On the Ruelle-Mayer Transfer Operators for Hölder Continuous Functions
Abstract
We consider a family of operators connected with the geodesic flow on the modular surface. We show certain spectral information is retained after expanding their domain to the space of -Hölder continuous functions on the unit interval. For example, the point spectra associated with the Maass cusp forms and non-trivial zeroes of the Riemann zeta function to the right of the critical line remain unchanged when the Hölder constant is and respectively. We briefly consider a three-term functional equation introduced by Lewis in the Hölder setting and provide a partial classification of solutions in this setting.
1 Introduction
Let be the Gauss map
| (1) |
We associate to the following family of transfer operators which we formally define as acting on some space of functions :
| (2) |
where . When , this is the familiar Ruelle-Perron-Frobenius operator for and can formally be interpreted as the right-adjoint of the Koopman operator . The operators in the case were first introduced by Mayer [15], who considered them acting on the Banach space of functions which are holomorphic on with continuous extension to .
For this choice of Banach space, Mayer also showed that the operator-valued function admits a meromorphic continuation on , with simple poles at for . The continuation for can be obtained by rewriting with and noting that
| (3) |
where is the Hurwitz zeta function, which has an analytic continuation for all and a simple pole of residue at . In fact, it can further be analytically continued to for any by noticing that
| (4) |
where . Furthermore is a nuclear operator on of order for all .
The Mayer transfer operator encodes geometric information about the modular surface, which is the quotient orbifold obtained from considering the Poincaré upper half plane equipped with the usual Riemannian metric and identifying points via equivalence relation for all with . Since is a nuclear operator, the Fredholm determinants and are well-defined. These relate to Selberg Zeta function by the formula
| (5) |
see [16]. The Selberg zeta function is closely connected with the spectral theory of the Laplace-Beltrami operator [11]. Equation (3) is thus a way of connecting the spectral theory of with that of . This connection has been elaborated upon by Lewis, Zagier, Mayer and Chang [14, 13, 4, 3].
From a dynamical perspective, it is interesting to see to what extent the aforementioned spectral properties of survive when we consider e.g. Hölder continuous functions instead. For , let denote the space of -Hölder functions on . If with and , let be the set of functions for which the -th derivative exists and is continuous or -Hölder if or respectively. Denote by the operator (4) acting on , which we shall show to be well-defined and continuous for large enough . These operators are not compact, but we can decompose the spectrum into the essential spectrum and the discrete spectrum which consists of isolated eigenvalues of finite algebraic multiplicity. We prove the following.
Theorem 1.1.
On the half-plane , the radius of the essential spectrum satisfies
Furthermore, the generalised eigenspaces belonging to eigenvalues with consist of analytic functions which extend to generalised eigenfunctions of .
The statement in the abstract concerning Maass cusp forms and nontrivial zeroes of the Riemann zeta function, which we make more precise in Remark 3.2, turns out to follow immediately from this theorem.
Comparing the behaviour of transfer operators for different levels of regularity is a somewhat common theme in dynamics. One typically expects for a transfer operator associated to a sufficiently ‘nice’ expanding map that the essential spectral radius vanishes for the operator acting on spaces of increasingly smooth functions, see e.g. [2]. In a sense the above theorem is therefore not too suprising. Indeed, we shall show that if is a natural number, the proof of Theorem 1.1 is a mundane application of standard techniques.
However, the proof of the case is perhaps more interesting. In that case, we still obtain uniformly bounded estimates for on vertical lines even when , where the function is unbounded. However, by (2), the unboundedness in is due to the contribution rank one operator , which does not contribute to the essential spectral radius. Nevertheless, we do not obtain this result using a Doeblin-Fortet-Lasota-Yorke inequality, so our proof is substantially different from the standard approaches to these type of problems.
