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

Dynamical Zeta Functions in the Nonorientable Case

Yonah Borns-Weil Department of Mathematics, University of California, Berkeley, CA 94720 [email protected]  and  Shu Shen Institut de Mathématiques de Jussieu-Paris Rive Gauche, Sorbonne Université, 4 place Jussieu, 75252 Paris Cedex 5, France. [email protected]
Abstract.

We use a simple argument to extend the microlocal proofs of meromorphicity of dynamical zeta functions to the nonorientable case. In the special case of geodesic flow on a connected non-orientable negatively curved closed surface, we compute the order of vanishing of the zeta function at the zero point to be the first Betti number of the surface.

1. Background

In this note we use a simple geometric argument to extend the results of Dyatlov and Zworski [5, 6] and of Dyatlov and Guillarmou [3, 4] to Axiom A flows with nonorientable stable and unstable bundles. It is classically known that on a closed manifold there are countably many closed orbits of such flows, and therefore one can define the Ruelle zeta function

ζR(λ)=γ(1eiλTγ)\zeta_{R}(\lambda)=\prod_{\gamma^{\sharp}}\left(1-e^{i\lambda T_{\gamma}^{\sharp}}\right)

where the product is taken over all primitive closed geodesics γ\gamma^{\sharp} with corresponding periods TγT_{\gamma}^{\sharp}. Note that by [3, Lemma 1.17] and [4, Section 3], this product converges for Im(λ)1\mathrm{Im\,}(\lambda)\gg 1 large enough. The meromorphic continuation of ζR\zeta_{R} to all of \mathbb{C} was conjectured by Smale [13], and proved by Fried [8] under analyticity assumptions. The case of smooth Anosov flows was first answered by Giulietti, Liverani and Policott [9] and then with microlocal methods by Dyatlov and Zworski [5] for manifolds with orientable stable and unstable bundles, and was extended to Axiom A flows by Dyatlov and Guillarmou [3, 4] under the same orientability assumptions. In [9, Appendix B], the authors also outlined ideas for removing the orientability assumptions.

We remove the orientability assumption and give a full proof for Axiom A flows. Specifically, we shall show

Theorem 1.

If (ϕt)t(\phi_{t})_{t\in\mathbb{R}} is an Axiom A flow on a closed manifold, the Ruelle zeta function ζR\zeta_{R} extends to a meromorphic function on \mathbb{C}.

The definition of an Axiom A flow is given as Definition 1.3.

We then restrict to the case of contact Anosov flow on a 33-manifold, and study the order of vanishing of ζR\zeta_{R} at λ=0\lambda=0. An important example is when M=SΣM=S^{*}\Sigma, the cosphere bundle of a connected negatively curved closed surface Σ\Sigma, and (ϕt)t(\phi_{t})_{t\in\mathbb{R}} is geodesic flow [1]. This problem was treated in [6] in the case where the stable bundle is orientable, and it was shown that the order of vanishing is b1(M)2b_{1}(M)-2, where b1(M)b_{1}(M) is the first Betti number of MM.

We shall show that for nonorientable stable bundle, the analogous result is the following:

Theorem 2.

Let (ϕt)t(\phi_{t})_{t\in\mathbb{R}} be the Reeb flow on a connected contact closed 33-manifold. If (ϕt)t(\phi_{t})_{t\in\mathbb{R}} is Anosov with nonorientable stable bundle EsE_{s}, the Ruelle zeta function has vanishing order at λ=0\lambda=0 equal to b1(𝒪(Es))b_{1}(\mathscr{O}(E_{s})), the dimension of the first de Rham cohomology with coefficients in the orientation line bundle of EsE_{s}.

The orientation line bundle is reviewed in Definition 1.5.

In the special case of the geodesic flow on M=SΣM=S^{*}\Sigma with Σ\Sigma nonorientable, the vanishing order at λ=0\lambda=0 is given by b1(Σ)b_{1}(\Sigma), as is shown in Proposition 3.10. This is in contrast to the orientable case, in which it is b1(Σ)2b_{1}(\Sigma)-2.

More precisely, let χ(Σ)\chi^{\prime}(\Sigma) be the derived Euler characteristic of Σ\Sigma, i.e.,

χ(Σ)=i=02(1)iibi(Σ)={b1(Σ)+2,if Σ is orientable,b1(Σ),otherwise.\chi^{\prime}(\Sigma)=\sum_{i=0}^{2}(-1)^{i}ib_{i}(\Sigma)=\begin{cases}-b_{1}(\Sigma)+2,&\text{if $\Sigma$ is orientable},\\ -b_{1}(\Sigma),&\text{otherwise}.\end{cases}
Corollary 3.

If (ϕt)t(\phi_{t})_{t\in\mathbb{R}} is the geodesic flow on the cosphere bundle of a connected negatively curved closed surface (orientable or not), the Ruelle zeta function has vanishing order at λ=0\lambda=0 equal to χ(Σ)-\chi^{\prime}(\Sigma).

1.1. Axiom A Flows

Let MM be a compact manifold without boundary of dimension nn, and let (ϕt)t(\phi_{t})_{t\in\mathbb{R}} be a flow on MM generated by the vector field VC(M;TM)V\in C^{\infty}(M;TM).

Definition 1.1.

A ϕt\phi_{t}-invariant set KMK\subseteq M is called hyperbolic for the flow (ϕt)t(\phi_{t})_{t\in\mathbb{R}} if VV does not vanish on KK and for each xKx\in K the tangent space TxMT_{x}M can be written as the direct sum

TxM=E0(x)Es(x)Eu(x)T_{x}M=E_{0}(x)\oplus E_{s}(x)\oplus E_{u}(x)

where E0(x)=span(V(x))E_{0}(x)=\text{span}(V(x)), Es,EuE_{s},E_{u} are continuous ϕt\phi_{t}-invariant vector bundles on KK, and for some Riemannian metric |||\cdot|, there are C,θ>0C,\theta>0 such that for all t>0t>0,

|dϕt(x)v|ϕt(x)Ceθt|v|xvEs(x)|dϕt(x)w|ϕt(x)Ceθt|w|xwEu(x).\displaystyle\begin{split}|d\phi_{t}(x)v|_{\phi_{t}(x)}&\leq Ce^{-\theta t}|v|_{x}\qquad v\in E_{s}(x)\\ |d\phi_{-t}(x)w|_{\phi_{-t}(x)}&\leq Ce^{-\theta t}|w|_{x}\quad\,\,\,\,w\in E_{u}(x).\end{split} (1)

In the important case where all of MM is hyperbolic, we call (ϕt)t(\phi_{t})_{t\in\mathbb{R}} an Anosov flow.

There is an analogous notion of hyperbolicity at fixed points.

Definition 1.2.

A fixed point xMx\in M, i.e., V(x)=0V(x)=0, is called hyperbolic if the differential DV(x)DV(x) has no eigenvalues with vanishing real part.

A generalization of Anosov flows is the following, given first by Smale [13, II.5 Definition 5.1]:

Definition 1.3.

The flow (ϕt)t(\phi_{t})_{t\in\mathbb{R}} is called Axiom A if

  1. (1)

    all fixed points of (ϕt)t(\phi_{t})_{t\in\mathbb{R}} are hyperbolic,

  2. (2)

    the closure 𝒦\mathcal{K} of the union of all closed orbits of (ϕt)t(\phi_{t})_{t\in\mathbb{R}} is hyperbolic,

  3. (3)

    the nonwandering set ([4, Definition 2.2]) of (ϕt)t(\phi_{t})_{t\in\mathbb{R}} is the disjoint union of the set of fixed points and 𝒦\mathcal{K}.

We now recall the definition of a locally maximal set, given in [4, Definition 2.4].

Definition 1.4.

