrary Riemannian metric
$g_{0}$
on M. More precisely, denoting by
$\Delta $
the non-negative Laplacian associated to this metric, we define
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu18.png?pub-status=live)
and
$H^s(M)$
is the completion of
$C^\infty (M)$
with respect to this norm. This definition is naturally extended to sections of vector bundles. What is important is that the spaces and the norm (up to a scaling factor) do not depend on the choice of metric
$g_{0}$
. For
$k\in \mathbb {N}, \alpha \in (0,1)$
, the spaces
$C^{k,\alpha }(M)$
are the usual Hölder spaces and
$\mathcal {D}^{\prime }(M)$
will denote the space of distributions dual to
$C^\infty (M)$
. We will denote by
$\langle \cdot ,\cdot \rangle _{L^2}$
the continuous extension of the pairing
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu19.png?pub-status=live)
to the pairing
$H^{s}(M) \times H^{-s}(M)\to \mathbb {C}$
for each
$s\in \mathbb {R}$
(and likewise for sections of bundles).
2.3 Symmetric tensors and X-ray transform
In this subsection, we assume that the metric g is fixed and that its geodesic flow
$\varphi ^g_t$
is Anosov on the unit tangent bundle
$SM$
of g. We denote by
$\mu ^{\textrm{L}}$
the Liouville measure, normalized to be a probability measure on
$SM$
. For the sake of simplicity, we drop the index g in the notation. Given an integer
$m \in \mathbb {N}$
, we denote by
$\otimes ^m T^*M \rightarrow M$
,
$S^mT^*M \rightarrow M$
the respective vector bundle of m-tensors and symmetric m-tensors on M. Given
$f \in C^\infty (M;S^m T^*M)$
, we denote by
$\pi _m^*f \in C^\infty (SM)$
the canonical morphism
$\pi _m^*f : (x,v) \mapsto f_x(v,\ldots ,v)$
. We also introduce the trace operator
$\operatorname {\textrm{Tr}} : C^\infty (M;S^{m+2} T^*M) \rightarrow C^\infty (M;S^{m} T^*M)$
defined pointwise in
$x \in M$
by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu20.png?pub-status=live)
where
$(\textbf{e}_1, \ldots , \textbf{e}_n)$
denotes an orthonormal basis of
$TM$
in a neighborhood of a fixed point
$x_0 \in M$
. Observe that, for
$f = \sum _{i,j=1}^n f_{ij} \textbf{e}_i^* \otimes \textbf{e}_j^* \in C^\infty (M;S^2T^*M)$
defined around
$x_0$
, we have
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu21.png?pub-status=live)
for some constant
$C_n = \int _{\mathbb {S}^{n-1}} v_1^2 dv$
depending on
$n=\dim M$
. This justifies the claim that the trace-free condition (1.2) was equivalent to the fact that the pullback of the symmetric tensor to
$SM$
was of average
$0$
.
The natural derivation of symmetric tensors is
$D := \sigma \circ \nabla $
, where
$\nabla $
is the Levi-Civita connection and
$\sigma : \otimes ^m T^*M \rightarrow S^m T^*M$
is the operation of symmetrization. This operator satisfies the important identity
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqn4.png?pub-status=live)
where X denotes the geodesic vector field on
$SM$
. The operator D is elliptic [Reference Goüezel and LefeuvreGL, Lemma 2.4] with trivial kernel when m is odd and one-dimensional kernel when m is even, given by the Killing tensors
$c\sigma (g^{\otimes m/2}), c \in \mathbb {R}$
(this is a simple consequence of (2.1) combined with the fact that the geodesic flow is ergodic in the Anosov setting). We denote by
$\langle \cdot , \cdot \rangle $
the scalar product on
$C^\infty (M;S^mT^*M)$
induced by the metric g (see [Reference Goüezel and LefeuvreGL, §2] for further details). The formal adjoint of D with respect to this scalar product is
$D^*=-\operatorname {\textrm{Tr}}\circ \nabla $
. We also denote by the same
$\langle \cdot , \cdot \rangle $
the natural
$L^2$
scalar product on
$C^\infty (SM)$
induced by the Liouville measure
$\mu ^{{\textrm{L}}}$
. The formal adjoint of
$\pi _m^*$
with respect to these two scalar products is denoted by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu22.png?pub-status=live)
where
$\mathcal {D}^{\prime }$
denotes the space of distributions, dual to
$C^\infty $
.
We recall that
$\mathcal {C}$
, the set of free homotopy classes in M, is in one-to-one correspondence with the set of conjugacy classes of
$\pi _1(M,x_0)$
for some arbitrary choice of
$x_0 \in M$
(see [Reference KlingenbergKli74]) and for each
$c \in \mathcal {C}$
there exists a unique closed geodesic
$\gamma (c) \in c$
. We denote its Riemannian length with respect to g by
$L(c)=\ell _g(\gamma (c))$
. The X-ray transform on
$SM$
is the operator defined by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu23.png?pub-status=live)
where
$z \in \gamma (c)$
is any point. This is a continuous linear operator when
$\ell ^\infty (\mathcal {C})$
is endowed with the sup norm on the sequences. Then the X-ray transform
$I_m$
of symmetric m-tensors is simply defined by
$I_m := I \circ \pi _m^*$
. Using (2.1), we immediately have
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqn5.png?pub-status=live)
Using the ellipticity of D, any tensor
$f \in C^\infty (M; S^m T^*M)$
can be decomposed uniquely as a sum
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqn6.png?pub-status=live)
with
$p \in C^\infty (M;S^{m-1}T^*M)$
and
$h \in C^\infty (M;S^{m}T^*M)$
is such that
$D^*h=0$
. We call
$Dp$
the potential part of f and h the solenoidal part. The same decomposition holds in Sobolev regularity
$H^s(M)$
,
$s \in \mathbb {R}$
, and in
$C^{k,\alpha }(M)$
regularity,
$k \in \mathbb {N}, \alpha \in (0,1)$
. We will write
$h = \pi _{\ker D^*}f$
and the solenoidal projection
is a pseudodifferential operator of order
$0$
[Reference Goüezel and LefeuvreGL, Lemma 2.6] (here
$\Delta _g := D^*_g D_g$
is the Laplacian on
$1$
-forms). The X-ray transform is said to be solenoidal injective (or s-injective for short) if (2.2) is an equality. It is conjectured that
$I_m$
is s-injective as long as the metric is Anosov, but it is only known in the following cases:
-
• for
$m=0,1$ [Reference Dairbekov and SharafutdinovDS10];
-
• for any
$m \in \mathbb {N}$ in dimension
$2$ [Reference GuillarmouGui17, Reference Paternain, Salo and UhlmannPSU14];
-
• for any
$m \in \mathbb {N}$ , in any dimension in non-positive curvature [Reference Croke and SharafutdinovCS98].
It is also known that
$\ker I_m/\operatorname {\textrm{ran}} D$
is finite-dimensional for general Anosov geodesic flow (see [Reference Dairbekov and SharafutdinovDS03, Theorem 1.5] or [Reference GuillarmouGui17, Remark 3.7]).
The direct study of the analytic properties of
$I_m$
is difficult as this operator involves integrals over the set of closed orbits, which is not a manifold. Nevertheless, in [Reference GuillarmouGui17], the second author introduced an operator
$\Pi _m$
that involves a sort of integration of tensors over ‘all orbits’, and this space is essentially the manifold
$SM$
. The construction of
$\Pi _m: C^\infty (M;S^mT^*M)\to \mathcal {D}^{\prime }(M;S^mT^*M)$
relies on microlocal tools coming from [Reference Dyatlov and ZworskiDZ16, Reference Faure and SjöstrandFS11], but a simpler definition that uses the fast mixing of the flow
$\varphi _t$
is given by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqn7.png?pub-status=live)
if
$\langle f,1\rangle =\int _{SM} f \, d\mu ^{\textrm{L}}=0$
and
$\Pi (1):=0$
. The convergence of the integral as
$T\to \infty $
is ensured by the exponential decay of correlations [Reference LiveraniLiv04] (but also follows from the existence of the variance [Reference Katsuda and SunadaKS90]). We can thus write, for
$\langle f,1\rangle =0$
,
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu24.png?pub-status=live)
We note the following useful properties of
$\Pi $
, proved in [Reference GuillarmouGui17, Theorem 1.1]:
-
•
$\Pi : H^s(SM)\to H^{-s}(SM)$ is bounded for all
$s>0$ ;
-
• if
$f\in H^s(SM)$ with
$s>0$ , then
$X\Pi f=0$ ;
-
• if f and
$Xf$ belong to
$H^s(SM)$ for
$s>0$ , then
$\Pi Xf=0$ . (In [Reference GuillarmouGui17], f is assumed to be in
$H^{s+1}(SM)$ , but one can reduce to the case
$f\in H^s(SM)$ by using a density argument and [Reference Dyatlov and ZworskiDZ19, Lemma E.45].)
As is well known (see; for example; [Reference Katsuda and SunadaKS90, Proof of Proposition 1.2.]), we can make a link between
$\Pi $
and the variance in the central limit theorem for Anosov geodesic flows. Let us quickly explain this fact by using the fast mixing of the flow. The variance of
$\varphi _t$
with respect to the Liouville measure
$\mu ^{\textrm{L}}$
is defined for
$u\in C^\alpha (SM), \alpha \in (0,1)$
real-valued, by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqn8.png?pub-status=live)
under the condition that
$\int _{SM}u\, d\mu ^{\textrm{L}}=0$
. We observe, since
$\varphi _t$
preserves
$\mu ^{\textrm{L}}$
, that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu25.png?pub-status=live)
where the
$L^2$
pairing is with respect to
$\mu ^{\textrm{L}}$
. By exponential decay of correlations [Reference LiveraniLiv04], we have, for
$|r|$
large,
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu26.png?pub-status=live)
for some
$\alpha>0,\nu >0$
,
$C>0$
independent of u. Thus, by the Lebesgue theorem,
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqn9.png?pub-status=live)
if
$\langle u,\textbf{1}\rangle =0$
, where
$\textbf{1}$
denotes the constant function equal to
$1$
, showing that the quadratic form associated to our operator
$\Pi $
is nothing more than the variance. For a symmetric
$2$
-tensor h satisfying
$\langle h,g\rangle _{L^2}=\int _{M}{\textrm{Tr}}_{g}(h)\, d{\textrm{vol}}_{g}=0$
, we have
$\int _{SM}\pi _2^*h \, d\mu _{g}^L=0$
and
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu27.png?pub-status=live)
We have the following properties for
$\Pi _m$
.
-
•
$\Pi _m$ is a positive self-adjoint pseudodifferential operator of order
$-1$ , elliptic on solenoidal tensors; see [Reference GuillarmouGui17, Theorem 3.5] and [Reference Goüezel and LefeuvreGL, Lemma 4.3].
-
•
$\Pi _mD = 0$ and
$D^*\Pi _m=0$ (by [Reference GuillarmouGui17, Theorem 3.5] and
$X\pi _{m-1}^*= \pi _m^*D$ ).
-
• If
$I_m$ is s-injective, then
$\Pi _m$ is invertible on solenoidal tensors in the sense that there exists a pseudodifferential operator Q of order
$1$ such that
$Q\Pi _m=\pi _{\ker D^*}$ ; see [Reference Goüezel and LefeuvreGL, Theorem 4.7].
-
• Conversely, if
$\Pi _m|_{\ker D^*}$ is injective, then
$I_m$ is s-injective. Indeed, by [Reference GuillarmouGui17, Corollary 2.8], if
$I_mh=0$ then
$\pi _m^*h=Xu$ for some
$u\in C^\infty (SM)$ and thus
$\Pi _mh={\pi _m}_*\Pi Xu=0$ .
In particular, using the spectral theorem, there is a bounded self-adjoint operator
$\sqrt {\Pi _m}$
on
$L^2$
such that
$\sqrt {\Pi _m}\sqrt {\Pi _m}=\Pi _m$
. We add the following property, the use of which will be crucial in this paper.
Lemma 2.1. If
$(M,g)$
has Anosov geodesic flow and
$I_2$
is s-injective, there exists a constant
$C> 0$
such that, for all tensors
$h \in H^{-1/2}(M; S^2T^*M)$
,
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu28.png?pub-status=live)
Proof In [Reference Goüezel and LefeuvreGL, Theorem 4.4 and Lemma 2.2], the principal symbol of
$\Pi _2$
was computed and turned out to be
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu29.png?pub-status=live)
for some positive definite diagonal endomorphism
$A_2$
which is constant on both subspaces
$S^2_0T^*M:=\{h\in S^2T^*M |\, {\textrm{Tr}}_g(h)=0\}$
and
$\mathbb {R} g=\{\unicode{x3bb} g\in S^2T^*M\, |\, \unicode{x3bb} \in \mathbb {R}\}$
. Here
$i_\xi $
is the interior product with the dual vector
$\xi ^\sharp \in T_xM$
of
$\xi $
with respect to the metric. We introduce the symbol
$b\in C^\infty (T^*M)$
of order
$-1/2$
defined by
$b :(x,\xi ) \mapsto \chi (x,\xi )|\xi |^{-1/2}A_2$
, where
$\chi \in C^\infty (T^*M)$
vanishes near the
$0$
section in
$T^*M$
and is equal to
$1$
for
$|\xi |>1$
, and define
$B := \operatorname {\textrm{Op}}(b)\in \Psi ^{-1/2}(M;S^2T^*M)$
, where
$\operatorname {\textrm{Op}}$
is a quantization on M. Using that the principal symbol of
$\pi _{\ker D^*}$
is
$\pi _{\ker i_\xi }$
(see [Reference Goüezel and LefeuvreGL, Lemma 2.6]), we observe that
$\Pi _2 = \pi _{\ker D^*}B^*B\pi _{\ker D^*} + R$
, where
$R\in \Psi ^{-2}(M;S^2T^*M)$
. Thus, given
$h \in H^{-1/2}(M,S^2 T^*M)$
,
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqn10.png?pub-status=live)
By ellipticity of B, there exists a pseudodifferential operator Q of order
$1/2$
such that
$QB\pi _{\ker D^*} = \pi _{\ker D^*} + R^{\prime }$
, where
$R^{\prime }\in \Psi ^{-\infty }(M;S^2T^*M)$
is smoothing. Thus there is
$C>0$
such that, for each
$h\in C^\infty (M;S^2T^*M)$
,
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu30.png?pub-status=live)
Since Lemma 2.1 is trivial on potential tensors, we can already assume that h is solenoidal, that is,
$\pi _{\ker D^*}h=h$
. Recalling (2.7), we obtain that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqn11.png?pub-status=live)
Now, assume by contradiction that the statement in Lemma 2.1 does not hold, that is, we can find a sequence of tensors
$f_n \in C^\infty (M;S^2 T^*M)$
such that
$\|f_n\|_{H^{-1/2}}=1$
with
$D^*f_n=0$
and
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu31.png?pub-status=live)
Up to a subsequence, and since R is of order
$-2$
, we can assume that
$Rf_n \rightarrow v_1$
in
$H^{1/2}$
for some
$v_1$
, and
$R^{\prime }f_n \rightarrow v_2$
in
$H^{-1/2}$
. Then, using (2.8), we obtain that
$(f_n)_{n \in \mathbb {N}}$
is a Cauchy sequence in
$H^{-1/2}$
which thus converges to an element
$v_3 \in H^{-1/2}$
such that
$\|v_3\|_{H^{-1/2}}=1$
and
$D^*v_3=0$
. By continuity,
$\Pi _2f_n \rightarrow \Pi _2v_3$
in
$H^{1/2}$
and thus
$\langle \Pi _2v_3,v_3 \rangle =0$
. Since
$v_3$
is solenoidal, we get
$\sqrt {\Pi _2}v_3=0$
, thus
$\Pi _2v_3=0$
. Since we assumed
$I_2$
s-injective,
$\Pi _2$
is also injective by [Reference Goüezel and LefeuvreGL, Lemma 4.6]. This implies that
$v_3 \equiv 0$
, thus contradicting
$\|v_3\|_{H^{-1/2}}=1$
.
We note that the same proof also works for tensors of any order
$m \in \mathbb {N}$
. In fact we can even get a uniform estimate.
