这是indexloc提供的服务,不要输入任何密码

On the Ruelle-Mayer Transfer Operators for Hölder Continuous Functions

Alexander Baumgartner Centro di Ricerca Matematica Ennio De Giorgi, Scuola Normale Superiore, Italy
(9th November 2025)
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 α\alpha-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 (1/2+ε)(1/2+\varepsilon) and 3/43/4 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 G:(0,1)[0,1)G:(0,1)\to[0,1) be the Gauss map

G(x)=1z1z.G(x)=\frac{1}{z}-\left\lfloor\frac{1}{z}\right\rfloor. (1)

We associate to GG the following family of transfer operators which we formally define as acting on some space of functions f:[0,1]f:[0,1]\to\mathbb{C}:

β(f)(z)=n=11(n+z)2βf(1n+z),\mathcal{L}_{\beta}(f)(z)=\sum_{n=1}^{\infty}\frac{1}{(n+z)^{2\beta}}f\left(\frac{1}{n+z}\right), (2)

where β{z:(β)>1/2}\beta\in\{z\in\mathbb{C}:\Re(\beta)>1/2\}. When β=1\beta=1, this is the familiar Ruelle-Perron-Frobenius operator for GG and can formally be interpreted as the right-adjoint of the Koopman operator ffGf\mapsto f\circ G. The operators in the case β1\beta\neq 1 were first introduced by Mayer [15], who considered them acting on the Banach space of functions A(D)A_{\infty}(D) which are holomorphic on D:={z:|z1|<3/2}D:=\{z\in\mathbb{C}:\left\lvert z-1\right\rvert<3/2\} with continuous extension to D¯\overline{D}.

For this choice of Banach space, Mayer also showed that the operator-valued function ββ\beta\mapsto\mathcal{L}_{\beta} admits a meromorphic continuation on \mathbb{C}, with simple poles at β=k/2\beta=k/2 for k{1,0,1,2,}k\in\{1,0,-1,-2,\ldots\}. The continuation for (β)>0\Re(\beta)>0 can be obtained by rewriting f=f(0)+ff=f(0)+f^{\star} with f:=ff(0)f^{\star}:=f-f(0) and noting that

β(f)(z)=f(0)ζH(2β,z+1)+n=11(n+z)2β(f)(1n+z),\mathcal{L}_{\beta}(f)(z)=f(0)\zeta_{H}(2\beta,z+1)+\sum_{n=1}^{\infty}\frac{1}{(n+z)^{2\beta}}(f^{\star})\left(\frac{1}{n+z}\right), (3)

where ζH(s,z)=n=0(n+z)s\zeta_{H}(s,z)=\sum_{n=0}^{\infty}(n+z)^{-s} is the Hurwitz zeta function, which has an analytic continuation for all s1s\neq 1 and a simple pole of residue 11 at s=1s=1. In fact, it can further be analytically continued to βk/2\beta\geq-k/2 for any kk\in\mathbb{N} by noticing that

β(f)(z)=n=0kdnfdzn(0)ζH(n+2β,z+1)n!+n=11(n+z)2β(f)(1n+z),\mathcal{L}_{\beta}(f)(z)=\sum_{n=0}^{k}\frac{d^{n}f}{dz^{n}}(0)\frac{\zeta_{H}(n+2\beta,z+1)}{n!}+\sum_{n=1}^{\infty}\frac{1}{(n+z)^{2\beta}}(f^{\star})\left(\frac{1}{n+z}\right), (4)

where f(z)=f(z)n=0kdnfdnz(0)zn/n!f^{\star}(z)=f(z)-\sum_{n=0}^{k}\frac{d^{n}f}{d^{n}z}(0)z^{n}/n!. Furthermore β\mathcal{L}_{\beta} is a nuclear operator on A(D)A_{\infty}(D) of order 0 for all β\beta.