A compact ϕt\phi_{t}-invariant set KMK\subseteq M is called locally maximal for (ϕt)t(\phi_{t})_{t\in\mathbb{R}} if there is a neighborhood VV of KK such that

K=tϕt(V).K=\bigcap_{t\in\mathbb{R}}\phi_{t}(V).

We may then state the key proposition, which generalises [4, Proposition 3.1] to the case where EsE_{s} or EuE_{u} is not necessarily orientable on 𝒦\mathcal{K}.

Proposition 1.1.

Let KMK\subseteq M be a locally maximal hyperbolic set for (ϕt)t(\phi_{t})_{t\in\mathbb{R}}, and let ζK\zeta_{K} be defined as the Ruelle zeta function where we only take the product over trajectories in KK. Then ζK\zeta_{K} has a continuation to a meromorphic function on all of \mathbb{C}.

Theorem 1 follows from Proposition 1.1, as we may remark that by [13, II.5 Theorem 5.2] we can write 𝒦=K1KN\mathcal{K}=K_{1}\sqcup\dots\sqcup K_{N} with KjK_{j} basic hyperbolic111These are locally maximal hyperbolic by definition (see [4, Definition 2.5]).. Then the product

ζR(λ)=j=1NζKj(λ),\zeta_{R}(\lambda)=\prod_{j=1}^{N}\zeta_{K_{j}}(\lambda),

which a priori holds for Im(λ)1\mathrm{Im\,}(\lambda)\gg 1, gives that ζR\zeta_{R} also has a meromorphic continuation to all of \mathbb{C}.

The goal of Section 2 is to prove Proposition 1.1.

1.2. The Orientation Bundle

To fix notation we recall the definition of transition functions of a vector bundle. Given a continuous real vector bundle EE of rank kk over a manifold MM with projection map π\pi, let Uα,UβMU_{\alpha},U_{\beta}\subseteq M be two small open sets with nonempty intersection, and let ψα:π1UαUα×n\psi_{\alpha}:\pi^{-1}U_{\alpha}\to U_{\alpha}\times\mathbb{R}^{n}, ψβ:π1UβUβ×n\psi_{\beta}:\pi^{-1}U_{\beta}\to U_{\beta}\times\mathbb{R}^{n} be local trivializations. Then the map ψαψβ1:(UαUβ)×n(UαUβ)×n\psi_{\alpha}\circ\psi_{\beta}^{-1}:(U_{\alpha}\cap U_{\beta})\times\mathbb{R}^{n}\to(U_{\alpha}\cap U_{\beta})\times\mathbb{R}^{n} is of the form

ψαψβ1(p,v)=(p,ταβ(p)v)\psi_{\alpha}\circ\psi_{\beta}^{-1}(p,v)=(p,\tau_{\alpha\beta}(p)v)

where ταβC0(UαUβ,GLk())\tau_{\alpha\beta}\in C^{0}(U_{\alpha}\cap U_{\beta},{\rm GL}_{k}(\mathbb{R})) is called a transition function. If the local trivializations can be chosen such that ταβ\tau_{\alpha\beta} are smooth, then EE is a smooth vector bundle. Similarly, if ταβ\tau_{\alpha\beta} can be chosen to be locally constant functions, then EE is a flat vector bundle.

Furthermore, suppose we are given an open cover (Uα)αA(U_{\alpha})_{\alpha\in A} of MM together with a set of continuous (resp. smooth, resp. locally constant) GLk(){\rm GL}_{k}(\mathbb{R})-valued functions (ταβ)α,βAUαUβ(\tau_{\alpha\beta})_{\begin{subarray}{c}\alpha,\beta\in A\\ U_{\alpha}\cap U_{\beta}\neq\emptyset\end{subarray}} with ταα=I\tau_{\alpha\alpha}=I on UαU_{\alpha}. Then there exists a continuous (resp. smooth, resp. flat) vector bundle EE with transition functions ταβ\tau_{\alpha\beta}, provided the following triple product property holds:

ταβ(p)τβγ(p)τγα(p)=I\tau_{\alpha\beta}(p)\tau_{\beta\gamma}(p)\tau_{\gamma\alpha}(p)=I

for any pUαUβUγ.p\in U_{\alpha}\cap U_{\beta}\cap U_{\gamma}.

Definition 1.5.

If EE is a continuous (but not necessarily smooth) real vector bundle over MM with transition functions ταβ\tau_{\alpha\beta}, the orientation bundle of EE is a smooth flat line bundle 𝒪(E)\mathscr{O}(E) with transition functions

σαβ(p)=sgndet(ταβ(p))={1det(ταβ(p))>01det(ταβ(p))<0.\sigma_{\alpha\beta}(p)={\rm sgn}\det(\tau_{\alpha\beta}(p))=\begin{cases}1&\det(\tau_{\alpha\beta}(p))>0\\ -1&\det(\tau_{\alpha\beta}(p))<0.\end{cases}

Recall that if f:MMf:M\to M is a map, we say ff lifts to a bundle map F:EEF:E\to E if πF=fπ\pi\circ F=f\circ\pi.

Since 𝒪(E)\mathscr{O}(E) is a flat vector bundle, using the associated flat connection, we can lift the flow (ϕt)t(\phi_{t})_{t\in\mathbb{R}} to a flow (Φ~t)t(\widetilde{\Phi}_{t})_{t\in\mathbb{R}} on 𝒪(E)\mathscr{O}(E). If the flow (ϕt)t(\phi_{t})_{t\in\mathbb{R}} on MM lifts to a flow (Φt)t(\Phi_{t})_{t\in\mathbb{R}} on EE, if ψ,η\psi,\eta are distinct trivializations of EE near pp, ϕt(p)\phi_{t}(p) respectively, and ψ~,η~\widetilde{\psi},\widetilde{\eta} are trivializations of 𝒪(E)\mathscr{O}(E) near pp, ϕt(p)\phi_{t}(p) respectively, we have for pMp\in M and l𝒪(E)pl\in\mathscr{O}(E)_{p}:

Φ~t(l)=η~1(ϕt(p),sgn(det(ηΦtψ1)|p)proj2ψ~(l)),\displaystyle\widetilde{\Phi}_{t}(l)=\widetilde{\eta}^{-1}\left(\phi_{t}(p),\mathrm{sgn}\,\left(\det\left.\left(\eta\Phi_{t}\psi^{-1}\right)\right|_{p}\right)\mathrm{proj}_{2}\widetilde{\psi}(l)\right), (2)

where proj2\mathrm{proj}_{2} is the obvious projection to the second component.

1.3. Geodesic flows

Let ZZ be a negatively curved closed Riemannian manifold. Let M=SZM=S^{*}Z be the cosphere bundle on ZZ. It is classical that the geodesic flow on MM is Anosov [1].

Let π:MZ\pi:M\to Z be the canonical projection. For xMx\in M, we have a morphism of linear spaces

π:TxMTπ(x)Z.\displaystyle\pi_{*}:T_{x}M\to T_{\pi(x)}Z. (3)

The following proposition is classical [1, Section 22] and [12, Proposition 6]. We include a proof for the sake of completeness.

Proposition 1.2.

The morphism π\pi_{*} induces an isomorphism of continuous vector bundles on MM,

EsE0π(TZ).\displaystyle E_{s}\oplus E_{0}\simeq\pi^{*}(TZ). (4)
Proof.

Since both sides of (4) have the same dimension, it is enough to show that π|EsE0\pi_{*}|_{E_{s}\oplus E_{0}} is injective. We will show this using Jacobi fields. It is convenient to work on the sphere bundle M=SZM^{\prime}=SZ. We identify MM^{\prime} with MM via the Riemannian metric on ZZ.