Lemma 2.2. Let
$(M,g_0)$
be a smooth compact Anosov Riemannian manifold with
$I_2^{g_0}$
being s-injective. There exist a
$C^\infty $
neighborhood
$\mathcal {U}_{g_0}$
of
$g_0$
and a constant
$C> 0$
such that for all
$g\in \mathcal {U}_{g_0}$
and all tensors
$h \in H^{-1/2}(M; S^2T^*M)$
,
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu32.png?pub-status=live)
Proof First, let
$g_0$
be fixed Anosov metric with
$I_2^{g_0}$
s-injective (in particular, it is the case if it has non-positive curvature). Proposition 4.1 (which will be proved later) shows that the operator
$\Pi _2=\Pi _2^g$
is a continuous family as a map
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu33.png?pub-status=live)
where
$\mathcal {U}_{g_0}\subset C^\infty (M;S^2T^*M)$
is a
$C^\infty $
-neighborhood of
$g_0$
and
$\Psi ^{-1}(M;S^2T^*M)$
is equipped with its Fréchet topology as explained before. Let
$h \in \ker D^*_g$
be a solenoidal (with respect to g) symmetric
$2$
-tensor, then
$h=\pi _{\ker D^*_g}h$
. Let
$C_{g_0}> 0$
be the constant provided by Lemma 2.1 applied to the metric
$g_0$
. We choose
$\mathcal {U}_{g_0}$
small enough so that
$\|\Pi ^g_2-\Pi ^{g_0}_2\|_{H^{-1/2}\rightarrow H^{1/2}} \leq C_{g_0}/3$
(this is made possible by the continuity of
$g \mapsto \Pi ^g_2 \in \Psi ^{-1}$
). Then
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu34.png?pub-status=live)
But the map
is continuous: this follows from the fact that one can construct a full parametrix
$Q_g\in \Psi ^{-2}(M)$
of
$\Delta _g$
modulo smoothing in a continuous way with respect to g (by standard elliptic microlocal analysis), the fact that
$\Delta _g$
is injective since
$\ker D_g=0$
for g Anosov (as
$D_gu=0$
implies
$X\pi _1^*u=0$
, thus
$\pi _1^*u$
has to be constant, thus
$0$
since
$\pi _1^*u(x,-v)=-\pi _1^*u(x,v)$
) and the continuity of composition of pseudodifferential operators. This implies that for g in a possibly smaller neighborhood
$\mathcal {U}_{g_0}$
of
$g_0$
, using
$h=\pi _{\ker D^*_g}h$
,
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu35.png?pub-status=live)
The proof is complete.
We also observe that the generalization of the previous lemma to tensors of any order is straightforward. As mentioned earlier, an immediate consequence of the previous lemma is the following proposition.
Proposition 2.3. Let
$(M,g_0)$
be a smooth Riemannian n-dimensional Anosov manifold with
$I_m^{g_0}$
s-injective. Then there exists a
$C^\infty $
-neighborhood
$\mathcal {U}_{g_0}$
of
$g_0$
in
$\mathcal {M}$
such that, for any
$g \in \mathcal {U}_{g_0}$
, for any
$m \in \mathbb {N}$
,
$I^g_m$
is s-injective.
Proof As mentioned above (before Lemma 2.1), the s-injectivity of
$I_m^g$
is equivalent to that of
$\Pi _m^g$
on solenoidal tensors and the previous lemma allows us to conclude.
2.4 The space of Riemannian metrics
We fix a smooth metric
$g_0 \in \mathcal {M}$
and consider an integer
$k \geq 2$
and
$\alpha \in (0,1)$
. We recall that the space
$\mathcal {M}$
of all smooth metrics is a Fréchet manifold. We denote by
$\mathcal {D}_0 := \operatorname {\textrm{Diff}}_0(M)$
the group of smooth diffeomorphisms on M that are isotopic to the identity; this is a Fréchet Lie group in the sense of [Reference HamiltonHam82, Section 4.6]. The right action
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu36.png?pub-status=live)
is smooth and proper [Reference EbinEbi68, Reference Ebin, Chern and SmaleEbi70]. Moreover, if g is a metric with Anosov geodesic flow, it is directly seen from ergodicity that there are no Killing vector fields and thus the isotropy subgroup
$\{\psi \in \mathcal {D}_0 ~|~ \psi ^*g=g\}$
of g is finite. For negatively curved metrics it is shown in [Reference FrankelFra66] that the action is free, that is, the isotropy group is trivial. One cannot apply the usual quotient theorem [Reference TrombaTro92, pp. 20] in the setting of Banach or Hilbert manifolds but rather smooth Fréchet manifolds instead (using the Nash–Moser theorem). Thus, in the setting of the space of smooth metrics with Anosov geodesic flows (the important fact, to apply Ebin’s slice theorem, is that metrics with Anosov geodesic flows do not have Killing vector fields, that is, infinitesimal isometries; this is due to the fact that
$\ker D|_{C^\infty (M,T^*M)} = \{ 0 \}$
as mentioned earlier, which itself follows from the ergodicity of the geodesic flow), which is an open set of a Fréchet vector space, the slice theorem says that there exist a neighborhood
$\mathcal {U}$
of
$g_0$
, a neighborhood
$\mathcal {V}$
of
${\textrm{Id}}$
in
$\mathcal {D}_0$
and a Fréchet submanifold
$\mathcal {S}$
containing
$g_0$
so that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqn12.png?pub-status=live)
is a diffeomorphism of Fréchet manifolds and
$T_{g_0}\mathcal {S}=\{h\in T_{g_0}\mathcal {M} ~|~ D_{g_0}^*h=0\}$
; see [Reference EbinEbi68, Reference Ebin, Chern and SmaleEbi70]. Moreover,
$\mathcal {S}$
parametrizes the set of orbits
$g\cdot \mathcal {D}_0$
for g near
$g_0$
and
$T_g\mathcal {S}\cap T(g \cdot \mathcal {D}_0)=0$
.
On the other hand, if one considers
$\mathcal {M}^{k,\alpha }$
, the space of metrics with
$C^{k,\alpha }$
regularity and
$\mathcal {D}^{k+1,\alpha }_0 := \operatorname {\textrm{Diff}}^{k+1,\alpha }_0(M)$
, the group of diffeomorphisms isotopic to the identity with
$C^{k+1,\alpha }$
regularity, then both spaces are smooth Banach manifolds. However, the action of
$\mathcal {D}_0^{k+1,\alpha }$
on
$\mathcal {M}^{k,\alpha }$
is no longer smooth but only topological, which also prevents us from applying the quotient theorem.
Nevertheless, recalling that
$g_0$
is smooth, if we consider
$\mathcal {O}^{k,\alpha }(g_0) := g_0\cdot \mathcal {D}^{k+1,\alpha }_0 \subset \mathcal {M}^{k,\alpha }$
, then this is a smooth submanifold of
$\mathcal {M}^{k,\alpha }$
and
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu37.png?pub-status=live)
Notice that (2.3) in
$C^{k,\alpha }$
regularity exactly says that given
$g \in \mathcal {O}^{k,\alpha }(g_0)$
, we have the decomposition
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqn13.png?pub-status=live)
Thus, an infinitesimal perturbation of a metric
$g \in \mathcal {O}^{k,\alpha }(g_0)$
by a symmetric
$2$
-tensor that is solenoidal with respect to g is actually an infinitesimal displacement transversally to the orbit
$\mathcal {O}^{k,\alpha }(g_0)$
.
We will need a stronger version of the previous decomposition (2.10) which can be understood as a slice theorem. Knowledge of it goes back to [Reference EbinEbi68, Reference Ebin, Chern and SmaleEbi70]; see also [Reference Guillarmou and LefeuvreGL19, Lemma 4.1] for a short proof in the
$C^{k,\alpha }$
category.
Lemma 2.4. Let k be an integer greater than or equal to
$2$
and
$\alpha \in (0,1)$
, let
$g_0$
be a
$C^{k+3,\alpha }$
metric with Anosov geodesic flow. There exists a neighborhood
$\mathcal {U} \subset \mathcal {M}^{k,\alpha }$
of
$g_0$
in the
$C^{k,\alpha }$
-topology such that for any
$g \in \mathcal {U}$
, there exists a unique
$C^{k+1,\alpha }$
-diffeomorphism
$\psi $
such that
$\psi ^*g$
is solenoidal with respect to
$g_0$
. Moreover, the following map is
$C^2$
:
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu38.png?pub-status=live)
Remark 2.5. The previous lemma is not stated exactly this way in [Reference Guillarmou and LefeuvreGL19, Lemma 4.1]. Indeed, the proof assumes that
$g_0$
is smooth and fixed. However, inspecting the proof, it readily applies to
$g_0\in C^{k+3,\alpha }$
and the implicit function theorem used in that proof shows the regularity of
$\psi $
with respect to
$g_0$
. We do not include the proof of these details in order not to burden the discussion.
We also see that we need to use to
$C^{k,\alpha }$
regularity for
$\alpha \neq 0,1$
instead of
$C^k$
: this is due to the fact that the pseudodifferential operator inverting the linearization
$D^*_{g_0}D_{g_0}$
that arises naturally in the proof of this lemma (see [Reference Guillarmou and LefeuvreGL19, Lemma 4.1]) acts on these spaces but on
$C^k$
, for
$k \in \mathbb {N}$
. Instead, one would have to resort to Zygmund spaces
$C^k_*$
. We refer to [Reference TaylorTay91, Appendix A] for further details.
2.5 Thermodynamic formalism
Let f be a Hölder-continuous function on
$S_{g_0}M$
. We recall that its pressure [Reference WaltersWal82, Theorem 9.10] is defined by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqn14.png?pub-status=live)
where
$\mathfrak {M}_{\textrm{inv}}$
denotes the set of invariant (by the flow
$\varphi ^{g_0}$
) Borel probability measures and
${\textbf {h}}_\mu (\varphi _1^{g_0})$
is the metric entropy of the flow
$\varphi _1^{g_0}$
at time
$1$
. It is actually sufficient to restrict the
$\textrm{sup}$
to ergodic measures
$\mathfrak {M}_{\textrm{inv,erg}}$
[Reference WaltersWal82, Corollary 9.10.1]. Since the flow is Anosov, the supremum is always achieved for a unique invariant ergodic measure
$\mu _f$
(by [Reference Bowen and RuelleBR75, Theorem 3.3]; see also [Reference Hasselblatt and FisherHF19, Theorem 9.3.4] and the following discussion therein) called the equilibrium state of f, and
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqn15.png?pub-status=live)
see [Reference Hasselblatt and FisherHF19, Theorem 9.3.16]. The measure
$\mu _f$
is also mixing and positive on open sets, which rules out the possibility of a finite combination of Dirac measures supported on a finite number of closed orbits. Moreover,
$\mu _f$
can be written as an infinite weighted sum of Dirac masses
$\delta _{g_0}(c_j)$
supported over the geodesics
$\gamma _{g_0}(c_j)$
, where
$c_j\in \mathcal {C}$
are the primitive classes (see [Reference Parry and AlexanderPar88] for the case
${\textbf {P}}(f)\geq 0$
or [Reference Paulin, Pollicott and SchapiraPPS15, Theorem 9.17] for the general case). For example, when
${\textbf {P}}(f)\geq 0$
,
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqn16.png?pub-status=live)
where
$N(T,f):=\sum _{j, L_{g_0}(c_j)\in [T,T+1]}L_{g_0}(c_j)e^{\int _{\gamma _{g_0}(c_j)}f}$
. When
$f = 0$
, this is the measure of maximal entropy, also called the Bowen–Margulis measure
$\mu _{g_0}^{\textrm{BM}}$
; in that case
$\textbf{P}(0) = {\textbf {h}}_{\textrm{top}}(\varphi _1^{g_0})$
is the topological entropy of the flow. When
$f =-J_{g_0}^u$
, where
$J_{g_0}^u : x \mapsto \partial _t (|\det d\varphi ^g_t(x)|_{E_u(x)})|_{t=0}$
is the unstable Jacobian, we obtain the Liouville measure
$\mu _{g_0}^{\textrm{L}}$
induced by the metric
$g_0$
; in that case,
$\textbf{P}(-J_{g_0}^u) = 0$
. If we fix an exponent of Hölder regularity
$\nu> 0$
, then the map
$C^\nu (S_{g_0}M) \ni f \mapsto \textbf{P}(f)$
is real analytic (see [Reference RuelleRue04, Corollary 7.10] for discrete systems and [Reference Parry and PollicottPP90, Proposition 4.7] for flows).
2.6 Geodesic stretch
We refer to [Reference Croke and FathiCF90, Reference KnieperKni95] for the original definition of this notion.
2.6.1 Structural stability and time reparametrization
We fix a smooth metric
$g_0 \in \mathcal {M}$
with Anosov geodesic flow and we view the geodesic flow and vector fields of any metric g close to
$g_0$
as living on the unit tangent bundle
$S_{g_0}M$
of
$g_0$
by simply pulling them back by the diffeomorphism
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu39.png?pub-status=live)
We fix some constant
$k \geq 2$
and
$\alpha \in (0,1)$
. There exist a regularity parameter
$\nu>0$
and a neighborhood
$\mathcal {U} \subset \mathcal {M}^{k,\alpha }$
of
$g_0$
such that, by the structural stability theorem ([Reference de la Llave, Marco and MoriyóndlLMM86, Appendix A] or [Reference Katok, Knieper, Pollicott and WeissKKPW89, Proposition 2.2] for the Hölder regularity case), for any
$g \in \mathcal {U}$
, there exists a
$C^\nu $
Hölder homeomorphism
$\psi _g : S_{g_0}M \rightarrow S_{g_0}M$
, differentiable in the flow direction, which is an orbit conjugacy that is such that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqn17.png?pub-status=live)
where
$a_g$
is in
$C^{\nu }(S_{g_0}M)$
. Moreover, the map
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu40.png?pub-status=live)
is
$C^{k-2}$
and
$\psi _g$
is homotopic to the identity. For the proof of Theorem 1.3, we will also need the continuity of
$a_g=a_{g_0,g}$
and of its g-derivatives of order
$\ell \leq k-2$
as a function of the base metric
$g_0$
. This continuity follows essentially from the proof of [Reference Katok, Knieper, Pollicott and WeissKKPW89, Proposition 2.2]; we give a proof of this fact in Proposition C.1 in the Appendix.
Note that neither
$a_g$
nor
$\psi _g$
is unique, but
$a_g$
is unique up to a coboundary and in all the following paragraphs; adding a coboundary to
$a_g$
will not affect the results. From (2.14), we obtain that for
$t \in \mathbb {R}$
,
$z \in S_{g_0}M$
,
$\varphi ^g_{\kappa _{a_g}(z,t)} (\psi _g(z)) = \psi _g(\varphi ^{g_0}_{t}(z))$
with
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqn18.png?pub-status=live)
If
$c \in \mathcal {C}$
is a free homotopy class, then
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqn19.png?pub-status=live)
for any
$z \in \gamma _{g_0}(c)$
, the unique
$g_0$
-closed geodesic in c.
2.6.2 Definition of the geodesic stretch
We denote by
$\widetilde {M}$
the universal cover of M. Given a metric
$g \in \mathcal {M}$
on M, we denote by
$\widetilde {g}$
its lift to the universal cover. Given two metrics
$g_1$
and
$g_2$
on M, there exists a constant
$c> 0$
such that
$c^{-1}g_1 \leq g_2 \leq cg_1$
. This implies that any
$\widetilde {g_1}$
-geodesic is a quasi-geodesic for
$\widetilde {g_2}$
. We now assume that the two metrics
$g_1,g_2$
are Anosov on M. The ideal (or visual) boundary
$\partial _\infty \widetilde {M}$
is independent of the choice of g and is naturally endowed with the structure of a topological manifold (see Appendix B) whose regularity inherits that of the foliation (that is, it is at least Hölder continuous and is
$C^{2-\varepsilon }$
for any
$\varepsilon> 0$
on negatively curved surfaces by [Reference Hurder and KatokHK90]). In negative curvature, we refer to [Reference Bridson and HaefligerBH99, Ch. H.3] and [Reference Knieper, Hasselblat and KatokKni02] for further details. For the general Anosov case, we refer to [Reference KnieperKni12] and Appendix B of the present paper.
We denote by
$\mathcal {G}_g := S_{\widetilde {g}}\widetilde {M}/\sim $
(where
$z \sim z^{\prime }$
if and only if there exists a time
$t \in \mathbb {R}$
such that
$\varphi _t(z) =z^{\prime }$
) the set of g-geodesics on
$\widetilde {M}$
: this is a smooth
$2n$
-dimensional manifold. Moreover, there exists a Hölder-continuous homeomorphism
$\Phi _g : \mathcal {G}_g \rightarrow \partial _\infty \widetilde {M} \times \partial _\infty \widetilde {M} \setminus \Delta $
, where
$\Delta $
is the diagonal in
$\partial _\infty \widetilde {M} \times \partial _\infty \widetilde {M}$
. Given a point
$z \in S_{\widetilde {g}}\widetilde {M}$
, we will denote by
$z_+, z_- \in \partial _\infty \widetilde {M}$
the points (in the future and in the past, respectively) on the boundary at infinity of the geodesic generated by z.
We now consider a fixed metric
$g_0$
on M and a metric g in a neighborhood of
$g_0$
. If
$\psi _g$
denotes an orbit equivalence between the two geodesic flows, then
$\psi _g$
induces a homeomorphism
$\Psi _g : \mathcal {G}_{g_0} \rightarrow \mathcal {G}_g$
. The map
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu41.png?pub-status=live)
is nothing more than the identity.
Given
$z = (x,v) \in S_{g_0}M$
, we denote by
$c_{g_0}(z) : t \mapsto c_{g_0}(z,t)\in M$
the unique geodesic (for the sake of simplicity, we identify the geodesic and its arc-length parametrization) such that
$c_{g_0}(z,0)=x, \dot {c}_{g_0}(z,0)=v$
. We consider
$\widetilde {c}_{g_0}(z)$
, a lift of
$c_{g_0}(z)$
to the universal cover
$\widetilde {M}$
, and introduce the function
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu42.png?pub-status=live)
which computes the
$\widetilde {g}$
-distance between the endpoints of the
$\widetilde {g_0}$
-geodesic joining
$\widetilde {c}_{g_0}(z,0)$
to
$\widetilde {c}_{g_0}(z,t)$
. It is an immediate consequence of the triangle inequality that
$(z,t) \mapsto b(z,t)$
is a subadditive cocycle for the geodesic flow
$\varphi ^{g_0}$
, that is,
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu43.png?pub-status=live)
As a consequence, by the subadditive ergodic theorem (see [Reference WaltersWal82, Theorem 10.1], for instance), we obtain the following lemma.