In Section 4, we interpret this result in terms of Lewis’ three-term functional equation. In particular, we can interpret this as an extension of the ‘bootstrapping’ result of Lewis and Zagier, who proved in [13] that real-analytic solutions of this equation on satisfying certain growth conditions automatically extend to holomorphic functions on the cut plane . We show that for , we may even replace real-analyticity with a Hölder condition and obtain that it is still the restriction of a holomorphic function on .
2 Preliminaries
2.1 On the Gauss Map
We briefly elucidate the Markov Structure of the Gauss map and define some notation which will be useful later on. There is a natural countable partition of the unit interval into intervals . Each of these intervals are mapped by homeomorphically onto . We denote by the inverse branch . Given a sequence we denote by the map defined by
| (6) |
We also denote the right-hand side by as a simplifying bit of notation and we let denote the set .
2.2 The Essential Spectral Radius
In Section 3, we provide estimates for the essential spectrum of , so we provide a brief reminder of some definitions. Let be the identity operator. We recall that the spectrum of an operator on a Banach space is the set of all for which is not invertible. Following Browder [1], we define the essential spectrum to be the set of all for which is not closed, is infinite-dimensional or for which is a limit point of . A nice consequence of this definition is that consists solely of isolated eigenvalues with finite-dimensional generalised eigenspaces.
There is a convenient formula for the essential spectral radius due to Nussbaum [19] that we now briefly introduce. Let be the subspace of compact operators on . Define the seminorm on the space of bounded linear operators on by
and let be a bounded linear operator. Then its essential spectral radius satisfies
| (7) |
2.3 Definition of Transfer Operators
In this subsection, we show that the Mayer transfer operators with are well-defined on Hölder-continuous functions with a sufficiently large Hölder exponent. We also provide a brief overview of some definitions and notations which will be used throughout this preprint.
Denote for by the space of -Hölder continuous functions on equipped with the norm
and . It will turn out to be useful to define for the auxiliary function on by
We remark that .
For , the operator (2) is also well defined on the space of bounded functions equipped with the -norm. We denote the operator acting on this space by .
Lemma 2.1.
Let , with . Then the operator is well-defined and continuous on . Furthermore there exists a constant and a family of constants depending only on and such that
Proof.
We note that for we have that as , hence we have that
| (8) |
To bound the auxiliary function, we start with a basic application of Taylor’s theorem. Let be a complex number with . Let . Then
| (9) |
for a constant which can be uniformly bounded over all and .
Note that
By first using the definition of the auxiliary function, and then applying (9) for we find that there exists some constant which is independent of for which
By estimating the second term using (9) for , we obtain that
| (10) |
for some uniformly bounded constant . Let . If , we sum over all in (10) to obtain
where denotes the function . Else let be large enough such that . Sum both sides of the inequality
for and add the respective sides in (10) over all . After taking the supremum over all we then obtain that for some . The lemma follows from boundedness of the operator on
∎
Remark 2.1.
To apply the usual method for obtaining quasi-compactness with essential spectral radius less than , we would need an estimate of the form
where and . It is indeed possible to find an upper bound for by iteratively using Lemma 2.1. Unfortunately, this does not yield a quantitatively good upper bound, both due to the dependence on in the estimate (8) and when considering for which is large.
2.4 Operator-Valued Holomorphic Functions
On occasion, we shall make use of the notion of holomorphy of functions, so we briefly recall the definition. We follow Chapter III, Section 1 and Section 3 in [12]. Let be a simply connected open set of . If is a Banach space with norm , or more generally a topological vector space, and is some -valued function, we define to be homeomorphic if has a limit as for .
If are Banach spaces, let be the set of bounded operators equipped with the usual operator norm, then a map is defined to be holomorphic if it is holomorphic according to the Banach space definition outlined above.
Many of the usual theorems in complex analysis continue to hold for Banach or topological vector space-valued functions. In particular, holomorphic implies infinitely differentiable with a locally converging Taylor expansion. The unique extension theorem also continues to hold. Finally, note that it is straightforward to show that is a meromorphic function on for any with derivative
3 Uniformly Bounded Essential Spectrum
By (7), we can estimate the essential spectral radius by finding for each a compact operator for which is small. A first naive approach would be to consider the operator which sends to a piecwise linear approximation that interpolates on the points and considering as . For , we indeed have by Lemma 2.1 that for any there exists some for which . However, we do not have sufficient control over the -norm. We remedy this by considering a countable interpolation which ensures remains a (nonlinear) interpolation of .