We follow [7, Section II.H]. Let \mathcal{M} be the total space of TZTZ. Denote still by π:Z\pi:\mathcal{M}\to Z the obvious projection. Let TVTT^{V}\mathcal{M}\subset T\mathcal{M} be the vertical subbundle of TT\mathcal{M}. The Levi-Civita connection on TZTZ induces a horizontal subbundle THTT^{H}\mathcal{M}\subset T\mathcal{M} of TT\mathcal{M}, so that

T=TVTH.\displaystyle T\mathcal{M}=T^{V}\mathcal{M}\oplus T^{H}\mathcal{M}. (5)

Since TVπ(TZ)T^{V}\mathcal{M}\simeq\pi^{*}(TZ) and THπ(TZ)T^{H}\mathcal{M}\simeq\pi^{*}(TZ), by (5), we can identify the smooth vector bundles,

T=π(TZTZ).\displaystyle T\mathcal{M}=\pi^{*}(TZ\oplus TZ).

For x=(z,v)x={(z,v)}\in\mathcal{M}, let γx\gamma_{x} be the unique geodesic on ZZ such that (γx(0),γ˙x(0))=(z,v).(\gamma_{x}(0),\dot{\gamma}_{x}(0))=(z,v). For wTxw\in T_{x}\mathcal{M}, let Jx,wC(γx,TZ|γx)J_{x,w}\in C^{\infty}(\gamma_{x},TZ|_{\gamma_{x}}) be the unique Jacobi field along γx\gamma_{x} such that (Jx,w(0),J˙x,w(0))=w,\left(J_{x,w}(0),\dot{J}_{x,w}(0)\right)=w, where J˙x,w\dot{J}_{x,w} is the covariant derivation of Jx,wJ_{x,w} in the direction γ˙x\dot{\gamma}_{x}. Recall that a Jacobi field JJ is called stable, if there is C>0C>0 such that for all t0t\geq 0,

|J(t)|C.\displaystyle\left|J(t)\right|\leq C.

By [7, Proposition VI.A], given xx\in\mathcal{M}, for any Y1TzZY_{1}\in T_{z}Z, there exists one and only one stable Jacobi field JJ along γx\gamma_{x} such that J(0)=Y1J(0)=Y_{1}.

For x=(z,v)Mx=(z,v)\in M^{\prime}, we have

TxM={(Y1,Y2)TzZTzZ:Y2,v=0}.\displaystyle T_{x}M^{\prime}=\{(Y_{1},Y_{2})\in T_{z}Z\oplus T_{z}Z:\langle Y_{2},v\rangle=0\}.

The morphism π\pi_{*} in (3) is just

wTxMJx,w(0)TzZ.\displaystyle w\in T_{x}M^{\prime}\to J_{x,w}(0)\in T_{z}Z.

By [7, Proposition VI.B], wEs(x)E0(x)w\in E_{s}(x)\oplus E_{0}(x) if and only if the Jacobi fields Jx,wJ_{x,w} is stable. By the uniqueness of stable Jacobi fields, we see that π|EsE0\pi_{*}|_{E_{s}\oplus E_{0}} is injective. ∎

Since E0E_{0} is a trivial line bundle, our proposition implies immediately:

Corollary 1.3.

We have the isomorphism of smooth flat line bundles

𝒪(Es)π(𝒪(TZ)).\displaystyle\mathscr{O}(E_{s})\simeq\pi^{*}(\mathscr{O}(TZ)).

2. Proof of Proposition 1.1

We use the notation of [5]. If 0kn10\leq k\leq n-1, let 0kΛk(TM)\mathcal{E}_{0}^{k}\subset\Lambda^{k}(T^{*}M) denote the subbundle of kk-forms ω\omega such that ιVω=0\iota_{V}\omega=0, where ι\iota denotes interior multiplication.

Let 0k~=0k𝒪(Es).\widetilde{\mathcal{E}_{0}^{k}}=\mathcal{E}_{0}^{k}\otimes\mathscr{O}(E_{s}). We consider the pullback ϕt\phi_{-t}^{*} on sections of 0k~\widetilde{\mathcal{E}_{0}^{k}}. Note that the flow (ϕt)t(\phi_{t})_{t\in\mathbb{R}} lifts to a flow (Φt)t(\Phi_{t})_{t\in\mathbb{R}} on 0k\mathcal{E}_{0}^{k}. Indeed, for pMp\in M, ω0,pk\omega\in\mathcal{E}_{0,p}^{k}, Φtω0,ϕt(p)k\Phi_{t}\omega\in\mathcal{E}_{0,\phi_{t}(p)}^{k} is defined for v1,,vkTϕt(p)Mv_{1},\cdots,v_{k}\in T_{\phi_{t}(p)}M by

Φtω(v1,,vk)=ω((dϕt|p)1v1,,(dϕt|p)1vk).\displaystyle\Phi_{t}\omega(v_{1},\dots,v_{k})=\omega\left((d\phi_{t}|_{p})^{-1}v_{1},\cdots,(d\phi_{t}|_{p})^{-1}v_{k}\right). (6)

Note that from the above formula, it is easy to check that ιVΦtω=0\iota_{V}\Phi_{t}\omega=0. Recall also that the flow (ϕt)t(\phi_{t})_{t\in\mathbb{R}} lifts to a flow Φ~t\widetilde{\Phi}_{t} on 𝒪(Es)\mathscr{O}(E_{s}) (see (2)). For a section ss in 0k~\widetilde{\mathcal{E}_{0}^{k}}, we have

ϕts(p)=(ΦtΦ~t)(s(ϕt(p))).\displaystyle\phi_{-t}^{*}s\,(p)=\left(\Phi_{t}\otimes\widetilde{\Phi}_{t}\right)\big{(}s(\phi_{-t}(p))\big{)}. (7)

Let χC(M)\chi\in C^{\infty}(M) be a smooth function whose support is contained in a small neighborhood of KK such that χ(x)=1\chi(x)=1 for all xKx\in K. We now invoke the Guillemin trace formula (see [11, pp. 501-502], [5, Appendix B], [3, (4.6)]) which says that the flat trace trχϕtχ|C(M;0k~)\mathrm{tr}^{\flat}\left.\chi\phi_{-t}^{*}\chi\right|_{C^{\infty}\left(M;\widetilde{\mathcal{E}_{0}^{k}}\right)} is a distribution on (0,)(0,\infty) given by

trχϕtχ|C(M;0k~)=γKTγtr~0,yk(ΦTγΦ~Tγ)|det(I𝒫γ)|δtTγ,\displaystyle\mathrm{tr}^{\flat}\left.\chi\phi_{-t}^{*}\chi\right|_{C^{\infty}\left(M;\widetilde{\mathcal{E}_{0}^{k}}\right)}=\sum_{\gamma\subset K}\frac{T_{\gamma}^{\sharp}\ \mathrm{tr}^{\widetilde{\mathcal{E}}_{0,y}^{k}}\left(\Phi_{T_{\gamma}}\otimes\widetilde{\Phi}_{T_{\gamma}}\right)}{|\det(I-\mathcal{P}_{\gamma})|}\delta_{t-T_{\gamma}}, (8)

where the sum is taken over all the periodic trajectories γ\gamma in KK with period TγT_{\gamma} and primitive period TγT_{\gamma}^{\sharp}, yy is any point on γ\gamma, and 𝒫γ=dϕTγ|(EsEu)y\mathcal{P}_{\gamma}=d\phi_{-T_{\gamma}}|_{(E_{s}\oplus E_{u})_{y}} is the linearized Poincaré map at yy. Note that as trace and determinant are invariant under conjugation, the right hand side does not depend on yy.