Lemma 2.6. Let
$\mu $
be an invariant probability measure for the flow
$\varphi _t^{g_0}$
. Then the quantity
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu44.png?pub-status=live)
exists for
$\mu $
-almost every
$z \in S_{g_0}M$
,
$I_\mu (g_0,g,\cdot ) \in L^1(S_{g_0}M,d\mu )$
, and this function is invariant by the flow
$\varphi _t^{g_0}$
.
We define the geodesic stretch of the metric g, relative to the metric
$g_0$
, with respect to the measure
$\mu $
by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu45.png?pub-status=live)
When the measure
$\mu $
in the previous definition is ergodic, the function
$I_\mu (g_0,g,\cdot )$
is thus (
$\mu $
-almost everywhere) equal to the constant
$I_\mu (g_0,g)$
. We recall that
$\delta _{g_0}(c)$
is the normalized measure supported on
$\gamma _{g_0}(c)$
, that is,
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu46.png?pub-status=live)
We can actually describe the stretch using the time reparametrization
$a_g$
.
Lemma 2.7. Let
$\mu $
be an ergodic invariant measure with respect to the flow
$\varphi _t^{g_0}$
. Then,
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu47.png?pub-status=live)
where
$(c_j)_{j \geq 0} \in \mathcal {C}^{\mathbb {N}}$
is such that (the existence of
$c_j$
follows from [Reference SigmundSig72, Theorem 1])
$\delta _{g_0}(c_j) \rightharpoonup _{j \rightarrow +\infty } \mu $
.
Proof We first prove the left-hand equality. Let
$\widetilde {M}$
be the universal covering of M and
$\Gamma $
the group of deck transformations. Denote as above by
$\widetilde {\psi }_{g} : S_{\widetilde {g}_0}\widetilde {M} \rightarrow S_{\widetilde {g}} \widetilde {M}$
the lift of the conjugacy between the geodesic flow of the metrics
$ \widetilde g$
and
$\widetilde g_0$
. Then, for all
$\gamma \in \Gamma $
,
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu48.png?pub-status=live)
If
$\pi : T\widetilde {M} \to \widetilde {M}$
is the canonical projection the function
$d_{\widetilde {g}}(\pi (\widetilde {\psi }_g(z)), \pi (z))$
is
$\Gamma $
-invariant. This follows since
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu49.png?pub-status=live)
Hence, by the compactness of M and the continuity of
$d_{\widetilde {g}}(\pi (\widetilde {\psi }_g(z)), \pi (z))$
there is a constant
$C>0$
such that
$d_{\widetilde {g}}(\pi (\widetilde {\psi }_g(z)), \pi (z)) \le C$
for all
$z\in S\widetilde {M}$
. Using the triangle inequality, we obtain
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu50.png?pub-status=live)
This implies, using (2.15), that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu51.png?pub-status=live)
for
$\mu $
-almost every
$z \in S_{g_0}M$
, by the Birkhoff ergodic theorem [Reference WaltersWal82, Theorem 1.14]. By (2.16) we also have
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu52.png?pub-status=live)
thus the proof is complete.
As a consequence, we immediately obtain the following corollary.
Corollary 2.8. Let g belong to a fixed neighborhood
$\mathcal {U}$
of
$g_0$
in
$\mathcal {M}^{k,\alpha }$
, and assume that for any sequence of primitive free homotopy classes
$(c_j)_{j \geq 0} \in \mathcal {C}^{\mathbb {N}}$
such that
$L_{g_0}(c_j) \rightarrow \infty $
, we have
$\lim _{j\to \infty }L_{g}(c_j)/L_{g_0}(c_j)= 1$
. Then, for any equilibrium state
$\mu _f$
with respect to
$\varphi _t^{g_0}$
associated to some Hölder function f, we have
$I_{\mu _f}(g_0,g)=1$
.
Combining this with the results of [Reference Guillarmou and LefeuvreGL19, Theorem 1], namely the local rigidity of the marked length spectrum, we also easily obtain the following theorem.
Theorem 2.9. Let
$(M,g_0)$
be a smooth Riemannian n-dimensional manifold with Anosov geodesic flow, topological entropy
${\textbf {h}}_{\textrm{top}}(g_0) = 1$
, and assume that its curvature is non-positive if
$n \geq 3$
. Then there exists
$k \in \mathbb {N}$
large enough, depending only on n,
$\varepsilon> 0$
small enough such that the following statement holds: there is
$C>0$
depending on
$g_0$
so that, for each
$g \in C^{k}(M;S^2T^*M)$
with
$\|g-g_0\|_{C^k} \leq \varepsilon $
, if
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu53.png?pub-status=live)
for some sequence
$(c_j)_{j \in \mathbb {N}}$
of primitive free homotopy classes such that
$\delta _{g_0}(c_j) \rightharpoonup _{j \rightarrow +\infty } \mu ^{\textrm{BM}}_{g_0}$
, then g is isometric to
$g_0$
.
Proof Given a metric g, we have by [Reference KnieperKni95, Theorem 1.2] (in [Reference KnieperKni95] the metric is assumed to be negatively curved, but the argument applies also for Anosov flows, as is shown in [Reference Bridgeman, Canary, Labourie and SambarinoBCLS15, Proposition 3.8]: it corresponds to Proposition 3.10 below in the case
$f:=1$
and
$f^{\prime }=a_g$
) that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqn20.png?pub-status=live)
with equality if and only if
$\varphi ^{g_0}$
and
$\varphi ^g$
are, up to a scaling, time-preserving conjugate, that is, there exists a homeomorphism
$\psi $
such that
$\psi \circ \varphi _{g_0}^{ct} = \varphi _g^t \circ \psi $
with
$c:= {\textbf { h}}_{\text {top}}(g)/{\textbf {h}}_{\text {top}}(g_0)$
.
In particular, restricting to metrics with entropy
$1$
, we obtain that
$I_{\mu ^{\textrm{BM}}_{g_0}}(g_0,g) \geq 1$
with equality if and only if the geodesic flows are conjugate, that is, if and only if
$L_g = L_{g_0}$
(by the Livsic theorem). As a consequence, given
$g_0, g$
with entropy
$1$
such that
$L_{g}(c_j)/L_{g_0}(c_j) \rightarrow _{j \rightarrow +\infty } 1$
for some sequence
$\delta _{g_0}(c_j) \rightharpoonup _{j \rightarrow +\infty } \mu ^{\textrm{BM}}_{g_0}$
, we obtain that
$I_{\mu ^{\textrm{BM}}_{g_0}}(g_0,g)=1$
, hence
$L_g=L_{g_0}$
. If
$k \in \mathbb {N}$
was chosen large enough at the beginning, we can then conclude by the local rigidity of the marked length spectrum [Reference Guillarmou and LefeuvreGL19, Theorem 1].
In Theorem 2.9, we assume that
$g_0$
has entropy
$1$
. This is actually a harmless assumption in so far as the same result holds true on metrics of constant topological entropy
${\textbf {h}}_{{\textrm{top}}}(g)=\unicode{x3bb}>0$
. Recall that by considering
$\unicode{x3bb} ^2 g_0$
for some constant
$\unicode{x3bb}> 0$
, the entropy scales as
${\textbf { h}}_{\textrm{top}}(\unicode{x3bb} ^2 g_0) = {\textbf {h}}_{\textrm{top}}(g_0)/\unicode{x3bb} $
[Reference PaternainPat99, Lemma 3.23] and we can thus always reduce to the previous case
${\textbf {h}}_{{\textrm{top}}}(g_0) = 1$
. We also observe that the previous theorem implies the local rigidity of the marked length spectrum: if
$L_g=L_{g_0}$
, then
${\textbf {h}}_{\text {top}}(g_0)={\textbf {h}}_{\text {top}}(g)$
because the topological entropy
${\textbf {h}}_{\text {top}}(g)$
is the first pole of the Ruelle zeta function [Reference Parry and PollicottPP90, Theorem 9.1]
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu54.png?pub-status=live)
We can then apply Theorem 2.9 to deduce that g is isometric to
$g_0$
. We will provide an alternative proof of this fact in the next section without using the proof of [Reference Guillarmou and LefeuvreGL19].
3 A functional on the space of metrics
Given a metric g in a
$C^{k,\alpha }$
-neighborhood
$\mathcal {U}$
of
$g_0$
, we define the potential
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqn21.png?pub-status=live)
for some
$\nu>0$
. We remark that
$\mathcal {U} \ni g \mapsto V_g \in C^\nu (S_{g_0}M)$
is
$C^{k-2}$
and, for
$g=g_0$
,
$V_{g_0} = J_{g_0}^{u}$
. Consider the map
$\psi : \mathcal {M}^{k,\alpha }\to \mathbb {R}$
, defined for
$g_0$
a fixed smooth metric with Anosov geodesic flow, by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqn22.png?pub-status=live)
We also define the maps
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqn23.png?pub-status=live)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqn24.png?pub-status=live)
satisfying
$\Psi (g)=F(g)+\Phi (g)$
. We note that
$\Psi ,\Phi , F$
are
$C^{k-2}$
by [Reference ContrerasCon92]. We also make the following observation: since
${\textbf {P}}(-J^u_{g_0})=0$
and
$a_{g_0}$
is cohomologous to
$1$
, we have
$\Psi (g_0)=0$
and
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqn25.png?pub-status=live)
by using the variational definition (2.11) of the pressure. This shows that, for all
$g\in \mathcal {M}^{k,\alpha }$
,
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu55.png?pub-status=live)
Moreover,
$\Psi (g)=0$
if and only if the inequality (3.5) becomes an equality, which means that
$\mu _{g_0}^L$
is the equilibrium measure of
$-J^u_{g_0}+1-a_g$
. Since
$\mu _{g_0}^L$
is also the equilibrium measure associated to
$-J^u_{g_0}$
, we conclude by (2.12) that
$1-a_g$
is cohomologous to a constant, or equivalently
$a_g$
is cohomologous to a constant. We have thus shown the following lemma.
Lemma 3.1. The map
$\Psi $
satisfies
$\Psi (g)\geq \Psi (g_0)=0$
, and
$\Psi (g)=\Psi (g_0)=0$
if and only if
$a_g$
is cohomologous to a constant, or equivalently
$L_g=\unicode{x3bb} L_{g_0}$
for some
$\unicode{x3bb}>0$
.
The proof of Theorem 1.2 will be a consequence of the fact that Taylor expansion of
$\Psi $
at
$g=g_0$
has leading term given by the Hessian, which turns out to be the variance operator
$\Pi _2$
studied before.
3.1 The proof of Theorem 1.2
In what follows, we will compute the derivatives of the map
$\Psi ,\Phi ,F$
. As mentioned earlier, they are
$C^{k-2}$
by [Reference ContrerasCon92, Theorem C], and explicit computations of their derivatives can be found in [Reference Parry and PollicottPP90, Proposition 4.10] (subshift case) and [Reference Katok, Knieper, Pollicott and WeissKKPW90, Reference Katok, Knieper and WeissKKW91] (topological entropy case). The first step in the proof is the following proposition.
Proposition 3.2. The non-negative functional
$\Psi :\mathcal {M}^{k,\alpha }\to \mathbb {R}^+$
defined in (3.2) satisfies the following property: there exist a neighborhood
$\mathcal {U}$
of
$g_0$
in
$C^{5,\alpha }(M,S^2T^*M)$
and a constant
$C_{g_0}$
depending on
$g_0$
such that, for all
$g\in \mathcal {U}$
,
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu56.png?pub-status=live)
Proof We shall compute the Taylor expansion of
$\Psi $
at
$g=g_0$
to second order. By [Reference Parry and PollicottPP90, Proposition 4.10], we have, for
$h\in T_{g}\mathcal {M}^{k,\alpha }$
,
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu57.png?pub-status=live)
where
$m_g$
is the equilibrium measure of
$-V_g$
. In particular, observe that for
$g=g_0$
, we have
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqn26.png?pub-status=live)
since
$m_{g_0} = \mu ^{\textrm{L}}_{g_0}$
. Next, we get, for
$h\in T_{g_0}\mathcal {M}^{k,\alpha }$
,
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqn27.png?pub-status=live)
thus
$d\Psi _{g_0}.h=0$
for all
$h\in T_{g_0}\mathcal {M}^{k,\alpha }$
.
Let us next compute the second derivative
$d^2\Psi _{g_0}(h,h)$
. First, we have
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu58.png?pub-status=live)
Then, by [Reference Parry and PollicottPP90, Proposition 4.11] we know that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqn28.png?pub-status=live)
where
$\operatorname {\textrm{Var}}_{\mu ^{\textrm{L}}_{g_0}}(h)$
is the variance defined in (2.5), equal to
$\langle \Pi ^{g_0} h,h\rangle _{L^2}$
by (2.6) and
${\Pi ^{g_0}1=0}$
. Therefore,
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu59.png?pub-status=live)
All together, we finally get
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu60.png?pub-status=live)
To conclude, we claim in Lemma 3.3 below that
$da_{g_0}.h-\tfrac {1}{2}\pi _2^*h$
is a coboundary, so that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu61.png?pub-status=live)
The statement of the proposition is then simply the Taylor expansion of
$\Psi (g)$
at
$g=g_0$
, with
$h=g-g_0$
. (We need the map to be
$C^3$
for the Taylor expansion, hence the need for the
$C^{5,\alpha }$
regularity since we lose two derivatives as mentioned at the beginning of §3.)
Lemma 3.3. Consider a smooth deformation
$(g_\unicode{x3bb} )_{\unicode{x3bb} \in (-1,1)}$
of
$g_0$
inside
$\mathcal {M}^{k,\alpha }$
. Then there exists a Hölder-continuous function
$f : S_{g_0}M \rightarrow \mathbb {R}$
such that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu62.png?pub-status=live)
Proof Let c be a fixed free homotopy class, and let
$\gamma _0 \in c$
be the unique closed
$g_0$
-geodesic in the class c, which we parametrize by unit-speed
$z_0 : [0,\ell _{g_0}(\gamma _0)]\rightarrow S_{g_0}M$
. We define
$z_\unicode{x3bb} (s)=\psi _\unicode{x3bb} (z_0(s)) = (\alpha _\unicode{x3bb} (s),\dot {\alpha }_\unicode{x3bb} (s))$
(the dot is the derivative with respect to s), where
$\psi _\unicode{x3bb} $
is the conjugacy between
$g_\unicode{x3bb} $
and
$g_0$
: this gives a non-unit-speed parametrization of
$\gamma _\unicode{x3bb} $
, the unique closed
$g_\unicode{x3bb} $
-geodesic in c. We recall that
$\pi : TM \rightarrow M$
is the projection. Using (2.14), we obtain
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu63.png?pub-status=live)
Since
$s \mapsto \alpha _0(s)$
is a unit-speed geodesic for
$g_0$
, it is a critical point of the energy functional (with respect to
$g_0$
). Thus, by differentiating the previous identity with respect to
$\unicode{x3bb} $
and evaluating at
$\unicode{x3bb} =0$
, we obtain
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu64.png?pub-status=live)
As a consequence,
$\pi _2^*(\partial _\unicode{x3bb} g_\unicode{x3bb} |_{\unicode{x3bb} =0}) - 2 \partial _\unicode{x3bb} a_\unicode{x3bb} |_{\unicode{x3bb} =0}$
is a Hölder-continuous function in the kernel of the X-ray transform: by the usual Livsic theorem, there exists a function f (with the same Hölder regularity), differentiable in the flow direction, such that
$\pi _2^*(\partial _\unicode{x3bb} g_\unicode{x3bb} |_{\unicode{x3bb} =0}) - 2 \partial _\unicode{x3bb} a_\unicode{x3bb} |_{\unicode{x3bb} =0} = X_{g_0}f$
.
As a corollary, we obtain the following result.
Corollary 3.4. For
$k\geq 5$
,
$\alpha \in (0,1)$
, there exist a neighborhood
$\mathcal {U}$
of
$g_0$
in
$C^{k,\alpha }(M;S^2T^*M)$
and constants
$C_{g_0},C_{g_0}^{\prime }>0$
depending on
$g_0$
such that, for all
$g\in \mathcal {U}$
,
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu65.png?pub-status=live)
There exist a neighborhood
$\mathcal {U}^{\prime }$
of
$g_0$
in
$C^{k,\alpha }(M;S^2T^*M)$
and a constant
$C_{g_0}^{\prime \prime }>0$
depending on
$g_0$
such that, for all
$g\in \mathcal {U}^{\prime }$
, there is a diffeomorphism
$\psi \in C^{k+1,\alpha }(M)$
such that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu66.png?pub-status=live)
Proof The first inequality follows from Proposition 3.2 and Lemma 2.2. For the second inequality, we apply the first inequality to
$\psi ^*g$
, where
$\psi $
is the diffeomorphism obtained from Lemma 2.4, and we use that
$\Psi (\psi ^*g)=\Psi (g)$
.
The next step is to control the term
$\langle (\psi ^*g-g_0),g_0\rangle _{L^2}$
by the geodesic stretch. We will show the following proposition.