The Mayer transfer operator encodes geometric information about the modular surface, which is the quotient orbifold 𝐌\mathbf{M} obtained from considering the Poincaré upper half plane :{x+iy:x,y,y>0}\mathbb{H}:\{x+iy:x,y\in\mathbb{R},y>0\} equipped with the usual Riemannian metric ds2=(dx2+dy2)/y2ds^{2}=(dx^{2}+dy^{2})/y^{2} and identifying points via equivalence relation z(az+b)/(cz+d)\mathbb{H}\ni z\sim(az+b)/(cz+d) for all a,b,c,da,b,c,d\in\mathbb{Z} with adbc=1ad-bc=1. Since β\mathcal{L}_{\beta} is a nuclear operator, the Fredholm determinants det(1β)\det(1-\mathcal{L}_{\beta}) and det(1+β)\det(1+\mathcal{L}_{\beta}) are well-defined. These relate to Selberg Zeta function Z(s)Z(s) by the formula

Z(s)=det(1s)(1+s),Z(s)=\det(1-\mathcal{L}_{s})(1+\mathcal{L}_{s}), (5)

see [16]. The Selberg zeta function is closely connected with the spectral theory of the Laplace-Beltrami operator Δ:=y2(d2/dx2+d2/dy2)\Delta:=-y^{2}(d^{2}/dx^{2}+d^{2}/dy^{2}) [11]. Equation (3) is thus a way of connecting the spectral theory of Δ\Delta with that of β\mathcal{L}_{\beta}. 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 s\mathcal{L}_{s} survive when we consider e.g. Hölder continuous functions instead. For α(0,1)\alpha\in(0,1), let Cα([0,1])C^{\alpha}([0,1]) denote the space of α\alpha-Hölder functions on [0,1][0,1]. If α=k+η\alpha=k+\eta with kk\in\mathbb{N} and η[0,1)\eta\in[0,1), let Cα([0,1])C^{\alpha}([0,1]) be the set of functions for which the kk-th derivative exists and is continuous or η\eta-Hölder if η=0\eta=0 or η(0,1)\eta\in(0,1) respectively. Denote by LβL_{\beta} the operator (4) acting on Cα([0,1])C^{\alpha}([0,1]), which we shall show to be well-defined and continuous for large enough α\alpha. These operators are not compact, but we can decompose the spectrum σ(Lβ)\sigma(L_{\beta}) into the essential spectrum σe(Lβ)\sigma_{e}(L_{\beta}) and the discrete spectrum σdisc(Lβ)\sigma_{\mathrm{disc}}(L_{\beta}) which consists of isolated eigenvalues of finite algebraic multiplicity. We prove the following.

Theorem 1.1.

On the half-plane (β)>(1α)/2\Re(\beta)>(1-\alpha)/2, the radius ρe(Lβ):=supλσε(Lβ)|λ|\rho_{e}(L_{\beta}):=\sup_{\lambda\in\sigma_{\varepsilon}(L_{\beta})}\left\lvert\lambda\right\rvert of the essential spectrum satisfies

ρe(Lβ)ρ((β)+α):=supλσ(β+α)|λ|.\rho_{e}(L_{\beta})\leq\rho(\mathcal{L}_{\Re(\beta)+\alpha}):=\sup_{\lambda\in\sigma(\mathcal{L}_{\beta+\alpha})}\left\lvert\lambda\right\rvert.

Furthermore, the generalised eigenspaces belonging to eigenvalues λ\lambda with |λ|>ρe(Lβ)\left\lvert\lambda\right\rvert>\rho_{e}(L_{\beta}) consist of analytic functions which extend to generalised eigenfunctions of β\mathcal{L}_{\beta}.

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 α\alpha is a natural number, the proof of Theorem 1.1 is a mundane application of standard techniques.

