Dynamical Zeta Functions in the Nonorientable Case
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
where the product is taken over all primitive closed geodesics with corresponding periods . Note that by [3, Lemma 1.17] and [4, Section 3], this product converges for large enough. The meromorphic continuation of to all of 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 is an Axiom A flow on a closed manifold, the Ruelle zeta function extends to a meromorphic function on .
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 -manifold, and study the order of vanishing of at . An important example is when , the cosphere bundle of a connected negatively curved closed surface , and 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 , where is the first Betti number of .
We shall show that for nonorientable stable bundle, the analogous result is the following:
Theorem 2.
Let be the Reeb flow on a connected contact closed -manifold. If is Anosov with nonorientable stable bundle , the Ruelle zeta function has vanishing order at equal to , the dimension of the first de Rham cohomology with coefficients in the orientation line bundle of .
The orientation line bundle is reviewed in Definition 1.5.
In the special case of the geodesic flow on with nonorientable, the vanishing order at is given by , as is shown in Proposition 3.10. This is in contrast to the orientable case, in which it is .
More precisely, let be the derived Euler characteristic of , i.e.,
Corollary 3.
If 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 equal to .
1.1. Axiom A Flows
Let be a compact manifold without boundary of dimension , and let be a flow on generated by the vector field .
Definition 1.1.
A -invariant set is called hyperbolic for the flow if does not vanish on and for each the tangent space can be written as the direct sum
where , are continuous -invariant vector bundles on , and for some Riemannian metric , there are such that for all ,
(1) |
In the important case where all of is hyperbolic, we call an Anosov flow.
There is an analogous notion of hyperbolicity at fixed points.
Definition 1.2.
A fixed point , i.e., , is called hyperbolic if the differential 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 is called Axiom A if
-
(1)
all fixed points of are hyperbolic,
-
(2)
the closure of the union of all closed orbits of is hyperbolic,
-
(3)
the nonwandering set ([4, Definition 2.2]) of is the disjoint union of the set of fixed points and .
We now recall the definition of a locally maximal set, given in [4, Definition 2.4].
Definition 1.4.
A compact -invariant set is called locally maximal for if there is a neighborhood of such that
We may then state the key proposition, which generalises [4, Proposition 3.1] to the case where or is not necessarily orientable on .
Proposition 1.1.
Let be a locally maximal hyperbolic set for , and let be defined as the Ruelle zeta function where we only take the product over trajectories in . Then has a continuation to a meromorphic function on all of .
Theorem 1 follows from Proposition 1.1, as we may remark that by [13, II.5 Theorem 5.2] we can write with basic hyperbolic111These are locally maximal hyperbolic by definition (see [4, Definition 2.5]).. Then the product
which a priori holds for , gives that also has a meromorphic continuation to all of .
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 of rank over a manifold with projection map , let be two small open sets with nonempty intersection, and let , be local trivializations. Then the map is of the form
where is called a transition function. If the local trivializations can be chosen such that are smooth, then is a smooth vector bundle. Similarly, if can be chosen to be locally constant functions, then is a flat vector bundle.
Furthermore, suppose we are given an open cover of together with a set of continuous (resp. smooth, resp. locally constant) -valued functions with on . Then there exists a continuous (resp. smooth, resp. flat) vector bundle with transition functions , provided the following triple product property holds:
for any
Definition 1.5.
If is a continuous (but not necessarily smooth) real vector bundle over with transition functions , the orientation bundle of is a smooth flat line bundle with transition functions
Recall that if is a map, we say lifts to a bundle map if .
Since is a flat vector bundle, using the associated flat connection, we can lift the flow to a flow on . If the flow on lifts to a flow on , if are distinct trivializations of near , respectively, and are trivializations of near , respectively, we have for and :
(2) |
where is the obvious projection to the second component.
1.3. Geodesic flows
Let be a negatively curved closed Riemannian manifold. Let be the cosphere bundle on . It is classical that the geodesic flow on is Anosov [1].
Let be the canonical projection. For , we have a morphism of linear spaces
(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 induces an isomorphism of continuous vector bundles on ,
(4) |
Proof.
Since both sides of (4) have the same dimension, it is enough to show that is injective. We will show this using Jacobi fields. It is convenient to work on the sphere bundle . We identify with via the Riemannian metric on .