Proposition 3.5. There is
$k \in \mathbb {N}$
large enough, depending only on
$n=\dim M$
, such that if
$g_0$
is smooth with Anosov geodesic flow, and non-positive curvature in the case
$n>2$
, there exist
$C_{g_0}>0$
and
$C_n>0$
, an open neighborhood
$\mathcal {U}$
in
$C^{k,\alpha }(M;S^2T^*M)$
of
$g_0$
, such that. for each
$g\in \mathcal {U}$
, there is a diffeomorphism
$\psi $
satisfying
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu67.png?pub-status=live)
Here
$C_{g_0}$
depends on
$g_0$
and
$C_n$
on n only.
Proof We write the Taylor expansion of
$\Phi (\psi ^*g)=\Phi (g)=I_{\mu _{g_0}^L}(g_0,g)-1$
at
$g=g_0$
: by Lemma 3.6 and Lemma 3.3,
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu68.png?pub-status=live)
for some
$C_n>0$
depending only on
$n=\dim M$
. Then
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu69.png?pub-status=live)
Combining with Corollary 3.4, we obtain
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqn29.png?pub-status=live)
if
$\|\psi ^*g-g_0\|_{C^{5,\alpha }}$
is small enough, which is the case if
$\|g-g_0\|_{C^{5,\alpha }}$
is small enough by Lemma 2.4. To obtain the first inequality of Proposition 3.5, we apply Sobolev embedding and interpolation estimates (the interpolation estimate
$\|u\|_{H^c}\leq \|u\|_{H^a}^{t}\|u\|_{H^b}^{1-t}$
for
$c=ta+(1-t)b$
is obtained by applying the Hadamard three-line theorem to the holomorphic function
$s\mapsto \sum _{j}(1+\unicode{x3bb} _j)^{s}\langle u,e_j\rangle _{L^2}^2$
on
${\textrm{Re}}(s)\in [a,b]$
, where
$e_j$
is an orthonormal basis of eigenfunctions of any positive elliptic self-adjoint differential operator of order
$2$
on symmetric tensors and
$\unicode{x3bb} _j$
gives the corresponding eigenvalues) [Reference TaylorTay96, Ch. 4] and get, for some constants
$c_{g_0}>0,c^{\prime }_{g_0}>0$
depending on
$g_0$
only,
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu70.png?pub-status=live)
if
$k>({3}/{2})n+16+3\alpha $
and
$\alpha ^{\prime }>\alpha $
. This means that if
$\|\psi ^*g-g_0\|_{H^k}$
is small enough, depending on the constants
$C_{g_0}, C_{g_0}^{\prime \prime },c_{g_0},c^{\prime }_{g_0}$
, one can absorb the
$\|\psi ^*g-g_0\|^3_{C^{5,\alpha }}$
term of (3.9) into the left-hand side and get the first inequality of Proposition 3.5. The smallness of
$\|\psi ^*g-g_0\|_{H^k}$
is implied by the smallness of
$\|g-g_0\|_{C^{k,\alpha }}$
by Lemma 2.4. The same exact argument applies by replacing
$\Phi (g)$
by
$F(g)$
using that
$dF_{g_0}=-d\phi _{g_0}$
; this proves the second inequality of Proposition 3.5. The last inequality is similar since
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu71.png?pub-status=live)
for
$h:=g-g_0$
. The proof is complete.
To conclude the proof of Theorem 1.2, we need to estimate
$\Phi (g)$
and
$F(g)$
in terms of
$\mathcal {L}_\pm (g)$
. Recall that (see [Reference Paulin, Pollicott and SchapiraPPS15, Corollary 9.17])
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu72.png?pub-status=live)
Thus, if we order
$\mathcal {C}=(c_j)_{j\in \mathbb {N}}$
by the lengths (that is,
$L_{g_0}(c_j)\geq L_{g_0}(c_{j-1})$
), and we define
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu73.png?pub-status=live)
we see that for all
$\delta>0$
small, there is
$T_0>0$
large so that, for all j with
$L_{g_0}(c_j)\in [T,T+1]$
with
$T\geq T_0$
,
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu74.png?pub-status=live)
We deduce, using
${\textbf {P}}(-V_{g_0})=0$
, that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu75.png?pub-status=live)
Since
$\delta>0$
is arbitrarily small, we obtain
$|F(g)|\leq \max (|\mathcal {L}_+(g)|,|\mathcal {L}_-(g)|)$
. Similarly, Lemma 2.7 shows that
$|\Phi (g)|\leq \max (|\mathcal {L}_+(g)|,|\mathcal {L}_-(g)|)$
. So the proof of Theorem 1.2 is complete by combining these bounds with Proposition 3.5 (the right-hand side in the first and second inequalities of Proposition 3.5 being
$F(g)+\Phi (g)+C_n \Phi (g)^2$
and
$F(g)+\Psi (g)+C_nF(g)^2$
).
3.2 A submanifold of the space of metrics
It is quite natural to describe the stretch functional
$\Phi $
on the space
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqn30.png?pub-status=live)
and on
$\mathcal {N}^{k,\alpha }_{\textrm{sol}} := \mathcal {N}^{k,\alpha } \cap \ker D^*_{g_0}$
. Indeed, as we shall see, this becomes a strictly convex functional near
$g_0\in \mathcal {N}_{\textrm{sol}}^{k,\alpha }$
when restricted to
$\mathcal {N}_{\textrm{sol}}^{k,\alpha }$
. It is possible that the map is strictly convex globally on
$\mathcal {N}_{\textrm{sol}}^{k,\alpha }$
, in which case that would prove the global rigidity of the marked length spectrum.
Given
$g \in \mathcal {N}^{k,\alpha }$
, we denote by
$m_g$
the unique equilibrium state for the potential
$V_g$
. We will also write
$\mathcal {N}$
for the case where
$k=\infty $
. First we check that these are (infinite-dimensional) manifolds.
Lemma 3.6. There exists a neighborhood
$\mathcal {U} \subset \mathcal {M}^{k,\alpha }$
of
$g_0$
such that
$\mathcal {N}^{k,\alpha } \cap \mathcal {U}$
is a codimension-one
$C^{k-2}$
-submanifold of
$\mathcal {U}$
and
$\mathcal {N}^{k,\alpha }_{\textrm{sol}} \cap \mathcal {U}$
is a
$C^{k-2}$
-submanifold of
$\mathcal {U}$
. Similarly, there is
$\mathcal {U}\subset \mathcal {M}$
an open neighborhood so that
$\mathcal {N} \cap \mathcal {U}$
is a Fréchet submanifold of
$\mathcal {M}$
.
Proof To prove this lemma, we will use the notion of differential calculus on Banach manifolds as it is stated in [Reference ZeidlerZei88, Ch. 73]. Note that
$\mathcal {M}^{k,\alpha }$
is a smooth Banach manifold and
$\mathcal {N}^{k,\alpha } \subset \mathcal {M}^{k,\alpha }$
is defined by the implicit equation
$F(g)=0$
for
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqn31.png?pub-status=live)
The map F being
$C^{k-2}$
, we need only prove that
$dF_{g_0}$
does not vanish by [Reference ZeidlerZei88, Theorem 73.C]. This will immediately give that
$T_{g_0}\mathcal {N}^{k,\alpha }=\ker dF_{g_0}$
. In order to do so, we need a deformation lemma. For the sake of simplicity, we write the objects
$\cdot _{\unicode{x3bb} }$
instead of
$\cdot _{g_\unicode{x3bb} }$
.
We can now complete the proof of Lemma 3.6. We first prove the first part concerning
$\mathcal {N}^{k,\alpha }$
. Recall formula (3.6) for
$dF_{g_0}$
. Using Lemma 3.3, we obtain
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqn32.png?pub-status=live)
for some constant
$C_n> 0$
depending on n. This is obviously surjective and we also obtain
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu76.png?pub-status=live)
where the orthogonal is understood with respect to the
$L^2$
-scalar product.
We now deal with
$\mathcal {N}^{k,\alpha }_{\textrm{sol}}$
. First observe that
$\ker D^*_{g_0}$
is a closed linear subspace of
$\mathcal {M}^{k,\alpha }$
and thus a smooth submanifold of
$\mathcal {M}^{k,\alpha }$
. By [Reference ZeidlerZei88, Corollary 73.50], it is sufficient to prove that
$\ker D^*_{g_0}$
and
$\mathcal {N}^{k,\alpha }$
are transverse at
$g_0$
. But observe that
$g_0 \in \ker D^*_{g_0} \simeq T_{g_0} \ker D^*_{g_0}$
and thus
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu77.png?pub-status=live)
showing transversality.
The case of
$\mathcal {N}$
follows directly from the Nash–Moser theorem: F is obviously a smooth tame map from
$C^\infty (M;S^2T^*M)$
to
$\mathbb {R}$
; moreover,
$dF_g$
has a right inverse
$H_g$
since
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu78.png?pub-status=live)
where we use the fact that
$2da_g.g$
is cohomologous to
$a_g$
. This can be seen by differentiating
$L_g(c)=\int _{\gamma _{g_0}(c)}a_g$
and applying the Livsic theorem. In particular,the right inverse is given by
$H_g.1:=-2g/I_{m_g}(g_0,g)$
. The family of right inverses
$g\mapsto H_g$
is smooth since
$g\mapsto a_g$
and
$g\mapsto m_g$
are smooth by [Reference ContrerasCon92, Theorem C], and it is clearly also tame, thus we can apply directly [Reference HamiltonHam82, Theorem 1.1.3, pp. 172] to deduce that F has a smooth tame right inverse, which shows that
$\mathcal {N}$
is a Fréchet submanifold.
We remark that if
$L_g=L_{g_0}$
, then
$a_g$
is cohomologous to
$1$
, so
${\textbf { P}}(-V_{g})={\textbf {P}}(-V_{g_0})=0$
in that case, which means that
$g\in \mathcal {N}^{k,\alpha }$
. From the second inequality in Proposition 3.5, we obtain the following corollary.
Corollary 3.7. Let
$g_0$
be a smooth metric with Anosov geodesic flow, with non-positive curvature if
$n>2$
. There exist
$C_{g_0}>0$
and a neighborhood
$\mathcal {U}\subset \mathcal {N}^{k,\alpha }$
such that, for all
$g\in \mathcal {U}$
, there is a diffeomorphism
$\psi \in \mathcal {D}_0^{k+1,\alpha }$
so that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu79.png?pub-status=live)
As suggested by this estimate, the functional
$\Phi $
turns out to be strictly convex near
$g_0$
when restricted on
$\mathcal {N}_{\textrm{sol}}^{k,\alpha }$
. First, we have, for
$h\in T_{g_0}\mathcal {N}^{k,\alpha }$
,
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu80.png?pub-status=live)
so that
$\Phi :\mathcal {N}^{k,\alpha }\to \mathbb {R}$
has a critical point at
$g_0$
. For the second derivative at
$g_0$
, the same computation as in the previous section easily gives the following result.
Lemma 3.8. The map
$\Phi :\mathcal {N}^{k,\alpha }_{\textrm{sol}} \to \mathbb {R}$
is strictly convex at
$g_0$
, and there is
$C>0$
such that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu81.png?pub-status=live)
for all
$h \in T_{g_0}\mathcal {N}^{k,\alpha }_{\textrm{sol}}$
.
Proof The proof follows exactly that of Proposition 3.2, using
$T_{g_0} \mathcal {N}^{k,\alpha } = (\mathbb {R} g_0)^\bot $
.
3.3 The pressure metric on the space of negatively curved metrics
The results of this paragraph are stated in negative curvature, but it is very likely that one could relax the assumption to the Anosov case. Again, the only obstruction for the moment is that it is still not known whether the X-ray transform
$I_2$
(hence the operator
$\Pi _2$
) is injective on solenoidal tensors in the Anosov case when
$\textrm{dim}(M) \geq 3$
.
3.3.1 Definition of the pressure metric using the variance
On
$\mathcal {M}^-$
, the cone of smooth negatively curved metrics, we introduce the non-negative symmetric bilinear form
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqn33.png?pub-status=live)
defined for
$g \in \mathcal {M}$
,
$h_j\in T_g \mathcal {M} \simeq C^{\infty }(M; S^2T^*M)$
. It is non-degenerate on
$T_g \mathcal {M} \cap \ker D^*_g$
, namely,
$G_g(h,h) \geq C_g \|h\|^2_{H^{-1/2}}$
by Lemma 2.2, and the constant
$C_g$
turns out to be locally uniform for g near a given metric
$g_0$
. Combining these facts, we obtain the following proposition.
Proposition 3.9. Let
$g_0\in \mathcal {M}^-$
. Then the bilinear form G defined in (3.13) produces a Riemannian metric on the quotient space
$\mathcal {M}^-/\mathcal {D}_0$
near the class
$[g_0]$
, where
$\mathcal {M}^-/\mathcal {D}_0$
is identified with the slice
$\mathcal {S}$
passing through
$g_0$
as in (2.9).
Proof It suffices to show that G is non-degenerate on
$T\mathcal {S}$
. Let
$h\in T_{g}\mathcal {S}$
and assume that
$G_g(h,h)=0$
. We can write
$h=\mathcal {L}_{V}g+h^{\prime }$
where
$D_g^*h^{\prime }=0$
and V is a smooth vector field and
$\mathcal {L}_V$
the Lie derivative with respect to V. By Lemma 2.1 we obtain
$0=G_g(h,h)\geq C\|h^{\prime }\|_{H^{-1/2}}$
. Thus
$h=\mathcal {L}_Vg$
, but we also know that
$T_{g}\mathcal {S}\cap \{\mathcal {L}_Vg ~|~ V\in C^\infty (M;T^*M)\}=\{0\}$
since
$\mathcal {S}$
is a slice. Therefore
$h=0$
.
3.3.2 Definition using the intersection number
We now want to relate the pressure metric previously introduced to some renormalized intersection numbers involving some well-chosen potentials. This will be needed to show that the pressure metric coincides with (a multiple of) the Weil–Petersson metric in the case where M is a surface and one restricts to hyperbolic metrics. This also makes a relation with recent work of [Reference Bridgeman, Canary, Labourie and SambarinoBCLS15].
Let us assume that g is in a fixed
$C^2$
-neighborhood of
$g_0$
. Since
$J^u_{g_0}> 0$
, we obtain that
$V_g = J^u_{g_0} + a_g-1> 0$
if g is close enough to
$g_0$
. By [Reference SambarinoSam14, Lemma 2.4], there exists a unique constant
${\textbf {h}}_{V_g} \in \mathbb {R}$
such that
$\textbf{P}(-{\textbf {h}}_{V_g} V_g)= 0$
. In particular,
$\mathcal {N}$
coincides in a neighborhood of
$g_0$
with the set
$\{ g \in \mathcal {M} ~|~ {\textbf {h}}_{V_g} = 1\}$
. One can express the constant
${\textbf {h}}_{V_g}$
as
${\textbf {h}}_{V_g} = {\textbf {h}}_{\textrm{top}}(\varphi _t^{g_0,V_g})$
, where
$\varphi _t^{g_0,V_g}$
is a time reparametrization of the geodesic flow of
$g_0$
(see [Reference Bridgeman, Canary, Labourie and SambarinoBCLS15, §3.1.1]). More precisely, given a Hölder-continuous positive function
$f \in C^\nu (S_{g_0}M)$
on
$S_{g_0}M$
, we introduce the unique real number
${\textbf {h}}_f$
such that
$\textbf{P}(-{\textbf {h}}_f f)=0$
and we set
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu82.png?pub-status=live)
For a fixed
$z\in S_{g_0}M$
, this is a homeomorphism on
$\mathbb {R}$
and thus allows us to define
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqn34.png?pub-status=live)
We now follow the approach of [Reference Bridgeman, Canary, Labourie and SambarinoBCLS15, §3.4.1]. Given two Hölder-continuous functions
$f,f^{\prime } \in C^\nu (S_{g_0}M)$
such that
$f> 0$
, one can define an intersection number [Reference Bridgeman, Canary, Labourie and SambarinoBCLS15, Eq. (13)]
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu83.png?pub-status=live)
where
$d\mu _{- {\textbf {h}}_f f}$
is the equilibrium measure for the potential
$-{\textbf {h}}_f f$
. We have the following result, which follows from [Reference Bridgeman, Canary, Labourie and SambarinoBCLS15, Proposition 3.8] stated for Anosov flows on compact metric spaces:
Proposition 3.10. (Bridgeman, Canary, Labourie and Sambarino [Reference Bridgeman, Canary, Labourie and SambarinoBCLS15])
Let
$f, f^{\prime } : S_{g_0}M \rightarrow \mathbb {R}_+$
be two Hölder-continuous positive functions. Then
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu84.png?pub-status=live)
with equality if and only if
${\textbf {h}_{\textbf {f}}} f$
and
$\textbf{h}_{\textbf{f}^{\prime }} f^{\prime }$
are cohomologous for the geodesic flow
$\varphi _t^{g_0}$
of
$g_0$
. The quantity
$\textbf{J}_{g_0}(f,f^{\prime })$
is called the renormalized intersection number.