However, the proof of the case 1/2<α<11/2<\alpha<1 is perhaps more interesting. In that case, we still obtain uniformly bounded estimates for ρe(Lβ)\rho_{e}(L_{\beta}) on vertical lines (β)=σ\Re(\beta)=\sigma even when σ<1/2\sigma<1/2, where the function βLβ\beta\mapsto L_{\beta} is unbounded. However, by (2), the unboundedness in β\beta is due to the contribution rank one operator ff(0)ζH(2β,z+1)f\mapsto f(0)\zeta_{H}(2\beta,z+1), 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 (1,)(1,\infty) satisfying certain growth conditions automatically extend to holomorphic functions on the cut plane \(,1]\mathbb{C}\backslash(-\infty,-1]. We show that for (β)>0\Re(\beta)>0, we may even replace real-analyticity with a Hölder condition and obtain that it is still the restriction of a holomorphic function on \(,1]\mathbb{C}\backslash(-\infty,-1].

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 (1/(n+1),1/n],n(1/(n+1),1/n],n\in\mathbb{N}. Each of these intervals are mapped by GG homeomorphically onto [0,1)[0,1). We denote by ψn:[0,1)(1/(n+1),1/n]\psi_{n}:[0,1)\to(1/(n+1),1/n] the inverse branch ψn(x)=1/(n+x)\psi_{n}(x)=1/(n+x). Given a sequence n1,n2,,nkn_{1},n_{2},\ldots,n_{k}\in\mathbb{N} we denote by ψn1,n2,,nk\psi_{n_{1},n_{2},\ldots,n_{k}} the map defined by

ψn1,n2,,nk:=ψn1ψnk1ψnk(x)=1n1+1+1nk+x.\psi_{n_{1},n_{2},\ldots,n_{k}}:=\psi_{n_{1}}\circ\cdots\psi_{n_{k-1}}\circ\psi_{n_{k}}(x)=\frac{1}{n_{1}+\frac{1}{\ddots+\frac{1}{n_{k}+x}}}. (6)

We also denote the right-hand side by [n1,,nk+x][n_{1},\ldots,n_{k}+x] as a simplifying bit of notation and we let n1,,nk\langle n_{1},\ldots,n_{k}\rangle denote the set {[n1,,nk+x]:x[0,1)}\{[n_{1},\ldots,n_{k}+x]:x\in[0,1)\}.

2.2 The Essential Spectral Radius

In Section 3, we provide estimates for the essential spectrum of LβL_{\beta}, so we provide a brief reminder of some definitions. Let II be the identity operator. We recall that the spectrum σ(T)\sigma(T)\subset\mathbb{C} of an operator TT on a Banach space BB is the set of all λ\lambda\in\mathbb{C} for which (TλI)(T-\lambda I) is not invertible. Following Browder [1], we define the essential spectrum σε(T)σ(T)\sigma_{\varepsilon}(T)\subset\sigma(T) to be the set of all λ\lambda for which (TλI)B(T-\lambda I)B is not closed, r=1+ker(TλI)r\bigcup_{r=1}^{+\infty}\ker(T-\lambda I)^{r} is infinite-dimensional or for which λ\lambda is a limit point of σ(T)\sigma(T). A nice consequence of this definition is that σ(T)σε(T)\sigma(T)\setminus\sigma_{\varepsilon}(T) 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 KK be the subspace of compact operators on BB. Define the seminorm K\|\cdot\|_{K} on the space of bounded linear operators on BB by

TK=min{TT:TK},\|T\|_{K}=\min\{\|T-T^{\prime}\|:T^{\prime}\in K\},

and let T:BBT:B\to B be a bounded linear operator. Then its essential spectral radius satisfies

ρe(T):=supλσε(T)|λ|=liml+TlK1/l.\rho_{e}(T):=\sup_{\lambda\in\sigma_{\varepsilon}(T)}\left\lvert\lambda\right\rvert=\lim_{l\to+\infty}\|T^{l}\|_{K}^{1/l}. (7)