To this end, let be the set of points . For , we let
i.e. is the set of all points of the form and is the union of the set of all such points for with .
We remark that the set is closed. Indeed, the set of accumulation points are the set of all fractions that can be written as for and these are in by definition. Hence its complement is a countable disjoint collection of open intervals and we can define the following.
Definition 3.1.
For , we define the operator by letting for . If , let and the respective largest and smallest element of such that and interpolate the function between the two points to obtain
We prove it is well-defined in Lemma A.2 in the appendix. It is fairly easy to show that is not a compact operator. However, we shall show that is compact. As an intermediate step, we prove the following proposition.
Proposition 3.1.
Let and let . Let . For each , let be a family of functions where all satisfy . Assume furthermore that for all we have that
| (11) |
Then
for all if and if .
Applying the above proposition times, we obtain the following corollary.
Corollary 3.1.
Let and let be a collection of subsets of as in Proposition 3.1. Then .
The corollary is thus a way of converting a statement of the convergence of the family of functions on individual sets of the form to convergence of with respect to the -norm.
Proof of Proposition 3.1..
Our first claim is that (11) implies that . Let . Let be a finite collection of digits such that the set
satisfies for all . By (11) , there exists some for which for all and with . For any , there exists a with . Hence by uniform boundedness of the Hölder norm,
Letting , we obtain the claim. It suffices therefore to show that . Assume for the moment that .
We need uniform estimates on the Hölder seminorm of the function.
| (12) |
where . Let and note that there exists some depending only on and such that
for all with , since is analytic and as . We now estimate the contribution to the Hölder seminorm from the second term in (12). For , denote and . Let such that for all with :
for all . The existence of such an follows from uniform boundedness of the Hölder norm and the estimate (10) we used in the proof of Lemma 2.1. Furthermore, it is a fairly straightforward exercise to show that (11) implies that
for all with greater than some and with . We obtain by (12) and the above three equations that
for all with . The lemma for follows from letting . For the case , replace , and with and respectively.
∎
Proposition 3.2.
The operator is compact for each and .
Proof.
Let denote the set of functions with . We shall prove is precompact by showing that any sequence in the set has a Cauchy subsequence.
Let be a sequence of functions in and let be a sequence of functions in with for all . For any there exists a subsequence such that is a Cauchy sequence. By a diagonal argument, we may assume is Cauchy for all .
For any , let be the closed interval spanned by and . The intersection is finite. Indeed, it consists of the endpoints, the points with and the points with and .
Since the functions restricted to are piecewise linear functions interpolated at a finite number of points for which is Cauchy, the functions form a Cauchy sequence in . Since is analytic on a neighbourhood of , it follows that is Cauchy on . In particular, the families of functions defined by
satisfy the conditions of Proposition 3.1. By Corollary 3.1, we thus obtain that
which shows is a Cauchy subsequence. ∎
Let us now prove most of Theorem 1.1.
Theorem 3.1.
Let . Then for all with the transfer operator satisfies
Proof.
By Nussbaum’s formula and Proposition 3.2, it suffices to prove that there is some constant for which
We show in Lemma A.2 in the appendix that there exists a constant for which for all .
For fixed , we first claim that
for all . Let . For notational simplicity, let , which by the aforementioned lemma satisfies for all . Since we have that
It follows from the above formula that since for all with , the equality holds for all . By repeating this argument, we obtain that for all . By definition of the -seminorm, we obtain the inequality , where is the maximal distance of a point to . We thus obtain
Letting proves the claim.
To show the theorem use the claim for and Lemma 2.1 to obtain for any that
Since is arbitrary, the above also holds for . Applying the claim for we obtain once again by Lemma 2.1 that
Letting proves the theorem.