We apply the previous proposition with
$f := J^u_{g_0}$
(then
${\textbf {h}}_{J^u_{g_0}} = 1$
) and
$f^{\prime } := V_g$
. Without assuming that
$g \in \mathcal {N}$
(that is, we do not necessarily assume that
${\textbf {h}}_{V_g}=1$
), we have
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu85.png?pub-status=live)
where
${\textbf {h}}_L(g_0)$
is the entropy of the Liouville measure for
$g_0$
. In the specific case where
$g \in \mathcal {N}$
,
${\textbf {h}}_{V_g}=1$
and we find that
$I_{\mu ^{\textrm{L}}_{g_0}}(g_0,g) \geq 1$
with equality if and only if
$a_g$
is cohomologous to
$1$
, that is, if and only if
$L_g=L_{g_0}$
, or alternatively if and only if
$\varphi ^g$
and
$\varphi ^{g_0}$
are time-preserving conjugate. This computation holds as long as
$J^u_{g_0} + a_g-1> 0$
(which is true in a
$C^2$
-neighborhood of
$g_0$
).
In particular, on
$\mathcal {N}$
, we have the linear relation
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu86.png?pub-status=live)
In the notations of [Reference Bridgeman, Canary, Labourie and SambarinoBCLS15, Proposition 3.11], the second derivative computed for the family
$(g_\unicode{x3bb} )_{\unicode{x3bb} \in (-1,1)} \in \mathcal {N}$
is
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqn35.png?pub-status=live)
and is called the pressure form. When considering a slice transverse to the
$\mathcal {D}_0$
action on
$\mathcal {N}$
, it induces a metric called the pressure metric by Lemma 2.1. To summarize, we have the following lemma.
Lemma 3.11. Given a smooth metric
$g_0$
, the metric
$G_{g_0}$
restricted to
$\mathcal {N}$
can be obtained from the renormalized intersection number by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu87.png?pub-status=live)
where
$(g_\unicode{x3bb} )_{\unicode{x3bb} \in (-1,1)}$
is any family of metrics such that
$g_\unicode{x3bb} \in \mathcal {N}$
and
$\dot {g}_0 = h\in T_{g_0}\mathcal {N}$
.
3.3.3 Link with the Weil–Petersson metric
We now assume that
$M=S$
is an orientable surface of genus at least
$2$
. Let
$\mathcal {T}(S)$
be the Teichmüller space of S. We show that the pressure metric coincides with (a multiple of) the Weil–Petersson metric in restriction to
$\mathcal {T}(S)$
. We fix a hyperbolic metric
$g_0$
. Given
$\eta , \rho \in \mathcal {T}(S)$
and
$g_\eta ,g_\rho $
the associated hyperbolic metrics, since
$\mathcal {T}(S)$
is connected (indeed, a ball in
$\mathbb {C}^{3({\textrm{genus}}(M)-1)}$
) there is topological conjugacy between
$g_\eta ,g_\rho $
and
$g_0$
and one can defined the time rescaling
$a_{g_\eta }$
and
$a_{g_{\rho }}$
by using a path of hyperbolic metrics relating
$g_0$
to
$g_\eta $
or to
$g_\rho $
. The intersection number is defined as
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu88.png?pub-status=live)
where
$[g_\eta ]=\eta , [g_\rho ]=\rho $
and
$\mu _\eta $
is the equilibrium state of
$-{\textbf {h}}_{a_{g_\eta }}a_{g_{\eta }}$
. Note that
${\textbf {h}}_{a_{g_\eta }}={\textbf { h}}_{\textrm{top}}(\varphi ^{g_0,a_\eta }_t)=1$
since
$\varphi ^{g_0,a_\eta }$
is conjugate to the geodesic flow of
$g_\eta $
, which in turn has constant curvature, and by [Reference SambarinoSam14, Lemma 2.4],
$a_{g_\eta }d\mu _{\eta }/\int _{S_{g_0}M}a_{g_\eta }d\mu _{\eta }$
is the measure of maximal entropy of the flow
$\varphi ^{g_0,a_\eta }_t$
, thus also the normalized Liouville measure of
$g_\eta $
(viewed on
$S_{g_0}M$
). This number
$\textbf{I}(\eta ,\rho )$
is in fact independent of
$g_0$
as it can alternatively be written
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu89.png?pub-status=live)
where
$N_T=\sharp \{c\in \mathcal {C} \, | L_{g_\eta }(c)\leq T\}$
(see [Reference Bridgeman, Canary and SambarinoBCS18, Proof of Theorem 4.3]). In particular, taking
$g_0 = g_\eta $
, we have
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu90.png?pub-status=live)
As explained in [Reference Bridgeman, Canary and SambarinoBCS18, Theorem 4.3], up to a normalization constant
$c_0$
depending on the genus only, the Weil–Petersson metric on
$\mathcal {T}(S)$
is equal to
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqn36.png?pub-status=live)
where
$\dot {\eta }_0=h$
and
$(g_{\eta _\unicode{x3bb} })_{\unicode{x3bb} \in (-1,1)}$
is a family of hyperbolic metrics such that
$[g_{\eta _\unicode{x3bb} }] = \eta _\unicode{x3bb} $
,
$\eta =\eta _0 = [g_0]$
. This fact follows from combined works of Thurston, Wolpert [Reference WolpertWol86] and McMullen [Reference McMullenMM08]: the length of a random geodesic
$\gamma $
on
$(S,g_0)$
with respect to
$g_{\eta _\unicode{x3bb} }$
has a local minimum at
$\unicode{x3bb} =0$
and the Hessian is positive definite (Thurston), is equal to the Weil–Petersson norm squared of
$\dot {g}$
(Wolpert [Reference Fathi and FlaminioFF93, Reference WolpertWol86]) and is given by a variance (McMullen [Reference McMullenMM08]); here random means equidistributed with respect to the Liouville measure of
$g_0$
. We can check that the metric G also corresponds to this metric.
Proposition 3.12. The metric G on
$\mathcal {T}(S)$
is a multiple of the Weil–Petersson metric.
Proof This follows directly from (3.15), (3.16) and the fact that
${\textbf {h}}_L(g_\eta )=1$
if
$g_\eta $
has curvature
$-1$
.
Remark 3.13. We notice that the positivity of the metric in the case of Teichmüller space follows only from some convexity argument in finite dimension. In the case of general metrics with negative curvature, the elliptic estimate of Lemma 2.1 on the variance is much less obvious due to the infinite dimensionality of the space. As it turns out, this is the key for the local rigidity in our results.
4 Uniform elliptic estimates on
$\Pi _2$
In this section we prove that the operator
$\Pi _2^g \in \Psi ^{-1}(M;S^2T^*M)$
depends continuously on g. Let
$\mathcal {M}^{\textrm{An}}$
be the space of smooth Riemannian metrics with Anosov geodesic flow.
Proposition 4.1. The map
$\mathcal {M}^{\textrm{An}} \ni g \mapsto \Pi _2^g \in \Psi ^{-1}(M;S^2T^*M)$
is continuous when
$\Psi ^{-1}(M;S^2T^*M)$
is equipped with its topology of Fréchet spaces.
Recall that the Fréchet topology was introduced at the beginning of §2.2. We fix a metric
$g_0$
and we work in a neighborhood
$\mathcal {U}$
of
$g_0$
in the
$C^\infty $
topology. In particular, we will always assume that this neighborhood
$\mathcal {U}$
is small enough that any
$g \in \mathcal {U}$
has an Anosov geodesic flow that is orbit-conjugated to that of
$g_0$
by structural stability. We will also see the geodesic flows
$(\varphi _t^g)_{t \in \mathbb {R}}$
as acting on the unit bundle
$SM:=S_{g_0}M$
for
$g_0$
by using the natural identification
$S_gM\to S_{g_0}M$
obtained by scaling in the fibers. The operator
$\pi _2^*$
associated to g becomes: for
$(x,v)\in S_{g_0}M$
,
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu91.png?pub-status=live)
4.1 The resolvents of
$X_g$
and anisotropic spaces
We first recall the construction of resolvents of
$X_g$
from Faure and Sjöstrand [Reference Faure and SjöstrandFS11] (see also [Reference Dyatlov and ZworskiDZ16]) and, in particular, the version used in Dang et al [Reference Dang, Guillarmou, Rivière and ShenDGRS20] that deals with the continuity with respect to the flow
$X_g$
. Let
$E_{u/s}^*(g)\subset T^*(SM)$
be the annihilators of
$E_{u/s}(g)\oplus E_0(g)$
, that is,
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu92.png?pub-status=live)
There are two resolvents bounded on
$L^2$
for
$X_{g}$
defined for
${\textrm{Re}}(\unicode{x3bb} )>0$
by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu93.png?pub-status=live)
for
$f\in L^2(SM,d\mu ^{\textrm{L}}_g)$
. They solve
$(-X_g\pm \unicode{x3bb} )R_g^\pm (\unicode{x3bb} )={\textrm{Id}}$
on
$L^2$
. The following results are proved in [Reference Faure and SjöstrandFS11], and we use here the presentation of [Reference Dang, Guillarmou, Rivière and ShenDGRS20, Sections 3.2 and 3.3] due to the need for uniformity with respect to g: there is
$c_0>0$
depending only on g, locally uniform with respect to g (
$c_0$
depends only on the Anosov exponents of contraction/dilation of
$d\varphi _1^g$
), such that for each
$N_0>0,N_1>16N_0$
,
$R_g^\pm (\unicode{x3bb} )$
admits a meromorphic extension in
${\textrm{Re}}(\unicode{x3bb} )>-c_0N_0$
as a bounded operator
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqn37.png?pub-status=live)
where
$\mathcal {H}^{\pm m_g^{N_0,N_1}}$
are Hilbert spaces depending on
$N_0>0,N_1>0$
satisfying the properties
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu94.png?pub-status=live)
and defined by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu95.png?pub-status=live)
and
$A_{m_g^{N_0,N_1}}$
is an invertible pseudo-differential operator with inverse having principal symbol
$e^{-m_g^{N_0,N_1}\log f}$
. Here
${\textrm{Op}}$
denotes a quantization (with a fixed small semi-classical parameter to ensure that
${\textrm{Op}}({e^{m_g^{N_0,N_1}\log f}})$
is invertible), while
$m_g^{N_0,N_1}\in S^{0}(T^*(SM))$
,
$f\in S^1(T^*(SM), [1,\infty ))$
(the usual classes of symbols) are homogeneous of respective degree
$0$
and
$1$
in
$|\xi |>R$
, for some
$R>1$
independent of g, and constructed from the lifted flow
$\Phi _t^g=((d\varphi _t^{g})^{-1})^T$
acting on
$T^*(SM)$
. The function f can be taken depending only on
$g_0$
for g in a small enough
$C^\infty $
neighborhood
$\mathcal {U}$
of
$g_0$
. Moreover, there are small conic neighborhoods
$C_{u}(g_0)$
and
$C_s(g_0)$
of
$E^*_u(g_0)$
and
$E^*_s(g_0)$
such that, for any smaller open conic neighborhood
$C^{\prime }_u(g_0)\subset C_u(g_0)$
of
$E_u^*(g_0)$
and
$C^{\prime }_s(g_0)\subset C_s(g_0)$
of
$E_s^*(g_0)$
,
$m_g^{N_1,N_1}$
satisfies
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqn38.png?pub-status=live)
and
$m_g(x,\xi )\in [-2N_0,2N_1]$
for all
$(z,\xi )\in T^*(SM)$
. We note that [Reference Dang, Guillarmou, Rivière and ShenDGRS20, Lemma 3.3] shows that
$m_g^{N_0,N_1}$
is smooth with respect to the metric g and that f can be taken to be independent of g for g close enough to
$g_0$
. The spaces
$\mathcal {H}^{m_g^{N_0,N_1}}$
are called anisotropic Sobolev spaces. The pseudodifferential operators
$A_{m_g^{N_0,N_1}}$
belong to the class
$\Psi ^{2N_1}(SM)$
but also to some anisotropic subclass denoted
$\Psi ^{m_g^{N_0,N_1}}(SM)$
admitting composition formulas; we refer to [Reference Faure, Roy and SjöstrandFRS08, Reference Faure and SjöstrandFS11] for details.
Eventually, [Reference Dang, Guillarmou, Rivière and ShenDGRS20, Proposition 6.1] shows that there is a small open neighborhood
$W_\delta $
of the circle
$\{\unicode{x3bb} \in \mathbb {C} ~|~ |\unicode{x3bb} |=\delta \}$
for some small
$\delta>0$
so that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqn39.png?pub-status=live)
is continuous. (In [Reference Dang, Guillarmou, Rivière and ShenDGRS20, Proposition 6.1], a small semi-classical parameter
$h>0$
appears: we can just fix this parameter small enough. It does not play any role here except in the quantization procedure
${\textrm{Op}}$
. We also add that in [Reference Dang, Guillarmou, Rivière and ShenDGRS20, Proposition 6.1],
$N_1$
is chosen to be equal to
$20N_0$
for notational convenience, but the proof does not use that fact.)
4.2 The operator
$\Pi _2^g$
in terms of resolvents
Following [Reference GuillarmouGui17], the link between
$\Pi ^g$
and the resolvent is given by the Laurent expansion
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu96.png?pub-status=live)
where
$R_g^{+}(\unicode{x3bb} )$
has a pole of order
$1$
,
$R_g^\pm (0)$
is defined by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu97.png?pub-status=live)
and
$R_g^-(0)=-(R_g^+(0))^*$
where the adjoint is with respect to the Liouville measure.
Lemma 4.2. Let
$\chi \in C_c^\infty (\mathbb {R})$
be even and equal to
$1$
in
$[-T,T]$
and supported in the interval
$(-T-1,T+1)$
. Then
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqn40.png?pub-status=live)
Proof For
${\textrm{Re}}(\unicode{x3bb} )>0$
, we can write, by integration by parts,
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu98.png?pub-status=live)
Then taking the limit as
$\unicode{x3bb} \to 0$
, we obtain
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu99.png?pub-status=live)
and summing gives the result.
Next, we remark that, using that
$\varphi _t^g(x,-v)=-\varphi _{-t}^g(x,v)$
(where multiplication by
$-1$
is the symmetry in the fibers of
$SM$
), it is straightforward to check that, for all
$t\in \mathbb {R}$
,
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu100.png?pub-status=live)
which also implies that
${\pi _2}_*R_g^+(0)e^{tX_g}\pi _2^*=-{\pi _2}_*R_g^-(0)e^{-tX_g}\pi _2^*$
and thus
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqn41.png?pub-status=live)
We will prove that these three terms depend continuously on g. Note that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu101.png?pub-status=live)
and thus the g-continuity of this term is immediate. Now, we claim the following result.
Lemma 4.3. There exist
$T>0$
large enough and a neighborhood
$\mathcal {U}^{\prime }\subset \mathcal {U}$
of
$g_0$
in
$\mathcal {M}^{\textrm{An}}$
so that for all
$x\in M$
and all
$g\in \mathcal {U}^{\prime }$
the exponential map of g in the universal cover
$\widetilde {M}$
,
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu102.png?pub-status=live)
is a diffeomorphism onto its image and
$\Phi _{t}^g(V^*\cap \ker \iota _{X_g})\subset C^{\prime }_u(g_0)$
for all
$t\geq T$
, if
$\Phi _t^g:=((d\varphi _t^g)^{-1})^T$
is the symplectic lift of
$\varphi _t^g$
,
$V^*\subset T^*(SM)$
is the annihilator of the vertical bundle
$V=\ker d\pi _0\subset T(SM)$
and
$\iota _X:T^*(SM)\to \mathbb {R}$
is the contraction
$\iota _{X_g}(\xi )=\xi (X_g)$
.
We also mention here as it is used in the following proof that, as a consequence of hyperbolicity,
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu103.png?pub-status=live)
This can be found in [Reference PaternainPat99, Theorem 2.50], for instance (formulated for the tangent bundle
$T(SM)$
but the adaptation to
$T^*(SM)$
is straightforward). The T in Lemma 4.2 will be chosen accordingly so that Lemma 4.3 is satisfied.
Proof By [Reference Dang, Guillarmou, Rivière and ShenDGRS20, Lemma 3.1], the cone
$\mathcal {C}^{\prime }_u(g_0)$
can be chosen so that there exist
$T>0$
and
$\mathcal {U}^{\prime }$
such that, for all
$t\geq T$
and all
$g\in \mathcal {U}^{\prime }$
,
$\Phi _{t}^g(C^{\prime }_u(g_0))\subset C^{\prime }_u(g_0)$
. We also know that
$\Phi _{T_0}^{g_0}(V^*\cap \ker \iota _{X_g}) \subset C^{\prime }_u(g_0)$
for some
$T_0>T$
by hyperbolicity of
$g_0$
(that is, the stable bundle
$E_s^*$
only intersects trivially the vertical bundle
$V^* \cap \ker \imath _{X_g}$
), but by continuity of
$g\mapsto \Phi _{T_0}^g$
, the same holds for all g in some possibly smaller neighborhood
$\mathcal {U}^{\prime \prime }\subset \mathcal {U}^{\prime }$
, thus for all
$t\geq T_0$
and all
$g\in \mathcal {U}^{\prime \prime }$
,
$\Phi _{t}^g(V^*\cap \ker \iota _{X_g})\subset C^{\prime }_u(g_0)$
. Now, we claim that, up to choosing
$\mathcal {U}^{\prime \prime }$
even smaller, the exponential map is a diffeomorphism on
$\{|v|_g\leq T\}$
in the universal cover: indeed, Anosov geodesic flows have no pair of conjugate points.
4.3 Proof or Proposition 4.1
Let us define
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu104.png?pub-status=live)
Proposition 4.1 is a consequence of the following two lemmas.