By (6), the trace of ΦTγ\Phi_{T_{\gamma}} on 0,yk\mathcal{E}_{0,y}^{k} is just tr(k𝒫γ)\mathrm{tr}\left(\bigwedge^{k}\mathcal{P}_{\gamma}\right). By (2), we may take trivializations ψ,ψ~\psi,\widetilde{\psi} of EsE_{s}, 𝒪(Es)\mathscr{O}(E_{s}) in a neighborhood of yy and have the induced lifting on 𝒪(Es)\mathscr{O}(E_{s}) to be sgn(det(ψdϕTγ|Es,yψ1)|).\mathrm{sgn}\,\left(\det\left.\left(\psi d\phi_{T_{\gamma}}|_{E_{s,y}}\psi^{-1}\right)\right|\right). By definition we get this to be equal to

sgn(detdϕTγ|Es,y)=sgn(detdϕTγ|Es,y)=sgndet(𝒫γ|Es),\mathrm{sgn}\,\left(\det\left.d\phi_{T_{\gamma}}\right|_{E_{s,y}}\right)=\mathrm{sgn}\,\left(\det\left.d\phi_{-T_{\gamma}}\right|_{E_{s,y}}\right)=\mathrm{sgn}\,\det\left(\left.\mathcal{P}_{\gamma}\right|_{E_{s}}\right),

and as it is a map between one dimensional spaces, the trace is given by that expression as well. By the above consideration, we can rewrite (8) as

trχϕtχ|C(M;0k~)=γKTγtr(k𝒫γ)sgn(det𝒫γ|Es)|det(I𝒫γ)|δtTγ.\displaystyle\mathrm{tr}^{\flat}\left.\chi\phi_{-t}^{*}\chi\right|_{C^{\infty}\left(M;\widetilde{\mathcal{E}_{0}^{k}}\right)}=\sum_{\gamma\subset K}\frac{T_{\gamma}^{\sharp}\ \mathrm{tr}(\bigwedge^{k}\mathcal{P}_{\gamma})\mathrm{sgn}\,(\det\mathcal{P}_{\gamma}|_{E_{s}})}{|\det(I-\mathcal{P}_{\gamma})|}\delta_{t-T_{\gamma}}. (9)

Let us follow [4, Section 3]. By [4, Lemma 3.2], we may and we will assume that near KK, (ϕt)t(\phi_{t})_{t\in\mathbb{R}} is an open hyperbolic system in the sense of [3, Assumptions (A1)–(A4)]. By [3, Lemma 1.17], there is C>0C>0 such that for all t0t\geq 0,

|{γ closed trajectory in K:Tγt}|CeCt.\displaystyle|\{\gamma\text{ closed trajectory in }K:T_{\gamma}\leq t\}|\leq Ce^{Ct}. (10)

For Im(λ)1{\rm Im}(\lambda)\gg 1 big enough, set

ζK,k(λ)=exp(γKTγTγtr(k𝒫γ)sgn(det𝒫γ|Es)|det(I𝒫γ)|eiλTγ).\displaystyle\zeta_{K,k}(\lambda)=\exp\left(-\sum_{\gamma\subset K}\frac{T_{\gamma}^{\sharp}}{T_{\gamma}}\frac{\mathrm{tr}(\bigwedge^{k}\mathcal{P}_{\gamma})\mathrm{sgn}\,(\det\mathcal{P}_{\gamma}|_{E_{s}})}{|\det(I-\mathcal{P}_{\gamma})|}e^{i\lambda T_{\gamma}}\right). (11)
Lemma 2.1.

For Im(λ)1{\rm Im}(\lambda)\gg 1 big enough, we have

λlogζK,k(λ)=i0eiλttrχϕtχ|C(M;0k~)dt.\displaystyle\partial_{\lambda}\log\zeta_{K,k}(\lambda)=-i\int_{0}^{\infty}e^{i\lambda t}\mathrm{tr}^{\flat}\left.\chi\phi_{-t}^{*}\chi\right|_{C^{\infty}\left(M;\widetilde{\mathcal{E}_{0}^{k}}\right)}dt. (12)

The function ζK,k(λ)\zeta_{K,k}(\lambda) has a holomorphic extension to \mathbb{C}.

Proof.

Let us first remark that by (9) and (10), the right hand side of (12) is well defined. Taking a logarithm and differentiating (11) and using Guillemin trace formula (9), we get (12). The last part of the lemma follows from the arguments of [3, Section 4]. ∎

Recall that for Im(λ)1{\rm Im}(\lambda)\gg 1 big enough, we have

ζK(λ)=γK(1eiλTλ)=exp(γKTγTγeiλTγ).\displaystyle\zeta_{K}(\lambda)=\prod_{\gamma^{\sharp}\subset K}\left(1-e^{i\lambda T_{\lambda}^{\sharp}}\right)=\exp\left(-\sum_{\gamma\subset K}\frac{T_{\gamma}^{\sharp}}{T_{\gamma}}e^{i\lambda T_{\gamma}}\right). (13)

Proposition 1.1 is a consequence of the following lemma. This lemma was stated in [2], but we restate and prove it for convenience.

Lemma 2.2.

The following identity of meromorphic functions on \mathbb{C} holds,

ζK(λ)=k=0n1(ζK,k(λ))(1)k+dimEs.\displaystyle\zeta_{K}(\lambda)=\prod_{k=0}^{n-1}\big{(}\zeta_{K,k}(\lambda)\big{)}^{(-1)^{k+\dim E_{s}}}. (14)
Proof.

Following [5, (2.4)-(2.5)], since det(I𝒫γ)=k=0n1(1)ktr(k𝒫γ)\det(I-\mathcal{P}_{\gamma})=\sum_{k=0}^{n-1}(-1)^{k}\mathrm{tr}\left(\bigwedge^{k}\mathcal{P}_{\gamma}\right), by (11) and (13), it is enough to show

|det(I𝒫γ)|=(1)dimEssgn(det𝒫γ|Es)det(I𝒫γ).\displaystyle|\det(I-\mathcal{P}_{\gamma})|=(-1)^{\dim E_{s}}\ \mathrm{sgn}\,\left(\det\left.\mathcal{P}_{\gamma}\right|_{E_{s}}\right)\det(I-\mathcal{P}_{\gamma}). (15)

Remark that

det(I𝒫γ)=det(I𝒫γ|Eu)det(I𝒫γ|Es)=(1)dimEsdet(I𝒫γ|Eu)det(I𝒫γ1|Es)det(𝒫γ|Es).\displaystyle\begin{split}\det(I-\mathcal{P}_{\gamma})&=\det(I-\mathcal{P}_{\gamma}|_{E_{u}})\det(I-\mathcal{P}_{\gamma}|_{E_{s}})\\ &=(-1)^{\dim E_{s}}\det(I-\mathcal{P}_{\gamma}|_{E_{u}})\det(I-\mathcal{P}_{\gamma}^{-1}|_{E_{s}})\det(\mathcal{P}_{\gamma}|_{E_{s}}).\end{split} (16)

As time is running in the negative direction, we have by (1) that the eigenvalues λ\lambda of 𝒫γ|Eu\mathcal{P}_{\gamma}|_{E_{u}} have |λ|<1|\lambda|<1, and the eigenvalues μ\mu of 𝒫γ1|Es\mathcal{P}_{\gamma}^{-1}|_{E_{s}} have |μ|<1|\mu|<1. This gives any eigenvalues of I𝒫γ|EuI-\mathcal{P}_{\gamma}|_{E_{u}} to be either 1λ1-\lambda for λ(1,1)\lambda\in(-1,1) or conjugate pairs 1λ,1λ¯1-\lambda,1-\overline{\lambda} when λ\lambda is not real. In any case, we get by multiplying that

det(I𝒫γ|Eu)>0\det(I-\mathcal{P}_{\gamma}|_{E_{u}})>0

and similarly

det(I𝒫γ1|Es)>0.\det(I-\mathcal{P}_{\gamma}^{-1}|_{E_{s}})>0.

Then taking signs in (16), we get (15). ∎

Remark.