∎
Remark 3.1.
By using the spectral radius formula and bounding every function by a constant, we obtain that and hence that
where for is the eigenvalue of maximum modulus of . The eigenvalue is guaranteed to be simple and positive by the Perron-Frobenius theorem. In fact is a real-analytic and decreasing function of with , see [15, 9].
Corollary 3.2.
Let satisfy . Then is an eigenvalue of if and only if it is an eigenvalue of . In this case the respective eigenspaces and are equal.
This is an example of a more general principle for operators on Banach spaces, see e.g. [10]. For our purposes, we may directly adapt a proof due to Ruelle (Corollary 3.3 in [21]), which we provide in the appendix.
Proof of Theorem 1.1.
Let and let . We first assume and adapt a standard trick (see e.g. [6] III) to reduce to the problem on . We note that for we obtain by first applying the product rule for higher derivatives and then the chain rule that there exists coefficients with for which
This holds using analytic continuation to . Recall that the any composition of operators is compact if one of them is compact and that the operator which sends to is compact for . Hence we can write
where is some compact operator from . Since the essential spectral radius does not depend on compact perturbations, we have that the essential spectral radius of as an operator on is the same as the essential spectral radius of as an operator on . This follows from (7), but is perhaps easier to see using the Banach space isomorphism . The case follows mutatis mutandis, replacing with . Hence Theorem 1.1 for follows from the case . ∎
Remark 3.2.
The zeroes of on the line are those for which there are eigenvalues of called ‘Maass cusp forms’ with eigenvalue . Furthermore the multiplicity of the zeroes and the eigenspaces of match. The other zeroes of are a simple zero at and at the values of for which is a nontrivial zero of the Riemann zeta function . Since when , we see that the eigenspaces of and for the eigenvalues are identical when . In particular, the eigenspaces associated to the Maass cusp forms are preserved when , and the eigenspaces associated to the nontrivial zeroes of the Riemann zeta function111In fact, the zeroes of associated with the nontrivial zeroes of always occur due to being an eigenvalue of . to the right of the critical line are preserved when .
We conclude with a fairly straightforward remark which will make the next section easier from a notational perspective.
Remark 3.3.
We can of course consider the operator acting on spaces other than or . For example, it is clear that the estimates on the essential spectrum still hold for acting on with and . As we will remark upon Section 4, eigenfunctions of extend to functions on , so the point spectrum beyond the estimate of Theorem 1.1 does not depend on . In fact by letting and , we see that a natural space for is the space , which we define to be the Fréchet space of functions such that on every compact we have that is -times differentiable such that the -th derivative is continuous if and is -Hölder continuous if not.
4 Relation with the three-term functional equation.
In this section, we consider the following functional equation.
| (13) |
where . This equation was first introduced (in slightly modified form) for in [14] and for in [13]. The formulation in this document can be found in [4]. In particular, we consider this functional equation for functions on of varying regularity and for holomorphic functions on .
It is fairly easy to prove that any eigenfunction of with eigenvalue satisfies (13). Indeed, this follows from the identity , see e.g. Proposition 7 in [4]. Furthermore, from the identity
we can use the contraction properties of the inverse branches inductively to extend to a function which is holomorphic on .
Conversely, the space of solutions to (13) is uncountable-dimensional, something already remarked upon in [13]. However, we obtain a one-to-one correspondence after imposing a condition on the asymptotic behaviour of . Let us first describe the asymptotic behaviour of a generic solution in the non-analytic case (e.g. at Hölder regularity).
Define be the space of all -times continuously differentiable functions on the torus with -Hölder derivative if is not an integer. We identify these functions with -periodic functions on by setting for .
Proposition 4.1.
Let solve (13) for some . Assume furthermore that and . Then there exist a function and coefficients depending only on and the derivatives such that
| (14) |
and
| (15) |
where we may replace the with respectively if .
Proof.