Lemma 4.4. For each
$g\in \mathcal {U}^{\prime }$
,
$\Omega _1^g\in \Psi ^{-1}(M;S^2T^*M)$
with principal symbol
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu105.png?pub-status=live)
for some
$c_n>0$
depending only on
$n=\dim M$
and
$A_2$
some positive definite endomorphism defined in Lemma 2.1, and the map
$g\mapsto \Omega _1^g$
is continuous with respect to the smooth topology on
$\mathcal {U}^{\prime }$
and the usual Fréchet topology on
$\Psi ^{-1}(M;S^2T^*M)$
.
Proof The fact that, for each
$g\in \mathcal {M}^{\textrm{An}}$
, the operator
$\Omega _1^g\in \Psi ^{-1}(M;S^2T^*M)$
is proved in [Reference GuillarmouGui17, Theorem 3.5]. The computation of the principal symbol follows from the computation [Reference Sharafutdinov, Skokan and UhlmannSSU05, Reference Stefanov and UhlmannSU04] and is done in detail in our setting in [Reference Goüezel and LefeuvreGL, Theorem 4.4]. We need to check the continuity with respect to g in the
$\Psi ^{-1}(M;S^2T^*M)$
topology and we can proceed as in [Reference Stefanov and UhlmannSU04, Propositions 1 and 2]. For
$h\in C^\infty (M;S^2T^*M)$
, we can write explicitly in
$(x_i)_i$
coordinates in the universal cover
$\widetilde {M}$
near a point
$p\in \widetilde {M}$
,
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu106.png?pub-status=live)
where
$p_{ij}(x,v)$
are homogeneous polynomials of order
$2$
in the v variable,
$\widetilde {h}\in C^\infty (\widetilde {M};S^2T^*M)$
is the lift of h to the universal cover
$\widetilde {M}$
, and
$dS_x$
is the natural measure on the sphere
$S_x\widetilde {M}$
. Using Lemma 4.3, we can perform the change of coordinates
$(t,v)\in (0,T)\times S_x\widetilde {M}\mapsto y:=\exp ^{\widetilde {g}}_{x}(tv)\in \widetilde {M}$
, and we get the distance
$t=d_{\widetilde {g}}(x,y)$
in
$\widetilde {M}$
, and
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu107.png?pub-status=live)
for some
$J^g_{x}(y)$
smooth in
$x,y,g$
. We claim that this implies that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu108.png?pub-status=live)
for some
$K_g(x,y)$
which is smooth in
$(g,x,y)$
outside the diagonal
$x=y$
and, near the diagonal, has the form (for some
$L<\infty )$
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu109.png?pub-status=live)
with
$c_\ell $
a matrix valued function, smooth in all its variables and
$\omega _{\ell , g,x}(v)$
a vector-valued function smooth in
$g,x$
, homogeneous of degree
$-(n-1)$
in
$v\in \mathbb {R}^n$
. Indeed, one can work in the universal cover
$\widetilde {M}$
where
$x_i$
are globally defined coordinates, so that, writing
$h(x)=\sum _{i,j}h_{ij}(x)dx_idx_j$
and
$p=\sum _{ij}p_{ij}(x)dx_idx_j$
, we get that
$K_g(x,y)$
is a matrix with coefficients
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu110.png?pub-status=live)
where
$F^g_i(x,y)=-dx_i(\nabla ^{\widetilde {g}}_yd_{\widetilde {g}}(x,y))$
and
$G_i^g(x,y)=dx_i(\nabla ^{\widetilde {g}}_x d_{\widetilde {g}}(x,y))$
. Now we can use the standard fact (see, for example, [Reference Stefanov and UhlmannSU04, Lemma 1]) that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu111.png?pub-status=live)
(and likewise for
$dx_i(\nabla ^g_y d_{\tilde {g}}(x,y))$
by symmetry) where
$H_{ij}^{k}(g,x,y)$
are smooth in all variables and positive definite for
$x=y$
. The kernel
$K_g$
is thus smooth outside the diagonal (as a function of
$g,x,y$
), and can be written near the diagonal as a sum of terms of the form
$c(g,x,y)\omega _{g,x}(x-y)$
where c is smooth in all its variables and
$\omega _{g,x}(v)$
is a homogeneous distribution of degree
$-(n-1)$
in the variable v, smooth in
$g,x$
. The off-diagonal term for the Fréchet topology is then clearly smooth in g, while the near-diagonal term has full local symbols that are Fourier transforms of
$c(g,x,x-v)\omega _{g,x}(v)$
:
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu112.png?pub-status=live)
It is then a standard and easy exercise to check that this provides uniform bounds on semi-norms of the symbol. (Alternatively, the semi-norms on the full symbol are equivalent to semi-norms in the space of distributions on
$M\times M$
that are conormal to the diagonal, defined through differentiations of
$K_g(x,y)$
with respect to smooth fields tangent to
${\textrm{diag}}(M\times M)$
; see [Reference MelroseMel, Ch. 5, Proposition 6.1.1 and its proof]. Such norms for
$K_g$
are clearly uniformly bounded in terms of g.) We deduce the continuity (and indeed, smoothness) of
$\Omega _1^g$
as an element of
$\Psi ^{-1}(M;S^2T^*M)$
with respect to the metric g.
Lemma 4.5. The operator
$\Omega _2^g$
has a smooth Schwartz kernel for each
$g\in \mathcal {U}^{\prime }$
, and the map
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu113.png?pub-status=live)
is continuous if we identify
$\Omega _g^2$
with its Schwartz kernel.
Proof First we observe that if
$B\in \Psi ^0(SM)$
is chosen, independently of g, so that
$B^*=B$
and B is microsupported in a small conic neighborhood of
$V^*$
not intersecting
$\mathcal {C}_u(g_0)$
and equal microlocally to the identity in a slightly smaller conic neighborhood of
$V^*$
, then
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu114.png?pub-status=live)
with
$S_g$
a continuous family of smoothing operators. This decomposition is a consequence of the fact that
$\pi _2^*$
maps
$C^{-\infty }(M;S^2T^*M)$
to the space
$C^{-\infty }_{V^*}(SM)$
of distributions with wavefront set contained in
$V^*$
(
$\pi _2^*$
being essentially a pullback, this follows, for instance, from [Reference HörmanderHör03, Theorem 8.2.4]). We will show that the operator
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu115.png?pub-status=live)
is a continuous family (with respect to g) of smoothing operators. We need to show that for each
$N>0$
,
$\Omega _3^g: H^{-N}(SM)\to H^{N}(SM)$
is a continuous family with respect to g of bounded operators. To study
$R_g^-(0)$
, it suffices to write it in the form
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqn42.png?pub-status=live)
with
$\delta $
small enough so that the only pole of
$R_g^-(\unicode{x3bb} )$
in
$|\unicode{x3bb} |\leq \delta $
is
$\unicode{x3bb} =0$
(this is possible for g close enough to
$g_0$
by continuity of
$g\mapsto R_g^-(\unicode{x3bb} )$
as proved in [Reference Dang, Guillarmou, Rivière and ShenDGRS20]; note that the spectrum (the Pollicott–Ruelle resonances) depends continuously on the metric, as was shown by [Reference BonthonneauBon20]), so that this amounts to analyzing
$R_g^-(\unicode{x3bb} )$
on
$\{|\unicode{x3bb} |=\delta \}$
. We decompose
$B=B^1+B^2$
with
$B^i\in \Psi ^0(SM)$
, where
${\textrm{WF}}(B^1)$
is contained in a conic neighborhood of
$\ker \iota _{X_{g_0}}$
not intersecting the annihilator
$E_0(g_0)^*$
of
$E_u(g_0)\oplus E_s(g_0)$
(the neutral direction) and
${\textrm{WF}}(B^2)\cap \ker \iota _{X_{g_0}}=\emptyset $
(
$B^2$
is microsupported in the elliptic region). For
$i=1,2$
we let
$B_T^{i}\in \Psi ^0(SM)$
be microsupported in a conic neighborhood of
$\bigcup _{t\in [T,T+1]}\Phi _{t}^g({\textrm{WF}}(B^i))$
, so that by Egorov (or simply the formula of composition of
$\Psi ^{0}(SM)$
with diffeomorphisms of
$SM$
),
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu116.png?pub-status=live)
for some continuous family
$(g,t)\mapsto S_{g,i}^{\prime }(t)$
of smoothing operators (for g close enough to
$g_0$
). We note that by taking
$\mathcal {U}^{\prime }$
small enough and
${\textrm{WF}}(B^1)$
close enough to
$V^*\cap \ker \iota _{X_{g_0}}$
, Lemma 4.3 ensures that we can choose
$B^1_T$
depending only on T (thus uniform in
$g\in \mathcal {U}^{\prime }$
) so that
${\textrm{WF}}(B_T^1)\subset C^{\prime }_u(g_0)$
. Thus
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu117.png?pub-status=live)
for some continuous family
$g\mapsto S_{g,1}^{\prime \prime }$
of smoothing operators. Next we use (4.1) with the choice
$N_0=N+1$
and
$N_1/16=N+2$
. Since, by (4.2),
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu118.png?pub-status=live)
we obtain, using the composition properties in [Reference Faure, Roy and SjöstrandFRS08, Theorem 8] that
$A_{m_g^{N_0,N_1}}B^1_T\in \Psi ^{-N-1}(SM)$
is uniformly bounded with respect to g and continuous as a map
$g\in \mathcal {U}^{\prime }\mapsto A_{m_g^{N_0,N_1}}B^1_T\in \mathcal {L}(H^{-N}(SM), H^1(SM))$
. In particular,
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqn43.png?pub-status=live)
is continuous. Next, we deal with the ‘elliptic region’ term, that is, the term
$B^2$
. The idea is to show it is smoothing, since it is a Schwartz function of
$X_g$
microlocalized in the elliptic region of
$X_g$
. First,
${\textrm{WF}}(B^2_T)$
does not intersect
$\ker \iota _{X_g}$
for
$g\in \mathcal {U}^{\prime }$
after possibly reducing
$\mathcal {U}^{\prime }$
since it does not intersect
$\ker \iota _{X_{g_0}}$
. Moreover, we have
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu119.png?pub-status=live)
and since
${\textrm{WF}}(B^2_T)$
does not intersect
$\ker \iota _{X_g}$
for
$g\in \mathcal {U}^{\prime }$
, there exist by microlocal ellipticity [Reference Dyatlov and ZworskiDZ19, Proposition E.32] a family
$Q_g\in \Psi ^{-2N}(SM)$
and
$Z_g\in \Psi ^{-\infty }(SM)$
, both continuous with respect to g, so that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu120.png?pub-status=live)
We write
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu121.png?pub-status=live)
where
$Z_g^{\prime }\in \mathcal {L}(H^{-N}(SM),H^N(SM))$
continuously in g. Since
$\int _T^{T+1} \chi ^{\prime }(t)e^{-tX_g}B^2\, dt$
is continuous in g as a bounded map
$\mathcal {L}(H^{-N}(SM))$
and
$Q_g$
is continuous in g as a bounded map
$\mathcal {L}(H^{-N}(SM),H^N(SM))$
, we get
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu122.png?pub-status=live)
continuously in
$g\in \mathcal {U}^{\prime }$
. Combine these facts with (4.7), (4.3) and (4.6), we deduce that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu123.png?pub-status=live)
is continuous as a map with values in
$\mathcal{L}(H^{-N}(SM),L^2(SM))$
. Finally, using that
${\textrm{WF}}(B)\cap C_u(g_0)=\emptyset $
and
$-m_{g}^{N_0,N_1}\leq -2N-4$
outside
$C_u(g_0)$
by (4.2), we have that
$B(A_{m_g^{N_0,N_1}})^{-1}\in \Psi ^{-2N-4}(SM)$
uniformly in g (again using [Reference Faure, Roy and SjöstrandFRS08, Theorem 8] and [Reference Dang, Guillarmou, Rivière and ShenDGRS20, Lemma 3.2]) and the following map is continuous:
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu124.png?pub-status=live)
This shows that
$\mathcal {U}^{\prime }\ni g\mapsto \Omega _3^g \in \mathcal {L}(H^{-N}(M;S^2T^*M),H^N(M;S^2T^*M))$
is continuous. The terms involving the smoothing remainders
$S_g$
appearing in the difference between
$\Omega _2^g$
and
$\Omega _3^g$
can be dealt with using the same argument, and indeed are even simpler to consider. The proof is then complete.
The proof of Proposition 4.1 is simply the combination of Lemmas 4.4 and 4.5.
As a corollary we prove Theorem 1.3.
4.4 Proof of Theorem 1.3
Let
$g_0\in \mathcal {M}^{\textrm{An}}$
and assume
$g_0$
has non-positive curvature if
$n\geq 3$
. Using Lemma 2.4, for
$g_1,g_2\in \mathcal {M}$
close enough to
$g_0$
in
$C^{k+3,\alpha }$
norm, we can find
$\psi \in \mathcal {D}_0^{k+1,\alpha }$
(with
$k\geq 5$
to be chosen later), depending in a
$C^2$
fashion on
$(g_1,g_2)$
such that
$D_{g_1}^*(\psi ^*g_2)=0$
. Moreover,
$g_2^{\prime }=\psi ^*g_2$
satisfies
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu125.png?pub-status=live)
for some C depending only on
$g_0$
. We can then rewrite the proof of Theorem 1.2, replacing
$g_0$
by
$g_1$
. Let
$\Psi _{g_1}(g_2)={\textbf {P}}(-J_{g_1}^u-a_{g_1,g_2}+\int _{S_{g_0}M}a_{g_1,g_2}\, d\mu _{g_1}^L)$
be the map (3.2) with
$(g_1,g_2)$
replacing
$(g_0,g)$
, and
$\Phi _{g_1}(g_2)=I_{\mu ^{\textrm{L}}_{g_1}}(g_1,g_2)$
, where
$a_{g_1,g_2}$
is the time reparametrization coefficient (2.14) in the conjugacy between the flows
$\varphi ^{g_1}$
and
$\varphi ^{g_2}$
, and the pressure and the stretch are taken with respect to the flow
$\varphi ^{g_1}$
. Combining [Reference ContrerasCon92, Theorem C] and Proposition C.1, the maps
$(g_1,g_2)\mapsto \Psi _{g_2}(g_1)$
and
$(g_1,g_2)\mapsto \Phi _{g_1}(g_2)$
are
$C^3$
in
$g_2$
if k is chosen large enough, and each
$g_2$
-derivative of order
$\ell \leq 3$
is continuous with respect to
$(g_1,g_2)\in C^{k+3}\times C^{k+3}$
(again k is fixed large enough). Following the proof of Proposition 3.5, this gives that for
$g_1,g_2$
smooth but close enough to
$g_0$
in
$\mathcal {M}^{k+3,\alpha }$
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu126.png?pub-status=live)
where
$C_n$
depends only on
$n=\dim M$
, and
$C^{\prime }_{g_1}$
depends on
$\|g_1\|_{C^{k_0,\alpha }}$
for some fixed
$k_0$
.
Combining Proposition 4.1 and Lemma 2.2, we deduce that there exist
$C_{g_0},C^{\prime }_{g_0}>0$
depending only on
$g_0$
so that for
$g_1,g_2\in \mathcal {M}$
in a small enough neighborhood of
$g_0$
in the
$C^{k+3,\alpha }$
topology (for
$k\geq k_0$
),
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu127.png?pub-status=live)
This means that there exists
$\varepsilon>0$
depending on
$g_0$
and k large enough so that, for all
$g_1,g_2\in \mathcal {M}$
smooth satisfying
$\|g_j-g_0\|_{C^{k+3,\alpha }(M)}\leq \varepsilon $
, the estimates of Proposition 3.5 with
$(g_1,g_2)$
replacing
$(g_0,g)$
hold uniformly with respect to
$(g_1,g_2)$
. This proves the theorem.
5 Distances from the marked length spectrum
In this section we discuss different notions of distances involving the marked length spectrum on the space of isometry classes of negatively curved metrics. Again, if the X-ray transform
$I_2$
were known to be injective, it is likely that one could only assume the Anosov property for the metrics in this paragraph.
5.1 Length distance
We define the following map.
Definition 5.1 Let k be as in Theorem 1.3. We define the marked length distance map
$d_L: \mathcal {M}^{k,\alpha }\times \mathcal {M}^{k,\alpha }\to \mathbb {R}^+$
by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu128.png?pub-status=live)
This is indeed well defined. If
$g_1, g_2$
are two such metrics, then there exists a constant
$C = C(g_1,g_2) \geq 1$
such that for all
$(x,v) \in TM$
,
$(1/C) \times |v|_{g_1(x)} \leq |v|_{g_2(x)} \leq C \times |v|_{g_1(x)}$
. As a consequence, using that a geodesic is a minimizer of the length among a free homotopy class, we obtain
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu129.png?pub-status=live)
and the lower bound follows from a similar computation. We obtain the following corollary of Theorem 1.3.
Corollary 5.2. The map
$d_L$
descends to the set of isometry classes near
$g_0$
and defines a distance in a small
$C^{k,\alpha }$
-neighborhood of the isometry class of
$g_0$
.
Proof It is clear that
$d_L$
is invariant by action of diffeomorphisms homotopic to the identity since
$L_g=L_{\psi ^*g}$
for such diffeomorphisms
$\psi $
. Now let
$g_1,g_2,g_3$
three metrics. We have
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu130.png?pub-status=live)
thus
$d_L$
satisfies the triangle inequality. Finally, by Theorem 1.3, if
$d_L(g_1,g_2)=0$
with
$g_1,g_2$
in the
$C^{k,\alpha }$
neighborhood
$\mathcal {U}_{g_0}$
of Theorem 1.3, we have
$g_1$
isometric to
$g_2$
, showing that
$d_L$
produces a distance on the quotient of
$\mathcal {U}_{g_0}$
by diffeomorphisms.