The key point of our argument is based on the smoothness of 𝒪(Es)\mathscr{O}(E_{s}). Thanks to this property, most of the analytic arguments in the proof of Proposition 1.1 are reduced to [3]. In [2, Section 2], Baladi and Tsujii used the orientation bundle in a different way for the flow with discrete time.

3. Vanishing Order at Zero on a Contact 33-Manifold

In this section, we assume that MM is a connected closed 33-manifold with a contact form α\alpha, and that VV is the associated Reeb vector field. We suppose also that the flow (ϕt)t(\phi_{t})_{t\in\mathbb{R}} of VV is Anosov. One such example would be when M=SΣM=S^{*}\Sigma, the cosphere bundle of a connected closed surface Σ\Sigma with negative (variable) curvature, and (ϕt)t(\phi_{t})_{t\in\mathbb{R}} is geodesic flow.

The following result was proven in [6]:

Theorem 3.1.

If (ϕt)t(\phi_{t})_{t\in\mathbb{R}} is a contact Anosov flow on a connected closed 33-manifold with orientable EuE_{u} and EsE_{s}, the Ruelle zeta function has vanishing order at λ=0\lambda=0 equal to b1(M)2b_{1}(M)-2, where b1(M)b_{1}(M) denotes the first Betti number of MM.

The goal of this section is to determine the order of vanishing of ζR\zeta_{R} at 0 in the case that EsE_{s}, EuE_{u} are not orientable, and hence give a proof of Theorem 2. We remark that since a contact manifold is orientable, orientability of EsE_{s} is equivalent to orientability of EuE_{u}.

3.1. The twisted cohomology

Let us recall some background and general facts on the twisted cohomology of a flat vector bundle. Let XX be a closed manifold. Let FF be a flat vector bundle on XX with flat connection \nabla. It induces a sheaf \mathcal{F} on XX defined by locally constant sections, i.e., if UXU\subset X is an open set, then

(U)={sC(U;F|U):s=0}.\mathcal{F}(U)=\{s\in C^{\infty}(U;F|_{U}):\nabla s=0\}.

The twisted cohomology H(X;F)H^{\bullet}(X;F) is defined by the cohomology of the sheaf \mathcal{F} [10, Section II.4.4]. They are the algebraic invariants which describe the rigidity properties of the global flat sections of FF. Let bk(F)b_{k}(F) be the twisted Betti number

bk(F)=dimHk(X;F).b_{k}(F)=\dim H^{k}(X;F).

If FF is the trivial line bundle, we get the classical de Rham cohomology with real coefficients.

To evaluate H(X;F)H^{\bullet}(X;F), one can use the twisted de Rham complex. Indeed, if we denote Fk=Λk(TX)FF^{k}=\Lambda^{k}(T^{*}X)\otimes F, the flat connection \nabla extends to an operator dk:C(X;Fk)C(X;Fk+1)d_{k}:C^{\infty}(X;F^{k})\to C^{\infty}(X;F^{k+1}) by Leibniz rule: if αC(X;Λk(TX))\alpha\in C^{\infty}(X;\Lambda^{k}(T^{*}X)) and sC(X;F)s\in C^{\infty}(X;F), we have

dk(αs)=dαs+(1)kαs.d_{k}(\alpha\cdot s)=d\alpha\cdot s+(-1)^{k}\alpha\wedge\nabla s.

By the flatness of \nabla, we have dk+1dk=0d_{k+1}d_{k}=0, so that (C(X;F),d)(C^{\infty}(X;F^{\bullet}),d_{\bullet}) is a complex. By the de Rham isomorphism [10, Théorème II.4.7.1], we have

Hk(X;F)=kerdk/Imdk1.\displaystyle H^{k}(X;F)=\ker d_{k}/{\rm Im}d_{k-1}. (17)

As an analogue of [6, Lemma 2.1], using the theory of elliptic operators, we can evaluate H(X;F)H^{\bullet}(X;F) using the complex of twisted currents, or more generally twisted currents with wavefront conditions.

More precisely, let ΓTX\Gamma\subset T^{*}X be a closed cone. We denote by 𝒟Γ(X;Fk)\mathcal{D}^{\prime}_{\Gamma}(X;F^{k}) the space of FkF^{k}-valued distributions whose wavefront set is contained in Γ\Gamma (see [6, Section 2.1]). By microlocality, we have

dk:𝒟Γ(X;Fk)𝒟Γ(X;Fk+1).d_{k}:\mathcal{D}^{\prime}_{\Gamma}\left(X;F^{k}\right)\to\mathcal{D}^{\prime}_{\Gamma}\left(X;F^{k+1}\right).

For simplicity, we will write dd sometimes.

Lemma 3.2.

If u𝒟Γ(X;Fk)u\in\mathcal{D}^{\prime}_{\Gamma}\left(X;F^{k}\right) and duC(X;Fk+1)du\in C^{\infty}\left(X;F^{k+1}\right), then there exist vC(X;Fk)v\in C^{\infty}\left(X;F^{k}\right) and w𝒟Γ(X;Fk1)w\in\mathcal{D}^{\prime}_{\Gamma}\left(X;F^{k-1}\right) such that

u=v+dw.u=v+dw.

In particular, if u𝒟Γ(X;F)u\in\mathcal{D}^{\prime}_{\Gamma}\left(X;F\right) and duC(X;F1)du\in C^{\infty}\left(X;F^{1}\right), then uC(X;F).u\in C^{\infty}\left(X;F\right).

Proof.

Take a Riemannian metric on XX and a Hermitian metric on FF. Remark that these two metrics induce a fibrewise scalar product ,\langle\cdot,\cdot\rangle on FkF^{k}. For u,vC(X;Fk)u,v\in C^{\infty}\left(X;F^{k}\right), we can define the L2L^{2}-product by

u,vL2(X;Fk)=Xu,v𝑑vol,\displaystyle\langle u,v\rangle_{L^{2}\left(X;F^{k}\right)}=\int_{X}\langle u,v\rangle\,d\mathrm{vol}, (18)

where dvold\mathrm{vol} is a volume form. Let δk+1:C(X;Fk+1)C(X;Fk)\delta_{k+1}:C^{\infty}\left(X;F^{k+1}\right)\to C^{\infty}\left(X;F^{k}\right) be the formal adjoint of dd with respect to the L2L^{2}-product (18). Define the twisted Hodge Laplacian by

Δk=dk1δk+δk+1dk:C(X;Fk)C(X;Fk).\Delta_{k}=d_{k-1}\delta_{k}+\delta_{k+1}d_{k}:C^{\infty}\left(X;F^{k}\right)\to C^{\infty}\left(X;F^{k}\right).

Then Δk\Delta_{k} is an essentially self-adjoint second order elliptic differential operator. The remainder of the proof carries over identically from that of [6, Lemma 2.1]. ∎

Remark that if FF is the orientation bundle of certain vector bundle and uC(X;F)u\in C^{\infty}(X;F), then for xXx\in X, |u(x)|2|u(x)|^{2} is independent of the choice of trivializations. It defines a Hermitian metric on FF.

3.2. Resonant State Spaces.

Let MM be a connected 33-dimensional closed manifold with a contact form αC(M,TM)\alpha\in C^{\infty}(M,T^{*}M). Let VV be the associated Reeb vector field. Then,

iVα=1,ιVdα=0.\displaystyle i_{V}\alpha=1,\quad\iota_{V}d\alpha=0. (19)

We assume that the flow (ϕt)t(\phi_{t})_{t\in\mathbb{R}} associated to VV is Anosov. Let EuTME^{*}_{u}\subset T^{*}M be the dual of EsE_{s}. We will apply the results of Section 3.1 to the case where (X,F,Γ)=(M,𝒪(Es),Eu).(X,F,\Gamma)=(M,\mathcal{O}(E_{s}),E^{*}_{u}).