This is essentially the proof of the proposition in Chapter III section 3 of [13] adapted to the non-analytic case and including the case .
Define the function , where we extend to an operator on as in Remark 3.3. For it is clear that
and this formula continues to hold for all (with ) by analytic continuation. It is also clear from the definition that .
Assume with We now consider the asymptotic behaviour of . We note that if and ,
| (16) |
where is the -Hölder coefficient of on . Hence we obtain by an application of the integral test that if ,
as with the implied constant easily computable and depending on and . We therefore obtain by (4) that
| (17) |
We obtain (14) for by writing using the identity and the well-known expansion (see [18] §25.11)
| (18) |
where are the Bernoulli numbers.
For , we have that the left hand side of (16) is as and we follow our previous argument, mutatis mutandis. The asymptotic behavior of for small follows from the identity , using the asymptotics (14) to obtain the expansion for as and the expansion .
∎
We see by Proposition that a generic solution of (13) in satisfies if . On the other hand, we see that the condition in (14) or in (15) is equivalent to the condition , which by Theorem 1.1 implies extends to a holomorphic function in if . For example, we have the following corollary of Proposition 4.1, which we can view as a stronger version of the Bootstrapping procedure by Lewis and Zagier in [13].
Corollary 4.1.
Let . Suppose is some solution of (13) for . Assume furthermore it satisfies the mild growth condition as . Then if is -Hölder for some , it is automatically holomorphically extendable to the cut plane .
This corollary immediately follows from Theorem 1.1 and the fact that if . We remark that when and or , the space of eigenvalues of is in bijection with the space of even, respectively uneven Maass cusp forms on the modular surface with eigenvalue and we have that , i.e. as , see [14, 13].
Proposition 4.1 showed that a solution of the three term equation has the associated periodic function . By applying the inverse, we obtain the following proposition.
Proposition 4.2.
Proof.
We interpret in the sense of Remark 3.3. More specifically, for acting on , the expression is meromorphic and satisfies
| (19) |
if is large enough. By Theorem 1.1, we see that is a meromorphic operator-valued function for all . Hence the functions are meromorphic functions over the half-plane consisting of all satisfying the conditions in the proposition. We see by the second line of (19) and uniqueness of analytic continuation that these functions agree for different on the intersection of their domains. Hence we may define for any by choosing in (19) small enough.
We may thus define
For large enough and use (19) to obtain the analytic continuation. It is clear from the above formula that satisfies (13) for large and hence by analytic continuation for all satisfying the condition in the proposition.
The proposition follows immediately from the proof of Proposition 4.1. ∎
From the discussion so far we obtain the following theorem in a straightforward manner.
Theorem 4.1.
Proof.
Remark 4.1.
Proposition 4.2 does not give us a construction of solutions of (13) when is an eigenvalue of . We remark that in the case , explicit examples are constructed in [13], showing that the spaces are still uncountable-dimensional. For example, the function , where and is -periodic. These constructions seem to rely on an additional symmetry for solutions of (13) with . We conjecture these might not necessarily exist for other values of .
5 Further avenues of exploration.
This preprint provides estimates for the essential spectral radius for the mayer transfer operator, but the technique should be generalizable to a large class of transfer operators associated to the geodesic flow on noncompact geometrically finite Fuchsian groups. We refer to e.g. [5, 17, 20, 8]. One could also think about applying a similar approach in higher dimensions, i.e. for families transfer operators associated to multidimensional continued fraction algorithms or other expanding maps on hypercubes. However, the significance behind the analytic continuation of these operators is less clear.
Appendix A Appendix
A.1 Hölder Estimates
We begin this appendix with a lemma that will be useful to estimate Hölder norms.
Lemma A.1.
Let . Let and let be a function. Then there exists a constant independent of for which
Proof.
We first start with the following basic inequality. Let . Then
This can be proved by noting that if the left hand side is maximised for and if then it is maximised when two of the parameters are zero. Assume first that . Then we can write as the dot product
In particular, by submultiplicativity of the dot product with respect to the usual vector norm and the aforementioned inequality we have
which proves the lemma for . For the proof is analogous . ∎
Lemma A.2.