We also note that Theorem 1.3 states that there is
$C_{g_0}>0$
such that for each
$g_1,g_2\in C^{k,\alpha }(M;S^2T^*M)$
close to
$g_0$
there is a diffeomorphism such that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu131.png?pub-status=live)
5.2 Thurston’s distance
We also introduce the Thurston distance on metrics with topological entropy
$1$
, generalizing the distance introduced by Thurston in [Reference ThurstonThu98] for surfaces on Teichmüller space (all hyperbolic metrics on surface have topological entropy equal to
$1$
). We denote by
$\mathcal {E}$
(respectively,
$\mathcal {E}^{k,\alpha }$
) the space of negatively curved metrics in
$\mathcal {M}$
(respectively, in
$\mathcal {M}^{k,\alpha }$
) with topological entropy
$\textbf{h}_{{\textrm{top}}}(g)=1$
. (Let us also recall here for the sake of clarity that
$\textbf{h}_{{\textrm{top}}}(\unicode{x3bb} ^2 g)=\textbf{h}_{{\textrm{top}}}(g)/\unicode{x3bb} $
, for
$\unicode{x3bb}> 0$
.) With the same arguments as in Lemma 3.6, this is a codimension-one submanifold of
$\mathcal {M}$
and if
$g_0 \in \mathcal {E}^{k,\alpha }$
, we have
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqn44.png?pub-status=live)
Definition 5.3 We define the Thurston non-symmetric distance map
$d_T: \mathcal {E}^{k,\alpha } \times \mathcal {E}^{k,\alpha } \to \mathbb {R}^+$
by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu132.png?pub-status=live)
Note that the finiteness of the previous quantity also follows from the same argument as the one justifying the finiteness of Definition 5.1. Its non-negativity will be a consequence of Lemma 5.5, where it is proved that this can be expressed in terms of the geodesic stretch. We will prove the following proposition.
Proposition 5.4. The map
$d_T$
descends to the set of isometry classes of metrics in
$\mathcal {E}^{k,\alpha }$
(for
$k \in \mathbb {N}$
large enough,
$\alpha \in (0,1)$
) with topological entropy equal to
$1$
and defines a non-symmetric distance in a small
$C^{k,\alpha }$
-neighborhood of the diagonal.
Moreover, this distance is non-symmetric in the pair
$(g_1,g_2)$
which is also the case of the original distance introduced by Thurston [Reference ThurstonThu98], but this is just an artificial limitation (Thurston [Reference ThurstonThu98]): ‘It would be easy to replace L (in Thurston’s notation,
$L(g,h) = \limsup _{j\to \infty } \log ({L_{g}(c_j)}/{L_{h}(c_j)})$
) by its symmetrization
$\frac 12(L(g, h)+L(h, g))$
, but it seems that, because of its direct geometric interpretations, L is more useful just as it is.’ In order to justify that this is a distance, we start with the following lemma.
Lemma 5.5. Let
$g_1,g_2 \in \mathcal {M}$
be negatively curved. Then
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu133.png?pub-status=live)
Note that there is no need to assume
$g_1$
and
$g_2$
are close in this lemma: this follows from Appendix B, where we discuss the fact that the stretch (and the time reparametrization) is well defined despite the fact that the metrics may not be close. Here m is seen as an invariant ergodic measure for the flow
$\varphi ^{g_1}_t$
living on
$S_{g_1}M$
. However, writing
$M = \Gamma \backslash \widetilde {M}$
with
$\Gamma \simeq \pi _1(M,x_0)$
for
$x_0 \in M$
, it can also be identified with a geodesic current on
$\partial _\infty \widetilde {M} \times \partial _\infty \widetilde {M} \setminus \Delta $
, that is, a
$\Gamma $
-invariant Borel measure, also invariant by the flip
$(\xi ,\eta ) \mapsto (\eta ,\xi )$
on
$\partial _\infty \widetilde {M} \times \partial _\infty \widetilde {M} \setminus \Delta $
. This point of view has the advantage of being independent of
$g_1$
(see [Reference Schapira and TapieSTar]).
Proof First of all, we claim that (as pointed out to us by one of the referees, the map
$\mathfrak {M}_{\textrm{inv}} \ni m \mapsto I_m(g_1,g_2)$
is continuous and linear on a compact convex set; it thus achieves its maximum on the extremal points of the convex sets (the ergodic measures) so the argument could be shortened)
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu134.png?pub-status=live)
Of course, it is clear that
$\mathop {\textrm{sup}}_{m \in \mathfrak {M}_{\textrm{inv,erg}}} I_m(g_1,g_2) \leq \mathop {\textrm{sup}}_{m \in \mathfrak {M}_{\textrm{inv}}} I_m(g_1,g_2)$
and thus we are left to prove the reverse inequality. By compactness, we can consider a measure
$m_0 \in \mathfrak {M}_{\textrm{inv}}$
realizing
$\mathop {\textrm{sup}}_{m \in \mathfrak {M}_{\textrm{inv}}} I_m(g_1,g_2)$
. By the Choquet representation theorem (see [Reference WaltersWal82, pp. 153]), there exists a (unique) probability measure
$\tau $
on
$\mathfrak {M}_{\textrm{inv,erg}}$
such that
$m_0$
admits the ergodic decomposition
$m_0 = \int _{\mathfrak {M}_{\textrm{inv,erg}}}m ~d\tau (m)$
. Thus
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu135.png?pub-status=live)
which eventually proves the claim.
Let
$(c_j)_{j \in \mathbb {N}}$
be a subsequence such that
$\lim _{j \rightarrow +\infty } L_{g_2}(c_j)/L_{g_1}(c_j)$
realizes the
$\limsup $
. Then, by compactness, we can extract a subsequence such that
$\delta _{g_1}(c_j) \rightharpoonup m \in \mathfrak {M}_{\textrm{inv}}$
. Thus:
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu136.png?pub-status=live)
which proves, using our preliminary remark, that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu137.png?pub-status=live)
To prove the reverse inequality, we consider a measure
$m_0 \in \mathfrak {M}_{\textrm{inv,erg}}$
such that
$I_{m_0}(g_1,g_2) = \mathop {\textrm{sup}}_{m \in \mathfrak {M}_{\textrm{inv,erg}}} I_m(g_1,g_2)$
(which is always possible by compactness). Since
$m_0$
is invariant and ergodic, there exists a sequence of free homotopy classes
$(c_j)_{j \in \mathbb {N}}$
such that
$\delta _{g_1}(c_j) \rightharpoonup m_0$
(by [Reference SigmundSig72]). Then, as previously, we have
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu138.png?pub-status=live)
which provides the reverse inequality.
We can now prove Proposition 5.4.
Proof of Proposition 5.4. By (2.17), for
$g_1,g_2 \in \mathcal {E}^{k,\alpha }$
, we have that
$I_{\mu _{g_1}^{\textrm{BM}}}(g_1,g_2) \geq 1$
and thus, by Lemma 5.5, we obtain that
$d_T(g_1,g_2)\geq 0$
(note that
$g_1$
and
$g_2$
do not need to be close for this property to hold). Moreover, the triangle inequality is immediate for this distance. Eventually, if
$d_T(g_1,g_2) = 0$
, then
$0 \leq \log I_{\mu _{g_1}^{\textrm{BM}}}(g_1,g_2) \leq d_T(g_1,g_2)=0$
, that is,
$I_{\mu _{g_1}^{\textrm{BM}}}(g_1,g_2) = 1$
and, by Theorem 2.9, it implies that
$g_1$
is isometric to
$g_2$
if
$g_2$
is close enough to
$g_1$
in the
$C^{k,\alpha }$
topology (note that this neighborhood depends on
$g_1$
).
We now investigate in more detail the structure of the distance
$d_T$
. A consequence of Lemma 5.5 is the following expression of the Thurston Finsler norm.
Lemma 5.6. Let
$g_0 \in \mathcal {E}^{k,\alpha }$
and
$(g_t)_{t \in [0,\varepsilon )}$
be a smooth family of metrics and let
$f := \partial _tg_t|_{t=0}$
. Then
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqn45.png?pub-status=live)
The norm
$\|\cdot \|_T$
is a Finsler norm on
$T_{g_0}\mathcal {E}^{k,\alpha }\cap \ker D^*_{g_0}$
Proof We introduce
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu139.png?pub-status=live)
and write
$a_t := a_{g_0,g_t}$
for the time reparametrization (as in (2.14)). Then
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu140.png?pub-status=live)
since
$\dot {a}_0=\partial _{t}a_{t}|_{t=0}$
and
$\pi _2^*f$
are cohomologous by Lemma 3.3. This also shows that the derivative exists. The inversion of the limit and the
$\mathop {\textrm{sup}}$
follows from the fact that, writing
$F_t(m) := \int _{S_{g_0}M} (a_t-1)/t ~~dm$
, we have
$\mathop {\textrm{sup}}_{m \in \mathfrak {M}_{\textrm{inv,erg}}} |F_t(m)-F_0(m)| \rightarrow _{t \rightarrow 0} 0$
. Note that, up to taking a large
$k \in \mathbb {N}$
and iterating the same computation for higher-order derivatives, this shows that
$t \mapsto u(t)$
(thus
$t \mapsto d_T(g_0,g_t)$
) is at least
$C^2$
.
We now prove that this is a Finsler norm in a neighborhood of the diagonal. We fix
$g_0 \in \mathcal {E}^{k,\alpha }$
. By Lemma 2.4, isometry classes near
$g_0$
can be represented by solenoidal tensors, namely, there exists a
$C^{k,\alpha }$
-neighborhood
$\mathcal {U}$
of
$g_0$
such that for any
$g \in \mathcal {U}$
, there exists a (unique)
$\psi \in \mathcal {D}_0^{k+1,\alpha }$
such that
$D^*_{g_0}\psi ^*g = 0$
. Moreover, if
$g \in \mathcal {E}^{k,\alpha }$
, then
$\psi ^*g \in \mathcal {E}^{k,\alpha }$
. As a consequence, using (5.1), the statement now boils down to proving that (5.2) is a norm for solenoidal tensors
$f \in C^{k,\alpha }(M;S^2T^*M)$
such that
$\int _{S_{g_0}M} \pi _2^*f \,d\mu _{g_0}^{\textrm{BM}} = 0$
. Since the triangle inequality,
$\mathbb {R}_+$
-scaling and non-negativity are immediate, we simply need to show that
$\|f\|_T = 0$
implies
$f=0$
. Now, for such a tensor f, we have
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu141.png?pub-status=live)
and this supremum is achieved for
$m=\mu ^{\textrm{BM}}_{g_0}$
and
$\textbf{P}(\pi _2^*f) = 1$
. As a consequence, the equilibrium state associated to the potential
$\pi _2^*f$
is the Bowen–Margulis measure
$\mu ^{\textrm{BM}}_{g_0}$
(the equilibrium state associated to the potential
$0$
) and thus
$\pi _2^*f$
is cohomologous to a constant
$c \in \mathbb {R}$
(see [Reference Hasselblatt and FisherHF19, Theorem 9.3.16]) which has to be
$c=0$
since the average of
$\pi _2^*f$
with respect to Bowen–Margulis is equal to
$0$
, that is, there exists a Hölder-continuous function u such that
$\pi _2^*f = Xu$
. Since
$f \in \ker D^*_{g_0}$
, the s-injectivity of the X-ray transform
$I_2^{g_0}$
implies that
$f \equiv 0$
.
The asymmetric Finsler norm
$\|\cdot \|_T$
induces a distance
$d_F$
between isometry classes, namely,
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu142.png?pub-status=live)
It is easy to prove that
$d_T(g_1,g_2) \leq d_F(g_1,g_2)$
, which shows that
$d_F$
is indeed a distance in a neighborhood of the diagonal, just like
$d_T$
. Indeed, consider a
$C^1$
-path
$\gamma : [0,1] \rightarrow \mathcal {E}$
such that
$\gamma (0)=g_1,\gamma (1)=g_2$
. Then, considering
$N \in \mathbb {N}, t_i := i/N$
, we have, by the triangle inequality,
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu143.png?pub-status=live)
which proves the claim (note that we here use the fact that
$t \mapsto d_T(g_0,g_t)$
is at least
$C^2$
). In [Reference ThurstonThu98], Thurston proves that, on restriction to Teichmüller space, the asymmetric Finsler norm induces the distance
$d_T$
, that is,
$d_T = d_F$
. We make the following conjecture.
Conjecture 5.7. The distances
$d_T$
coincide with
$d_F$
for isometry classes of negatively curved metrics with topological entropy equal to
$1$
.
This conjecture would imply the marked length spectrum rigidity conjecture. Indeed, as mentioned just after Theorem 2.9, two metrics with the same marked length spectrum have the same topological entropy and there is no harm (up to a scaling of the metrics) in assuming that this topological entropy is equal to
$1$
. Then, if the previous conjecture is true, using that their Thurston distance
$d_T$
is zero, we obtain that their Finsler distance
$d_F$
is zero. But this implies that the metrics are isometric.
Acknowledgements
We warmly thank the referees for their many helpful comments. In particular, one of the referees suggested a short and elegant argument to show that
$L_g/L_{g_0} \rightarrow 1$
implies
$L_g=L_{g_0}$
(see Appendix A). This project has received funding from the European Research Council (ERC) under the European Union’s Horizon 2020 research and innovation program (grant agreement No. 725967). This material is based upon work supported by the National Science Foundation under grant no. DMS-1440140 while C.G. and T.L. were in residence at the Mathematical Sciences Research Institute in Berkeley, California, during the Fall 2019 semester. The second author was partially supported by the SFB/TRR 191 ‘Symplectic structures in geometry, algebra and dynamics’.
A Appendix. Asymptotic marked length spectrum
In this appendix, we show the following lemma (the proof was communicated to us by one of the referees).
Lemma A.1. Let g and
$g_0$
be two metrics with Anosov geodesic flows on a fixed manifold M and assume that g is close to
$g_0$
in
$C^{k,\alpha }$
norm. Assume that for all sequences
$(c_j)_{j \geq 0}$
in
$\mathcal {C}$
,
$L_g(c_j)/L_{g_0}(c_j) \rightarrow _{j \rightarrow +\infty } 1$
. Then
$L_g = L_{g_0}$
.
Proof By Sigmund [Reference SigmundSig72, Theorem 1], the set
$\mathfrak {D} := \{\delta _{g_0}(c) ~|~ c \in \mathcal {C} \}$
is dense in
$\mathfrak {M}_{\textrm{inv}}$
(the set of invariant measures by the
$g_0$
-geodesic flow on
$S_{g_0}M$
). If
$\mu \in \mathfrak {M}_{\textrm{inv}} \setminus \mathfrak {D}$
, we can therefore find a sequence such that
$\delta _{g_0}(c_j) \rightharpoonup _{j \rightarrow +\infty } \mu $
and
$L_{g_0}(c_j) \rightarrow +\infty $
. (Indeed, if
$L_{g_0}(c_j) \leq C$
for some
$C \geq 0$
, then the sequence
$(\delta _{g_0}(c_j))_{j \geq 0}$
only achieves a finite number of measures, which would imply that
$\mu $
is a Dirac mass on a closed orbit and this is excluded since
$\mu \notin \mathfrak {D}$
.) Then, the condition
$L_g/L_{g_0} \rightarrow 1$
immediately implies that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu144.png?pub-status=live)
Now, for
$c \in \mathcal {C}$
and
$t> 0$
small, the linear combination
$t \mu + (1-t)\delta _{g_0}(c) \notin \mathfrak {D}$
. Indeed, if not, we would have
$t \mu + (1-t)\delta _{g_0}(c) = \delta _{g_0}(c_t)$
but by continuity,
$c_t=c_0$
for t small, which contradicts
$\mu \notin \mathfrak {D}$
. Therefore
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu145.png?pub-status=live)
that is,
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu146.png?pub-status=live)
B Appendix. Global conjugacy for Riemannian Anosov flows
Let
$(M, g)$
be a closed Riemannian manifold whose geodesic flow is Anosov. As has been shown by Klingenberg [Reference KlingenbergKli74] the geodesic flow has no conjugate points. Let
$(\widetilde M, g)$
be the universal cover of M where for simplicity the lifted metric is also denoted by
$ g$
. Let
$\Gamma $
be the group of deck transformations. As has been remarked in [Reference KnieperKni12], the universal cover
$\widetilde M$
is Gromov hyperbolic (see [Reference Bridson and HaefligerBH99, Section III.H.1] for a definition of Gromov hyperbolicity). Denote by
$\partial _{\infty } \widetilde M$
the Gromov boundary which is equipped with the visibility topology (see, for example, [Reference Knieper, Hasselblat and KatokKni02] for more details). For
$\xi \in \partial _{\infty } \widetilde M$
and
$x_0 \in \widetilde M$
, the Busemann function
$x \mapsto b_\xi ^g(x_0, x)$
is defined by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqn46.png?pub-status=live)
It has the following properties:
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqn47.png?pub-status=live)
and
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqn48.png?pub-status=live)
for all
$\gamma \in \Gamma $
. We introduce the gradient of the Busemann function
$B^g(x, \xi ) : = \nabla _{x} b_\xi ^g(x_0,x)$
which is independent of
$x_0$
by property (B.2). Also observe that
$B^g(x,\xi ) \in S_g \widetilde M$
by the very definition (B.1). Here,
$S_g \widetilde M$
is the unit tangent bundle on the universal cover and
$\pi : S_g \widetilde M \rightarrow \widetilde M$
denotes the projection. Given
$z=(x,v) \in S_g \widetilde M $
, we introduce
$c_g(z,t) := \pi (\varphi _t^g(x,v))$
, where
$(\varphi _t^g)_{t \in \mathbb {R}}$
is the (lift of) the geodesic flow on
$\widetilde M$
. We set
$z^g_\pm = c_g(z,\pm \infty ) \in \partial _\infty \widetilde M$
.