Since the flow (ϕt)t(\phi_{t})_{t\in\mathbb{R}} is Anosov, we have K=MK=M. For 0k20\leq k\leq 2, we write ζk=ζK,k\zeta_{k}=\zeta_{K,k}. By (14), we have

ζR(λ)=ζ1(λ)ζ0(λ)ζ2(λ).\displaystyle\zeta_{R}(\lambda)=\frac{\zeta_{1}(\lambda)}{\zeta_{0}(\lambda)\zeta_{2}(\lambda)}. (20)

We consider the operator Pk=iVP_{k}=-i\mathcal{L}_{V}, where V\mathcal{L}_{V} (in a slight abuse of notation) denotes the natural action on sections of 0k~\widetilde{\mathcal{E}_{0}^{k}}, given by the Lie derivative on sections of 0k\mathcal{E}_{0}^{k} tensored with the flat connection on 𝒪(Es)\mathscr{O}(E_{s}). For Imλ1\text{Im}\lambda\gg 1 large enough, the integral Rk(λ)=i0eiλtϕt𝑑tR_{k}(\lambda)=i\int_{0}^{\infty}e^{i\lambda t}\phi_{-t}^{*}dt converges and defines a bounded operator on the L2L^{2}-space; this is nothing more than the resolvent operator of PkP_{k}. Then by [6, Section 2.3] we have that RkR_{k} extends meromorphically to the entire complex plane,

Rk(λ):C(M;0k~)𝒟(M;0k~).R_{k}(\lambda):C^{\infty}\left(M;\widetilde{\mathcal{E}_{0}^{k}}\right)\to\mathcal{D}^{\prime}\left(M;\widetilde{\mathcal{E}_{0}^{k}}\right).

More precisely, near λ0\lambda_{0}\in\mathbb{C}, we have

Rk(λ)=Rk,H(λ)j=1J(λ0)(Pkλ)j1Πk(λλ0)jR_{k}(\lambda)=R_{k,H}(\lambda)-\sum_{j=1}^{J(\lambda_{0})}\frac{(P_{k}-\lambda)^{j-1}\Pi_{k}}{(\lambda-\lambda_{0})^{j}}

where Rk,HR_{k,H} is a holomorphic family defined near λ0\lambda_{0}, J(λ0)J(\lambda_{0})\in\mathbb{N}, and Πk\Pi_{k} has rank mk(λ0)<m_{k}(\lambda_{0})<\infty. By the arguments at the end of [5], we have that at λ0\lambda_{0}, the function ζk\zeta_{k} has a zero of order mk(λ0)m_{k}(\lambda_{0}).

We define the space of resonant states at λ0\lambda_{0} to be

Resk(λ0)={u𝒟Eu(M;0k~):(Pkλ0)u=0}.\mathrm{Res}_{k}(\lambda_{0})=\left\{u\in\mathcal{D}^{\prime}_{E^{*}_{u}}\left(M;\widetilde{\mathcal{E}_{0}^{k}}\right):(P_{k}-\lambda_{0})u=0\right\}.

Then a special case of [6, Lemma 2.2] gives the following:

Lemma 3.3.

Suppose PkP_{k} satisfies the semisimplicity condition:

u𝒟Eu(M;0k~),(Pkλ0)2u=0(Pkλ0)u=0.u\in\mathcal{D}^{\prime}_{E_{u}^{*}}\left(M;\widetilde{\mathcal{E}_{0}^{k}}\right),\quad(P_{k}-\lambda_{0})^{2}u=0\,\,\implies\,\,(P_{k}-\lambda_{0})u=0.

Then mk(λ0)=dimResk(λ0)m_{k}(\lambda_{0})=\dim\mathrm{Res}_{k}(\lambda_{0}).

Recall that we are trying to find the order at λ=0\lambda=0 of ζR\zeta_{R}, which by (20) is simply

mR(0)=m1(0)m0(0)m2(0).\displaystyle m_{R}(0)=m_{1}(0)-m_{0}(0)-m_{2}(0). (21)

We will compute each of these individually, by computing dimResk(0)\dim\mathrm{Res}_{k}(0) and checking that the semisimplicity condition in Lemma 3.3 holds.

We begin with twisted “0-forms’, which are just sections of the orientation bundle 𝒪(Es)\mathscr{O}(E_{s}).

Proposition 3.4.

If EsE_{s} is nonorientable, the space Res0(0)\mathrm{Res}_{0}(0) is {0}\{0\}.

Proof.

Suppose uRes0(0)u\in\mathrm{Res}_{0}(0), i.e.,

P0u=0.\displaystyle P_{0}u=0. (22)

Since the flow (ϕt)t(\phi_{t})_{t\in\mathbb{R}} preserves the contact volume form αdα\alpha\wedge d\alpha, P0:C(M;𝒪(Es))C(M;𝒪(Es))P_{0}:C^{\infty}(M;\mathscr{O}(E_{s}))\to C^{\infty}(M;\mathscr{O}(E_{s})) is a symmetric operator with respect to the L2L^{2}-product (18). By [6, Lemma 2.3], uC(M;𝒪(Es))u\in C^{\infty}\left(M;\mathscr{O}(E_{s})\right). Using t(ϕtu)=ϕtVu\partial_{t}\left(\phi_{-t}^{*}u\right)=-\phi_{-t}^{*}\nabla_{V}u (where \nabla is the flat connection), we see that uu is constant on the flow line: for all tt\in\mathbb{R},

u=ϕtu.\displaystyle u=\phi_{-t}^{*}u. (23)

Let (x,v)TM.(x,v)\in TM. The pairing du(x),v\langle du(x),v\rangle is an element of 𝒪(Es)x\mathscr{O}(E_{s})_{x}. By (7) and (23), we have

du(x),v=ϕt(du)(x),v=Φ~tdu(ϕt(x)),dϕt(x)v.\langle du(x),v\rangle=\langle\phi_{-t}^{*}(du)(x),v\rangle=\widetilde{\Phi}_{t}\langle du(\phi_{-t}(x)),d\phi_{-t}(x)v\rangle.

If vEu(x)v\in E_{u}(x), then sending tt\to\infty gives du(x),v=0\langle du(x),v\rangle=0 by (1). Similarly, if vEs(x)v\in E_{s}(x), then sending tt\to-\infty gives du(x),v=0\langle du(x),v\rangle=0. This shows that

du|EsEu=0.\displaystyle du|_{E_{s}\oplus E_{u}}=0. (24)

By Cartan’s formula and by (22), we have ιVdu=0\iota_{V}du=0, i.e.,

du|E0=0.\displaystyle du|_{E_{0}}=0. (25)

By (24) and (25), we have du=0du=0. So uH0(M;𝒪(Es))u\in H^{0}(M;\mathscr{O}(E_{s})). Since EsE_{s} is nonorientable, we have H0(M;𝒪(Es))=0H^{0}(M;\mathscr{O}(E_{s}))=0, so u=0u=0 and Res0(0)\mathrm{Res}_{0}(0) is trivial. ∎

Corollary 3.5.

If EsE_{s} is nonorientable, the multiplicity for 0-forms is m0(0)=0m_{0}(0)=0.

Proof.

If P02(u)=0P_{0}^{2}(u)=0, then P0uRes0(0)P_{0}u\in\mathrm{Res}_{0}(0). By Proposition 3.4, P0u=0P_{0}u=0, so uRes0(0)u\in\mathrm{Res}_{0}(0). This shows semisimplicity, so by Lemma 3.3 we see that m0(0)=dimRes0(0)=0m_{0}(0)=\dim\mathrm{Res}_{0}(0)=0. ∎

Proposition 3.6.

If EsE_{s} is nonorientable, the space Res2(0)\mathrm{Res}_{2}(0) is {0}\{0\}.

Proof.