The operator is well-defined, continuous and has uniformly bounded operator norm over all .
Proof.
It is fairly straightforward to show that for the assignment is linear. We therefore just need to show continuity and Hölder continuity of and bound the operator norm.
We note that if is a connected component, then by the intermediate value theorem, which proves continuity at the accumulation points and hence continuity of everywhere.
We now estimate the Hölder norm. Let be two points in . We estimate the quantity . By perturbing the two points by an arbitrarily small amount, we may assume and are not accumulation points of . Let and be the connected components of such that and . In case the two intervals are the same, it is clear that . In case they are different, we let on as in Lemma A.1, whence we obtain that
The lemma follows after noting that coincides with on and using that
∎
A.2 Ruelle’s Corollary
If is a bounded linear operator between Banach spaces, denote its kernel by . If has closed image , denote its cokernel by . We say is Fredholm of index if has closed image, and are finite and . The composition of two Fredholm operators with respective indices is Fredholm with index . For a Banach space , we denote its dual by and denote by the induced adjoint operator . The following is a straightforward exercise in functional analysis.
Lemma A.3.
If has closed image and is finite-dimensional, then .
Suppose now , so . We recall the following facts about the spectral theory of operators on Banach spaces
-
•
If , then is Fredholm of index .
-
•
If is a compact operator, so is .
We refer the reader to [7] for more details on the spectral theory of Banach operators, noting that in this preprint corresponds to in their book.
Proof of Corollary 3.2..
We remind the reader that this proof is a direct adaptation from [21]. It suffices to prove that any eigenvalue of with is an eigenvalue of and that the dimensions of the respective generalised eigenspaces and are equal. We note that the unions are over a finite number of nontrivial kernels by definition of the essential spectral radius. By Lemma A.3 and the fact that is compact operator and that the relevant operators are Fredholm of index 0, it suffices to prove that the generalised eigenspaces and belonging respectively to and have equal dimension.
There is a natural (injective) inclusion map defined by restriction to . The dual operator is defined by restricting distributions on to distributions on . This restricts to a natural linear map from to . Surjectivity of this map follows from the Hahn-Banach theorem. We now show injectivity. Suppose that for some , its restriction is identically zero. We show .
In order to do so, let such that . Use Hahn-Banach to extend to . For any , let be a series of smooth approximations which converge in the -norm. By cutting of fourier expansions outside of balls of increasing radii in Fourier space, we may assume the maps are analytically extendable to elements of . Hence we obtain
which is what we had to show. ∎
Appendix B Acknowledgements
References
- [1] Felix E. Browder “On the spectral theory of elliptic differential operators. I” In Math. Ann. 142, 1961, pp. 22–130 DOI: 10.1007/BF01343363
- [2] Oliver Butterley, Niloofar Kiamari and Carlangelo Liverani “Locating Ruelle-Pollicott resonances” In Nonlinearity 35.1, 2022, pp. 513–566 DOI: 10.1088/1361-6544/ac3ad5
- [3] Cheng-Hung Chang and Dieter Mayer “The period function of the nonholomorphic Eisenstein series for ” In Math. Phys. Electron. J. 4, 1998, pp. Paper 6\bibrangessep8