We follow [7, Section II.H]. Let be the total space of . Denote still by the obvious projection. Let be the vertical subbundle of . The Levi-Civita connection on induces a horizontal subbundle of , so that
(5) |
Since and , by (5), we can identify the smooth vector bundles,
For , let be the unique geodesic on such that For , let be the unique Jacobi field along such that where is the covariant derivation of in the direction . Recall that a Jacobi field is called stable, if there is such that for all ,
By [7, Proposition VI.A], given , for any , there exists one and only one stable Jacobi field along such that .
Since is a trivial line bundle, our proposition implies immediately:
Corollary 1.3.
We have the isomorphism of smooth flat line bundles
2. Proof of Proposition 1.1
We use the notation of [5]. If , let denote the subbundle of -forms such that , where denotes interior multiplication.
Let We consider the pullback on sections of . Note that the flow lifts to a flow on . Indeed, for , , is defined for by
(6) |
Note that from the above formula, it is easy to check that . Recall also that the flow lifts to a flow on (see (2)). For a section in , we have
(7) |
Let be a smooth function whose support is contained in a small neighborhood of such that for all . We now invoke the Guillemin trace formula (see [11, pp. 501-502], [5, Appendix B], [3, (4.6)]) which says that the flat trace is a distribution on given by
(8) |
where the sum is taken over all the periodic trajectories in with period and primitive period , is any point on , and is the linearized Poincaré map at . Note that as trace and determinant are invariant under conjugation, the right hand side does not depend on .
By (6), the trace of on is just . By (2), we may take trivializations of , in a neighborhood of and have the induced lifting on to be By definition we get this to be equal to
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
(9) |
Let us follow [4, Section 3]. By [4, Lemma 3.2], we may and we will assume that near , is an open hyperbolic system in the sense of [3, Assumptions (A1)–(A4)]. By [3, Lemma 1.17], there is such that for all ,
(10) |
For big enough, set
(11) |
Lemma 2.1.
For big enough, we have
(12) |
The function has a holomorphic extension to .
Proof.
Recall that for big enough, we have
(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 holds,
(14) |
Proof.
Following [5, (2.4)-(2.5)], since , by (11) and (13), it is enough to show
(15) |
Remark that
(16) |
As time is running in the negative direction, we have by (1) that the eigenvalues of have , and the eigenvalues of have . This gives any eigenvalues of to be either for or conjugate pairs when is not real. In any case, we get by multiplying that
and similarly
3. Vanishing Order at Zero on a Contact -Manifold
In this section, we assume that is a connected closed -manifold with a contact form , and that is the associated Reeb vector field. We suppose also that the flow of is Anosov. One such example would be when , the cosphere bundle of a connected closed surface with negative (variable) curvature, and is geodesic flow.
The following result was proven in [6]:
Theorem 3.1.
If is a contact Anosov flow on a connected closed -manifold with orientable and , the Ruelle zeta function has vanishing order at equal to , where denotes the first Betti number of .
The goal of this section is to determine the order of vanishing of at in the case that , are not orientable, and hence give a proof of Theorem 2. We remark that since a contact manifold is orientable, orientability of is equivalent to orientability of .
3.1. The twisted cohomology
Let us recall some background and general facts on the twisted cohomology of a flat vector bundle. Let be a closed manifold. Let be a flat vector bundle on with flat connection . It induces a sheaf on defined by locally constant sections, i.e., if is an open set, then
The twisted cohomology is defined by the cohomology of the sheaf [10, Section II.4.4]. They are the algebraic invariants which describe the rigidity properties of the global flat sections of . Let be the twisted Betti number
If is the trivial line bundle, we get the classical de Rham cohomology with real coefficients.
To evaluate , one can use the twisted de Rham complex. Indeed, if we denote , the flat connection extends to an operator by Leibniz rule: if and , we have
By the flatness of , we have , so that is a complex. By the de Rham isomorphism [10, Théorème II.4.7.1], we have
(17) |
As an analogue of [6, Lemma 2.1], using the theory of elliptic operators, we can evaluate using the complex of twisted currents, or more generally twisted currents with wavefront conditions.
More precisely, let be a closed cone. We denote by the space of -valued distributions whose wavefront set is contained in (see [6, Section 2.1]). By microlocality, we have
For simplicity, we will write sometimes.
Lemma 3.2.
If and , then there exist and such that
In particular, if and , then
Proof.
Take a Riemannian metric on and a Hermitian metric on . Remark that these two metrics induce a fibrewise scalar product on . For , we can define the -product by
(18) |
where is a volume form. Let be the formal adjoint of with respect to the -product (18). Define the twisted Hodge Laplacian by
Then 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 is the orientation bundle of certain vector bundle and , then for , is independent of the choice of trivializations. It defines a Hermitian metric on .