We claim that

α:023\displaystyle\alpha\wedge:\mathcal{E}^{2}_{0}\to\mathcal{E}^{3} (26)

is a bundle isomorphism. Indeed, using (19), it is easy to see that the inverse of (26) is given by ιV\iota_{V}. Tensoring with 𝒪(Es)\mathscr{O}(E_{s}), we get a bundle isomorphism

α:02~3~.\displaystyle\alpha\wedge:\widetilde{\mathcal{E}^{2}_{0}}\to\widetilde{\mathcal{E}^{3}}. (27)

Let uRes2(0)u\in\mathrm{Res}_{2}(0). Since 3\mathcal{E}^{3} is generated by αdα\alpha\wedge d\alpha, by (27), there is v𝒟Eu(M;𝒪(Es))v\in\mathcal{D}^{\prime}_{E_{u}^{*}}(M;\mathscr{O}(E_{s})) such that αu=vαdα\alpha\wedge u=v\alpha\wedge d\alpha. Applying ιV\iota_{V} and using ιVu=0\iota_{V}u=0, we have u=vdαu=vd\alpha. Then

0=P2(u)=(P0v)dα.0=P_{2}(u)=(P_{0}v)d\alpha.

But this gives P0v=0P_{0}v=0, so by Proposition 3.4 we have v=0v=0. Therefore, u=0u=0. ∎

The following is then clear for the same reason as Corollary 3.5.

Corollary 3.7.

If EsE_{s} is nonorientable, the multiplicity for 22-forms is m2(0)=0m_{2}(0)=0.

We now turn to the case of P1P_{1} acting on the space of twisted 11-form-valued distributions 𝒟Eu(M;01~)\mathcal{D}^{\prime}_{E_{u}^{*}}\left(M;\widetilde{\mathcal{E}_{0}^{1}}\right). We can now state the analogous proposition for 11-forms:

Proposition 3.8.

If EsE_{s} is nonorientable, the space Res1(0)\mathrm{Res}_{1}(0) has dimension b1(𝒪(Es))b_{1}(\mathscr{O}(E_{s})).

Proof.

The proof is analogous to that of [6, Lemma 3.4], but slightly easier due to the holomorphy of the resolvent R0R_{0} near 0. Let uRes1(0)u\in\mathrm{Res}_{1}(0). Then duRes2(0)du\in\mathrm{Res}_{2}(0) by Proposition 3.6, so du=0du=0. By Lemma 3.2 there is a ϕ𝒟Eu(M;𝒪(Es))\phi\in\mathcal{D}^{\prime}_{E_{u}^{*}}(M;\mathscr{O}(E_{s})) such that

udϕC(M;1~),d(udϕ)=0.\displaystyle u-d\phi\in C^{\infty}\left(M;\widetilde{\mathcal{E}^{1}}\right),\quad d(u-d\phi)=0.

We shall show that the map:

Θ:u[udϕ]H1(M;𝒪(Es))\Theta:u\mapsto[u-d\phi]\in H^{1}(M;\mathscr{O}(E_{s}))

is well-defined, linear and bijective, which is enough to prove the lemma.

Well-Definedness and linearity:

Suppose there is another section ψ𝒟Eu(M;𝒪(Es))\psi\in\mathcal{D}^{\prime}_{E_{u}^{*}}\left(M;\mathscr{O}(E_{s})\right) with udψC(M;1~)u-d\psi\in C^{\infty}\left(M;\widetilde{\mathcal{E}^{1}}\right). Then subtracting gives d(ϕψ)C(M;1~)d(\phi-\psi)\in C^{\infty}\left(M;\widetilde{\mathcal{E}^{1}}\right), so ϕψC(M;𝒪(Es))\phi-\psi\in C^{\infty}\left(M;\mathscr{O}(E_{s})\right) by Lemma 3.2. This shows that the map Θ\Theta is well-defined. It is also easy to see that Θ\Theta is linear.

Injectivity:

If Θ(u)=0\Theta(u)=0, then udϕu-d\phi is exact, so without loss of generality we can assume that u=dϕu=d\phi. Combining with ιVu=0\iota_{V}u=0, we get ϕRes0(0)\phi\in\mathrm{Res}_{0}(0), so ϕ=0\phi=0 by Proposition 3.4. Therefore u=0u=0, and this shows Θ\Theta to be injective.

Surjectivity:

Let vC(M;1~)v\in C^{\infty}\left(M;\widetilde{\mathcal{E}^{1}}\right) with dv=0dv=0. Then as m0(0)=0m_{0}(0)=0, the resolvent R0R_{0} is holomorphic near 0. Take ϕ=iR0(0)ιVv𝒟Eu(M;𝒪(Es))\phi=iR_{0}(0)\iota_{V}v\in\mathcal{D}^{\prime}_{E_{u}^{*}}\left(M;\mathscr{O}(E_{s})\right). Then P0ϕ=iιVvP_{0}\phi=i\iota_{V}v. This rearranges to ιV(v+dϕ)=0\iota_{V}(v+d\phi)=0, so v+dϕRes1(0)v+d\phi\in\mathrm{Res}_{1}(0). This gives that Θ\Theta is surjective, and completes the proof of our proposition. ∎

Proposition 3.9.

If EsE_{s} is nonorientable, the multiplicity for 11-forms is m1(0)=b1(𝒪(Es))m_{1}(0)=b_{1}(\mathscr{O}(E_{s})).

Proof.

By Lemma 3.3, we must only check that the semisimplicity condition is satisfied. Take u𝒟Eu(M;01~)u\in\mathcal{D}^{\prime}_{E_{u}^{*}}\left(M;\widetilde{\mathcal{E}_{0}^{1}}\right) such that (P1)2u=0(P_{1})^{2}u=0. Then v=ιVduRes1(0)v=\iota_{V}du\in\mathrm{Res}_{1}(0). It is enough to show that v=0v=0.

Recall that in the proof of Proposition 3.8, we have seen that elements in Res1(0){\rm Res}_{1}(0) are closed. In particular,

dv=0.\displaystyle dv=0. (28)

Note that αdu𝒟Eu(M;3~)\alpha\wedge du\in\mathcal{D}^{\prime}_{E_{u}^{*}}(M;\widetilde{\mathcal{E}^{3}}). We claim that

αdu=0.\displaystyle\alpha\wedge du=0. (29)

Indeed, there is some a𝒟Eu(M;𝒪(Es))a\in\mathcal{D}^{\prime}_{E_{u}^{*}}(M;\mathscr{O}(E_{s})) such that

αdu=aαdα.\displaystyle\alpha\wedge du=a\,\alpha\wedge d\alpha.

Since V(α)=0\mathcal{L}_{V}(\alpha)=0, by (28), we have

(Va)αdα=αV(du)=αdιVdu=αdv=0.\displaystyle({\mathcal{L}}_{V}a)\alpha\wedge d\alpha=\alpha\wedge{\mathcal{L}}_{V}(du)=\alpha\wedge d\iota_{V}du=\alpha\wedge dv=0.

Then Va=0\mathcal{L}_{V}a=0, so a=0a=0 by Proposition 3.4. This gives (29).