- [4] Cheng-Hung Chang and Dieter H. Mayer “The transfer operator approach to Selberg’s zeta function and modular and Maass wave forms for ” In Emerging applications of number theory (Minneapolis, MN, 1996) 109, IMA Vol. Math. Appl. Springer, New York, 1999, pp. 73–141 DOI: 10.1007/978-1-4612-1544-8“˙3
- [5] Cheng-Hung Chang and Dieter H. Mayer “An extension of the thermodynamic formalism approach to Selberg’s zeta function for general modular groups” In Ergodic theory, analysis, and efficient simulation of dynamical systems Springer, Berlin, 2001, pp. 523–562
- [6] Pierre Collet and Stefano Isola “On the essential spectrum of the transfer operator for expanding Markov maps” In Comm. Math. Phys. 139.3, 1991, pp. 551–557 URL: http://projecteuclid.org/euclid.cmp/1104203470
- [7] D.. Edmunds and W.. Evans “Spectral theory and differential operators”, Oxford Mathematical Monographs Oxford University Press, Oxford, 2018, pp. xviii+589 DOI: 10.1093/oso/9780198812050.001.0001
- [8] Ksenia Fedosova and Anke Pohl “Meromorphic continuation of Selberg zeta functions with twists having non-expanding cusp monodromy” In Selecta Math. (N.S.) 26.1, 2020, pp. Paper No. 9\bibrangessep55 DOI: 10.1007/s00029-019-0534-3
- [9] Philippe Flajolet and Brigitte Vallée “Continued fraction algorithms, functional operators, and structure constants” In Theoret. Comput. Sci. 194.1-2, 1998, pp. 1–34 DOI: 10.1016/S0304-3975(97)00123-0
- [10] Sandy Grabiner “Spectral consequences of the existence of intertwining operators” In Comment. Math. Prace Mat. 22.2, 1981, pp. 227–238
- [11] Henryk Iwaniec “Introduction to the spectral theory of automorphic forms”, Biblioteca de la Revista Matemática Iberoamericana. [Library of the Revista Matemática Iberoamericana] Revista Matemática Iberoamericana, Madrid, 1995, pp. xiv+247
- [12] Tosio Kato “Perturbation theory for linear operators” Band 132, Grundlehren der Mathematischen Wissenschaften Springer-Verlag, Berlin-New York, 1976, pp. xxi+619
- [13] J. Lewis and D. Zagier “Period functions for Maass wave forms. I” In Ann. of Math. (2) 153.1, 2001, pp. 191–258 DOI: 10.2307/2661374
- [14] John B. Lewis “Spaces of holomorphic functions equivalent to the even Maass cusp forms” In Invent. Math. 127.2, 1997, pp. 271–306 DOI: 10.1007/s002220050120
- [15] Dieter H. Mayer “On the thermodynamic formalism for the Gauss map” In Comm. Math. Phys. 130.2, 1990, pp. 311–333 URL: http://projecteuclid.org/euclid.cmp/1104200514
- [16] Dieter H. Mayer “The thermodynamic formalism approach to Selberg’s zeta function for ” In Bull. Amer. Math. Soc. (N.S.) 25.1, 1991, pp. 55–60 DOI: 10.1090/S0273-0979-1991-16023-4
- [17] Takehiko Morita “Markov systems and transfer operators associated with cofinite Fuchsian groups” In Ergodic Theory Dynam. Systems 17.5, 1997, pp. 1147–1181 DOI: 10.1017/S014338579708632X
- [18] “NIST Digital Library of Mathematical Functions” F. W. J. Olver, A. B. Olde Daalhuis, D. W. Lozier, B. I. Schneider, R. F. Boisvert, C. W. Clark, B. R. Miller, B. V. Saunders, H. S. Cohl, and M. A. McClain, eds., https://dlmf.nist.gov/, Release 1.2.4 of 2025-03-15 URL: https://dlmf.nist.gov/25.11
- [19] Roger D. Nussbaum “The radius of the essential spectrum” In Duke Math. J. 37, 1970, pp. 473–478 URL: http://projecteuclid.org/euclid.dmj/1077379127
- [20] Anke D. Pohl “Symbolic dynamics, automorphic functions, and Selberg zeta functions with unitary representations” In Dynamics and numbers 669, Contemp. Math. Amer. Math. Soc., Providence, RI, 2016, pp. 205–236 DOI: 10.1090/conm/669/13430
- [21] David Ruelle “The thermodynamic formalism for expanding maps” In Comm. Math. Phys. 125.2, 1989, pp. 239–262 URL: http://projecteuclid.org/euclid.cmp/1104179466