3.2. Resonant State Spaces.
Let be a connected -dimensional closed manifold with a contact form . Let be the associated Reeb vector field. Then,
(19) |
We assume that the flow associated to is Anosov. Let be the dual of . We will apply the results of Section 3.1 to the case where
Since the flow is Anosov, we have . For , we write . By (14), we have
(20) |
We consider the operator , where (in a slight abuse of notation) denotes the natural action on sections of , given by the Lie derivative on sections of tensored with the flat connection on . For large enough, the integral converges and defines a bounded operator on the -space; this is nothing more than the resolvent operator of . Then by [6, Section 2.3] we have that extends meromorphically to the entire complex plane,
More precisely, near , we have
where is a holomorphic family defined near , , and has rank . By the arguments at the end of [5], we have that at , the function has a zero of order .
We define the space of resonant states at to be
Then a special case of [6, Lemma 2.2] gives the following:
Lemma 3.3.
Suppose satisfies the semisimplicity condition:
Then .
Recall that we are trying to find the order at of , which by (20) is simply
(21) |
We will compute each of these individually, by computing and checking that the semisimplicity condition in Lemma 3.3 holds.
We begin with twisted “-forms’, which are just sections of the orientation bundle .
Proposition 3.4.
If is nonorientable, the space is .
Proof.
Suppose , i.e.,
(22) |
Since the flow preserves the contact volume form , is a symmetric operator with respect to the -product (18). By [6, Lemma 2.3], . Using (where is the flat connection), we see that is constant on the flow line: for all ,
(23) |
Let The pairing is an element of . By (7) and (23), we have
If , then sending gives by (1). Similarly, if , then sending gives . This shows that
(24) |
Corollary 3.5.
If is nonorientable, the multiplicity for -forms is .
Proof.
Proposition 3.6.
If is nonorientable, the space is .
Proof.
The following is then clear for the same reason as Corollary 3.5.
Corollary 3.7.
If is nonorientable, the multiplicity for -forms is .
We now turn to the case of acting on the space of twisted -form-valued distributions . We can now state the analogous proposition for -forms:
Proposition 3.8.
If is nonorientable, the space has dimension .
Proof.
The proof is analogous to that of [6, Lemma 3.4], but slightly easier due to the holomorphy of the resolvent near . Let . Then by Proposition 3.6, so . By Lemma 3.2 there is a such that
We shall show that the map:
is well-defined, linear and bijective, which is enough to prove the lemma.
Well-Definedness and linearity:
Suppose there is another section with . Then subtracting gives , so by Lemma 3.2. This shows that the map is well-defined. It is also easy to see that is linear.
Injectivity:
If , then is exact, so without loss of generality we can assume that . Combining with , we get , so by Proposition 3.4. Therefore , and this shows to be injective.
Surjectivity:
Let with . Then as , the resolvent is holomorphic near . Take . Then . This rearranges to , so . This gives that is surjective, and completes the proof of our proposition. ∎
Proposition 3.9.
If is nonorientable, the multiplicity for -forms is .
Proof.
By Lemma 3.3, we must only check that the semisimplicity condition is satisfied. Take such that . Then . It is enough to show that .
Recall that in the proof of Proposition 3.8, we have seen that elements in are closed. In particular,
(28) |
Since , we have . By (29), we have
(30) |
By Lemma 3.2 and by (28), there are , such that
(31) |
Then
(32) |
In particular, is smooth. We compute by Stokes’ Theorem and by (30)-(32),
where the fourth equality comes from the fact the
In the above formula, we use the fact that a product of two twisted forms is untwisted. By [6, Lemma 2.3] we have , so . Then by the same argument as in Proposition 3.4 (see [6, Lemma 3.5]) we have . ∎
Let be a connected negatively curved closed surface. Take . By Corollary 1.3, we have
Proposition 3.10.
If is a connected negatively curved closed surface (oriented or not), we have
(33) |
Proof.
By the Gysin long exact sequence, we have the exact sequence
where is the pullback, is the integration along the fibre of , and is the Euler class of .
We claim that the last map
in the Gysin exact sequence is an isomorphism. Indeed, since is connected, we have , and by Poincaré duality, . It is enough to show that is non zero, or equivalently . This is a consequence of the fact that has negative curvature, as where is the Riemannian density and is the Gauss curvature.
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.