Since α(V)=1\alpha(V)=1, we have (α)ιV+ιV(α)=id(\alpha\wedge)\circ\iota_{V}+\iota_{V}\circ(\alpha\wedge)=\mathrm{id}. By (29), we have

du=((α)ιV+ιV(α))du=αv.\displaystyle du=((\alpha\wedge)\circ\iota_{V}+\iota_{V}\circ(\alpha\wedge))du=\alpha\wedge v. (30)

By Lemma 3.2 and by (28), there are wC(M;1~)w\in C^{\infty}\left(M;\widetilde{\mathcal{E}^{1}}\right), ϕ𝒟Eu(M;𝒪(Es))\phi\in\mathcal{D}^{\prime}_{E_{u}^{*}}\left(M;\mathscr{O}(E_{s})\right) such that

v=w+dϕ,dw=0.\displaystyle v=w+d\phi,\quad dw=0. (31)

Then

ιVw=ιV(vdϕ)=Vϕ.\displaystyle\iota_{V}w=\iota_{V}(v-d\phi)=-\mathcal{L}_{V}\phi. (32)

In particular, Vϕ\mathcal{L}_{V}\phi is smooth. We compute by Stokes’ Theorem and by (30)-(32),

0\displaystyle 0 =ReM𝑑uw¯=ReMαdϕw¯=ReMϕw¯dα\displaystyle=\mathrm{Re\,}\int_{M}du\wedge\overline{w}=\mathrm{Re\,}\int_{M}\alpha\wedge d\phi\wedge\overline{w}=\mathrm{Re\,}\int_{M}\phi\overline{w}\wedge d\alpha
=ReMιV(ϕw¯)αdα=ReMϕ(Vϕ¯)αdα=ReVϕ,ϕL2(M;𝒪(Es)),\displaystyle=\mathrm{Re\,}\int_{M}\iota_{V}(\phi\overline{w})\,\alpha\wedge d\alpha=-\mathrm{Re\,}\int_{M}\phi(\overline{\mathcal{L}_{V}\phi})\alpha\wedge d\alpha=-\mathrm{Re\,}\langle\mathcal{L}_{V}\phi,\phi\rangle_{L^{2}\left(M;\mathscr{O}(E_{s})\right)},

where the fourth equality comes from the fact the

(α)ιV(ϕw¯dα)=((α)ιV+ιV(α))(ϕw¯dα)=ϕw¯dα.(\alpha\wedge)\circ\iota_{V}(\phi\overline{w}\wedge d\alpha)=((\alpha\wedge)\circ\iota_{V}+\iota_{V}\circ(\alpha\wedge))(\phi\overline{w}\wedge d\alpha)=\phi\overline{w}\wedge d\alpha.

In the above formula, we use the fact that a product of two twisted forms is untwisted. By [6, Lemma 2.3] we have ϕC(M;𝒪(Es))\phi\in C^{\infty}\left(M;\mathscr{O}(E_{s})\right), so vC(M;01~)v\in C^{\infty}\left(M;\widetilde{\mathcal{E}_{0}^{1}}\right). Then by the same argument as in Proposition 3.4 (see [6, Lemma 3.5]) we have v=0v=0. ∎

Now Theorem 2 is a consequence of (21), Corollaries 3.5, 3.7, and Proposition 3.9.

Let Σ\Sigma be a connected negatively curved closed surface. Take M=SΣM=S^{*}\Sigma. By Corollary 1.3, we have

H1(M;𝒪(Es))=H1(M;π𝒪(TΣ)).\displaystyle H^{1}(M;\mathscr{O}(E_{s}))=H^{1}(M;\pi^{*}\mathscr{O}(T\Sigma)).
Proposition 3.10.

If Σ\Sigma is a connected negatively curved closed surface (oriented or not), we have

dimH1(M;π𝒪(TΣ))=dimH1(Σ).\displaystyle\dim H^{1}(M;\pi^{*}\mathscr{O}(T\Sigma))=\dim H^{1}(\Sigma). (33)
Proof.

By the Gysin long exact sequence, we have the exact sequence

where π\pi^{*} is the pullback, π\pi_{*} is the integration along the fibre of MΣM\to\Sigma, and eH2(Σ;𝒪(TΣ))e\in H^{2}(\Sigma;\mathscr{O}(T\Sigma)) is the Euler class of TΣT\Sigma.

We claim that the last map

e:H0(Σ)H2(Σ;𝒪(TΣ))\displaystyle e\wedge:H^{0}(\Sigma)\to H^{2}(\Sigma;\mathscr{O}(T\Sigma))

in the Gysin exact sequence is an isomorphism. Indeed, since Σ\Sigma is connected, we have dimH0(Σ)=1\dim H^{0}(\Sigma)=1, and by Poincaré duality, dimH2(Σ;𝒪(TΣ))=1\dim H^{2}(\Sigma;\mathscr{O}(T\Sigma))=1. It is enough to show that eH2(Σ;𝒪(TΣ))e\in H^{2}(\Sigma;\mathscr{O}(T\Sigma)) is non zero, or equivalently Σe0\int_{\Sigma}e\neq 0. This is a consequence of the fact that Σ\Sigma has negative curvature, as e=Kμe=K\mu where μ\mu is the Riemannian density and K<0K<0 is the Gauss curvature.

Therefore, we get an isomorphism

π:H1(Σ;𝒪(TΣ))H1(M;π𝒪(TΣ)).\displaystyle\pi^{*}:H^{1}(\Sigma;\mathscr{O}(T\Sigma))\simeq H^{1}(M;\pi^{*}\mathscr{O}(T\Sigma)). (34)

By Poincaré duality, we have

H1(Σ;𝒪(TΣ))(H1(Σ)).\displaystyle H^{1}(\Sigma;\mathscr{O}(T\Sigma))\simeq\left(H^{1}(\Sigma)\right)^{*}. (35)

By (34) and (35), we get (33). ∎

Now Corollary 3 is a consequence of Theorem 2 and Proposition 3.10.

4. Acknowledgements

We would like to thank Maciej Zworski for suggesting the problem and giving guidance along the way, and Kiran Luecke for some helpful commments. We are indebted to the referees for reading the paper very carefully and providing us the reference [12]. Y. B-W also acknowledges the partial support under the NSF grant DMS-1952939. S.S. acknowledges the partial support under the ANR grant ANR-20-CE40-0017.

References

  • [1] Anosov, D. V. Geodesic flows on closed Riemannian manifolds of negative curvature. Trudy Mat. Inst. Steklov. 90 (1967), 209.
  • [2] Baladi, V., and Tsujii, M. Dynamical determinants and spectrum for hyperbolic diffeomorphisms. Contemp. Math 469 (2008), 29–68.
  • [3] Dyatlov, S., and Guillarmou, C. Pollicott-Ruelle resonances for open systems. Ann. Henri Poincaré 17, 11 (2016), 3089–3146.
  • [4] Dyatlov, S., and Guillarmou, C. Afterword: Dynamical zeta functions for axiom a flows. Bulletin of the American Mathematical Society 55, 3 (2018), 337–342.
  • [5] Dyatlov, S., and Zworski, M. Dynamical zeta functions for Anosov flows via microlocal analysis. Ann. Sci. Éc. Norm. Supér. (4) 49, 3 (2016), 543–577.
  • [6] Dyatlov, S., and Zworski, M. Ruelle zeta function at zero for surfaces. Inventiones mathematicae 210, 1 (2017), 211–229.
  • [7] Eberlein, P. Geodesic flows in manifolds of nonpositive curvature. In Smooth ergodic theory and its applications (Seattle, WA, 1999), vol. 69 of Proc. Sympos. Pure Math. Amer. Math. Soc., Providence, RI, 2001, pp. 525–571.
  • [8] Fried, D. Meromorphic zeta functions for analytic flows. Communications in mathematical physics 174, 1 (1995), 161–190.
  • [9] Giulietti, P., Liverani, C., and Pollicott, M. Anosov flows and dynamical zeta functions. Annals of Mathematics (2013), 687–773.
  • [10] Godement, R. Topologie algébrique et théorie des faisceaux. Hermann, Paris, 1973.
  • [11] Guillemin, V., et al. Lectures on spectral theory of elliptic operators. Duke Mathematical Journal 44, 3 (1977), 485–517.
  • [12] Klingenberg, W. Riemannian manifolds with geodesic flow of Anosov type. Ann. of Math. (2) 99 (1974), 1–13.
  • [13] Smale, S. Differentiable dynamical systems. Bulletin of the American mathematical Society 73, 6 (1967), 747–817.