For
$\xi=z^g_+$
the submanifolds
$ W^{ss}(z)=\{(x,- B^g(x, \xi )) \in S_g \widetilde M \mid b_\xi ^g(x_0,x) = b_\xi ^g(x_0,\pi z) \} $
and
$W^{uu}(z) = \{ (x,B^g(x, \xi )) \in S_g \widetilde M \mid b_\xi ^g(x_0,x) = b_\xi ^g(x_0,\pi z) \}$
are the lifts of the leafs of strong stable and unstable foliations through
$z \in S_g \widetilde M$
. Since the leafs are smooth and the foliations are Hölder continuous, the Busemann functions
$(x,\xi ) \mapsto b_\xi ^g(x_0, x)$
are smooth with respect to x and Hölder continuous with respect to
$\xi $
. The following lemma was proved in [Reference Schapira and TapieSTar] (see also [Reference GromovGro00]) in negative curvature.
Lemma B.1. Let
$M = \widetilde M/ \Gamma $
be a closed manifold, and let
$g_1, g_2$
be two Riemannian metrics with Anosov geodesic flow. Consider the map
$\psi _{g_1, g_2}: S_{g_1}\widetilde M \to S_{g_2}\widetilde M $
defined by
$\psi _{g_1, g_2}(z) =w$
where
$w \in S^{g_2}\widetilde M $
is the unique vector with
$w^{g_2}_+ = z^{g_1}_+$
and
$w^{g_2}_-= z^{g_1}_-$
and
$b^{g_2}_{z^{g_1}_+}( \pi (z), \pi (w)) =0$
. Then
$\psi _{g_1, g_2}$
is a Hölder-continuous surjective map with
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu147.png?pub-status=live)
where
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu148.png?pub-status=live)
for all
$z \in S^{g_1}\widetilde M $
. Furthermore, for all
$\gamma \in \Gamma $
we have
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu149.png?pub-status=live)
and
${\tau (\gamma _{\ast }z,t)} = \tau (z,t)$
and therefore
$\psi _{g_1, g_2}$
descends to a map between the quotients
$S_{g_i}M $
.
Proof We show first that for each
$(z,t) \in S_{g_1}\widetilde M \times \mathbb {R}$
we have
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu150.png?pub-status=live)
where
$\tau (z,t) = b^{g_2}_{z^{g_1}_+} ( \pi (z),\pi ( \widetilde \varphi ^{g_1}_t(z))$
. From the cocycle property (B.2) of the Busemann function we obtain
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu151.png?pub-status=live)
By the definition of
$\psi _{g_1, g_2}$
this yields
$ \widetilde \varphi ^{g_2}_{\tau (z,t)}\psi _{g_1, g_2}(z) = \psi _{g_1, g_2}(\widetilde \varphi ^{g_1}_t(z))$
. The regularity of the Busemann function shows that
$\psi _{g_1, g_2}$
is Hölder continuous. The remaining assertions follow from the
$\Gamma $
-equivariance (B.3) of the Busemann function.
Remark B.2. Note that
$\psi _{g_1, g_2}: S_{g_1}M \to S_{g_2}M$
maps orbits of the geodesic flow of
$g_1$
surjectively onto orbits of the geodesic flow of
$g_2$
but is not necessarily injective. To obtain a injective map the following modification due to Gromov [Reference GromovGro00] (see also [Reference Knieper, Hasselblat and KatokKni02]) can be made. Choose
$r_0>0$
such that
$\tau (z,r_0)>0$
for all
$z \in S_{g_1}M $
. Define
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu152.png?pub-status=live)
and consider
$\psi _{g_1, g_2}^r (z) := \varphi ^{g_2}_{r(z)} \circ \psi _{g_1, g_2}(z)$
Then
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu153.png?pub-status=live)
Since
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu154.png?pub-status=live)
and
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu155.png?pub-status=live)
the map
$t \mapsto \hat \tau (z,t)$
is strictly monotone increasing and therefore
$\psi _{g_1, g_2}^r$
is injective and yields a conjugacy between the geodesic flows.
C Appendix. Anosov stability
The proof of the Anosov stability theorem is given using the implicit function theorem in [Reference de la Llave, Marco and MoriyóndlLMM86] in the
$C^0$
category; the extension to the Hölder setting (with the same method) appears in [Reference Katok, Knieper, Pollicott and WeissKKPW89]. We need the continuity with respect to the two metrics here; the proof of [Reference de la Llave, Marco and MoriyóndlLMM86, Reference Katok, Knieper, Pollicott and WeissKKPW89] indeed shows this, as we explain below. Let
$\nu \in (0,1)$
. Then if X is a
$C^k$
vector field for
$k\geq 4$
with flow
$\varphi _t^X$
, we will denote by
$C_{X}^\nu (\mathcal {M},\mathcal {M})$
the space of
$C^\nu $
maps
$\psi $
on a closed manifold
$\mathcal {M}$
so that
$d\psi .X:=\partial _{t}(\psi \circ \varphi ^X_t)|_{t=0}$
exists and belongs to
$C^\nu (\mathcal {M};T\mathcal {M})$
. This is a Banach manifold [Reference Katok, Knieper, Pollicott and WeissKKPW89, Proposition 2.2].
Proposition C.1. Let
$g_0$
be a smooth metric, and assume that
$X_{g_0}$
its geodesic vector field on
$\mathcal {M}:=S_{g_0}M$
is Anosov. We view all geodesic vector fields
$X_g$
associated to g near
$g_0$
as vector fields on
$\mathcal {M}$
(by pulling back from
$S_gM$
to
$S_{g_0}M$
). For
$k\geq 4$
, there exist
$\nu>0$
and two open neighborhoods
$\mathcal {U}_0\subset \mathcal {U}$
of
$X_{g_0}$
in
$C^{k+1}(\mathcal {M};T\mathcal {M})$
such that, for each
$Y\in \mathcal {U}$
and each
$g\in C^{k+2}(M;S^2T^*M)$
so that
$X_g\in \mathcal {U}_0$
, there exist a homeomorphism
$\psi _{g,Y}\in C_{X_{g}}^{\nu }(\mathcal {M},\mathcal {M})$
and
$a_{g,Y}\in C^{\nu }(\mathcal {M},\mathbb {R}^+)$
such that
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu156.png?pub-status=live)
where
$X_g$
is the geodesic vector field of g. Moreover,
$Y\in \mathcal {U}\mapsto a_{g,Y}\in C^\nu (\mathcal {M},\mathbb {R}^+)$
and
$Y\mapsto \psi _{g,Y}\in C^\nu _{X_g}(\mathcal {M},\mathcal {M})$
are
$C^{k}$
, and each derivative of order
$\ell \leq k$
with respect to Y is continuous with respect to
$(g,Y)$
with values in
$C^{\nu }$
.
Proof The proof is essentially contained in [Reference Katok, Knieper, Pollicott and WeissKKPW89, Proposition 2.2], except for the statement about the continuity with respect to
$X_g$
. Consider, for
$\nu \in (0,1) \mbox{ and } E := C^{k+1}(\mathcal{M};T\mathcal{M}) \times C^\nu(\mathcal{M},T\mathcal{M})$
, the map
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu157.png?pub-status=live)
defined by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu158.png?pub-status=live)
This is a
$C^k$
map between Banach manifolds. The differential of
$F^{X_g}$
at
$(X_g,{\textrm{Id}},1)$
is given (as in [Reference Katok, Knieper, Pollicott and WeissKKPW89]) by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqn49.png?pub-status=live)
where
$V\in C_{X_g}^{\nu }(\mathcal {M};T\mathcal {M}):=\{ V\in C^\nu (\mathcal {M};T\mathcal {M})\,|\, \mathcal {L}_{X_g}V\in C^{\nu }\}$
. Let
$\alpha _g$
be the contact form of g, so that
$\ker \alpha _g=E_u(g)\oplus E_s(g)$
is the smooth bundle of stable or unstable vectors for g. By [Reference Katok, Knieper, Pollicott and WeissKKPW89, Proposition 2.2. and Lemma 2.3], the operator
$\mathcal {L}_{X_g}: V\mapsto \mathcal {L}_{X_g}V$
is invertible from
$C^{\nu }_{X_g}(\mathcal {M};\ker \alpha _g)\to C^\nu (\mathcal {M};\ker \alpha _g)$
for some
$\nu $
depending on the maximal/minimal expansion rates of the flow
$\varphi _t^{X_g}$
. The inverse is given by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu159.png?pub-status=live)
where the integrals converge due to the contraction of the differential: for all
$t\geq 0$
,
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqn50.png?pub-status=live)
This operator maps continuously
$C^\nu (\mathcal {M};\ker \alpha _g)$
to
$C^{\nu }_{X_g}(\mathcal {M};\ker \alpha _g)$
if
$\nu>0$
is small enough, depending on
$\unicode{x3bb} _+$
and
$\|\varphi _T^{X_g}\|_{C^2}$
for
$T>0$
large (see below). Moreover, by continuity of the bundles
$E_s(g),E_u(g)$
with respect to g [Reference Hirsch, Pugh, Chern and SmaleHP68, Theorem 3.2], for g close enough to
$g_0$
in
$C^{k+5}$
,
$E_u(g)$
and
$E_s(g)$
are contained in a small conic neighborhood of
$E_u(g_0)$
and
$E_s(g_0)$
respectively, and the contraction exponents
$\unicode{x3bb} _\pm (g)$
are also close to
$\unicode{x3bb} _\pm (g_0)$
(see, for example, [Reference BonthonneauBon20, Lemma 3]), so this will give the boundedness of
$\mathcal {L}_{X_g}^{-1}$
in
$C^\nu $
for some fixed
$\nu>0$
for g close enough to
$g_0$
in
$C^{k+5}$
. From the expression of
$\mathcal {L}_{X_g}^{-1}$
, and the fact that (C.2) holds uniformly for g close to
$g_0$
for some
$0<\unicode{x3bb} _-<\unicode{x3bb} _+$
(and similarly on
$E_u(g)$
), we claim that, if
$\pi _{g}:T\mathcal {M}\to \ker \alpha _g$
is the projection given by
$\pi _{g}(V)=V-\alpha _{g}(V)X_g$
, then
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu160.png?pub-status=live)
is continuous with respect to g (in
$C^{k+5}$
) for
$\nu>0$
small enough. To prove this, we rewrite
$\mathcal {L}_{X_g}^{-1}\pi _g$
as
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqn51.png?pub-status=live)
where
$\pi _g^u: C^\nu (\mathcal {M};T\mathcal {M})\to C^\nu (\mathcal {M};T\mathcal {M})$
is the projection on
$E_u$
parallel to
$E_s$
and
$\pi _g^s: C^\nu (\mathcal {M};T\mathcal {M})\to C^\nu (\mathcal {M};T\mathcal {M})$
is the projection on
$E_s$
parallel to
$E_u$
, and
$e^{t\mathcal {L}_{X_g}} Y := d\varphi _{-t}^{X_g} Y \circ \varphi _t^{X_g}$
is the propagator. Here
$\nu $
is chosen small so that
$E_u$
and
$E_s$
are
$C^\nu $
bundles (see [Reference Hirsch, Pugh, Chern and SmaleHP68]), and by [Reference ContrerasCon92] the maps
$g\mapsto \pi _g^u$
and
$g\mapsto \pi _g^s$
are continuous (actually
$C^r$
for some r depending on the smoothness of g). Next, there exist
$C>0$
and
$\Lambda>0$
such that, for all t,
$\|\varphi ^{X_g}_t\|_{C^2}\leq Ce^{\Lambda |t|}$
for all g near
$g_0$
in
$C^{k+5}$
, which implies that, for all
$V\in C^1(\mathcal {M};T\mathcal {M})$
,
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu161.png?pub-status=live)
thus if
$\nu _0\in (0,1)$
is such that
$E_u\in C^{\nu _0}$
, we have by interpolation that
$\|e^{t\mathcal {L}_{X_g}}\|_{\mathcal {L}(C^{\nu })}\leq Ce^{(1+\nu _0) \Lambda |t|}$
for each
$\nu \leq \nu _0$
. Since
$\|e^{t\mathcal {L}_{X_g}}\pi _g^u\|_{\mathcal {L}(C^0)}+\|e^{-t\mathcal {L}_{X_g}}\pi _g^s\|_{\mathcal {L}(C^0)}\leq Ce^{-\unicode{x3bb} _- t}$
for all
$t\geq 0$
, we obtain by interpolating
$C^\nu $
between the spaces
$C^0$
and
$C^{\nu _0}$
with
$\nu = \theta \times 0 + (1-\theta ) \times \nu _0$
(for
$\theta \in (0,1)$
) that, for all
$t\geq 0$
,
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu162.png?pub-status=live)
We can now fix
$\nu $
small enough (that is,
$\theta $
close enough to
$1$
) to guarantee
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu163.png?pub-status=live)
which implies that (C.3) is uniformly converging with respect to g near
$g_0$
in
$C^{k+5}$
. Since
$g\mapsto e^{t\mathcal {L}_{X_g}}\pi _g^u$
and
$g\mapsto e^{-t\mathcal {L}_{X_g}}\pi _g^s$
are continuous for each
$t\geq 0$
, we can apply the Lebesgue theorem to deduce the continuity of
$g\mapsto \mathcal {L}_{X_g}^{-1}\pi _g\in \mathcal {L}(C^\nu )$
for
$\nu>0$
small enough.
Next, we consider the map
$\widetilde {F}^{X_g}: E\rightarrow E$
defined by
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu164.png?pub-status=live)
where we recall that
$E = C^{k+1}(\mathcal {M};T\mathcal {M}) \times C^\nu (\mathcal {M},T\mathcal {M}) $
and
$\exp _{g_0}$
is the exponential map of
$g_0$
. This map satisfies
$\widetilde {F}^{X_g}(X_g,0)=(X_g,0)$
. We want to apply the inverse function theorem to find a pre-image to each
$(Y,0)$
close to
$(X_g,0)$
. As in [Reference Katok, Knieper, Pollicott and WeissKKPW89, Proposition 2.2] (see also [Reference de la Llave, Marco and MoriyóndlLMM86, Appendix A]), the map
$\widetilde {F}^{X_g}$
is
$C^{k}$
, and moreover it depends continuously on
$g\in C^{k+5}(M;S^2 T^*M)$
, with all its derivatives of order
$\ell \leq k$
being also continuous with respect to g, due to the continuity of
$g\mapsto \mathcal {L}_{X_g}^{-1}\pi _g$
as a map
$C^{k+5}(M;S^2 T^*M)\to \mathcal {L}(C^\nu (\mathcal {M};T\mathcal {M}))$
. Now we have
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu165.png?pub-status=live)
by using (C.1) and
$\pi _g(X_g)=0$
. In particular, there is
$\varepsilon>0$
such that if
$\|g-g_0\|_{C^{k+5}}<\varepsilon $
,
$\|Y-X_g\|_{C^{k+1}}<\varepsilon $
and
$\|V\|_{C^\nu }<\varepsilon $
, then
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu166.png?pub-status=live)
For each Y close to
$X_g$
, we can then apply the fixed point theorem (as in the proof of the inverse function theorem) to the map
$(Z,V)\in E\mapsto (Z+Y,V)-\widetilde {F}^{X_g}(Z,V)$
and obtain that there is a unique
$(Y,V(Y))$
such that
$(Y,0)=\widetilde {F}^{X_g}(Y,V(Y))$
, and
$V(Y)\in C^{\nu }(\mathcal {M};T\mathcal {M})$
depends in a
$C^{k}$
fashion on Y and is continuous with respect to g. Moreover, the usual argument in the inverse function theorem used to prove the
$C^k$
property of
$Y\mapsto V(Y)$
also shows that the derivatives of order
$\ell \leq k$
are continuous with respect to
$(X_g,Y)$
, by using the continuity of
$\widetilde {F}^{X_g}$
and its derivatives with respect to g. This shows that for each Y close to
$X_g$
in
$C^{k+1}$
norm and g close to
$g_0$
in
$C^{k+5}$
norm, there is a
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu167.png?pub-status=live)
so that
$\gamma du.X_g=Y\circ u$
, with
![](https://static.cambridge.org/binary/version/id/urn:cambridge.org:id:binary:20220210092230446-0959:S0143385721000754:S0143385721000754_eqnu168.png?pub-status=live)
and where
$C^{k}$
and all the derivatives of order
$\ell \leq k$
are continuous in
$(g,Y)$
(with values in
$C^\nu (\mathcal {M},\mathcal {M})\times C^\nu (\mathcal {M})$
).