2.3 Definition of Transfer Operators

In this subsection, we show that the Mayer transfer operators with (β)>0\Re(\beta)>0 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 0<α<10<\alpha<1 by Cα([0,1])C^{\alpha}([0,1]) the space of α\alpha-Hölder continuous functions on [0,1][0,1] equipped with the norm

f:=f+fα,where fα=supx,y:xy|f(x)f(y)||xy|α\|f\|:=\|f\|_{\infty}+\|f\|_{\alpha},\text{where }\|f\|_{\alpha}=\sup_{\begin{subarray}{c}x,y:\\ x\neq y\end{subarray}}\frac{\lvert f(x)-f(y)\rvert}{\left\lvert x-y\right\rvert^{\alpha}}

and f=sup[0,1]|f|\|f\|_{\infty}=\sup_{[0,1]}\left\lvert f\right\rvert. It will turn out to be useful to define for fCα([0,1])f\in C^{\alpha}([0,1]) the auxiliary function |f|α\lvert f\rvert_{\alpha} on [0,1][0,1] by

|f|α(x)=supy:xy|f(x)f(y)||xy|α.\lvert f\rvert_{\alpha}(x)=\sup_{\begin{subarray}{c}y:\\ x\neq y\end{subarray}}\frac{\lvert f(x)-f(y)\rvert}{\left\lvert x-y\right\rvert^{\alpha}}.

We remark that |f|α=fα\|\lvert f\rvert_{\alpha}\|_{\infty}=\|f\|_{\alpha}.

For (β)>1/2\Re(\beta)>1/2, the operator (2) is also well defined on the space L([0,1])L^{\infty}([0,1]) of bounded functions equipped with the \|\cdot\|_{\infty}-norm. We denote the operator acting on this space by 𝔏β\mathfrak{L}_{\beta}.

Lemma 2.1.

Let α(0,1]\alpha\in(0,1], (β)>(1α)/2\Re(\beta)>(1-\alpha)/2 with β1/2\beta\neq 1/2. Then the operator LβL_{\beta} is well-defined and continuous on Cα([0,1])C^{\alpha}([0,1]). Furthermore there exists a constant CC and a family of constants (Dε)ε>0(D_{\varepsilon})_{\varepsilon>0} depending only on α\alpha and (β)\Re(\beta) such that

Lβ(f)Cf, and \left\|L_{\beta}(f)\right\|_{\infty}\leq C\|f\|\text{, and }
|Lβ(f)|α(x)𝔏(β)+α((1+ε)|f|α)(x)+Dεf for all x[0,1],ε>0.\lvert L_{\beta}(f)\rvert_{\alpha}(x)\leq\mathfrak{L}_{\Re(\beta)+\alpha}((1+\varepsilon)|f|_{\alpha})(x)+D_{\varepsilon}\|f\|_{\infty}\text{ for all }x\in[0,1],\varepsilon>0.
Proof.

We note that for fCα([0,1])f\in C^{\alpha}([0,1]) we have that |f(x)f(0)|fαxα\left\lvert f(x)-f(0)\right\rvert\leq\|f\|_{\alpha}x^{\alpha} as x0x\to 0, hence we have that

|Lβ(f)(z)|f|ζH(2β,z+1)|+fαn=11(n+z)2(β)+α.\left\lvert L_{\beta}(f)(z)\right\rvert\leq\|f\|_{\infty}\left\lvert\zeta_{H}(2\beta,z+1)\right\rvert+\|f\|_{\alpha}\sum_{n=1}^{\infty}\frac{1}{(n+z)^{2\Re(\beta)+\alpha}}. (8)

To bound the auxiliary function, we start with a basic application of Taylor’s theorem. Let θ\theta be a complex number with (θ)>0\Re(\theta)>0. Let x,y[0,1]x,y\in[0,1]. Then

|(n+y)θ(n+x)θ+θ(n+y)θ1(yx)|K(yx)2\left\lvert(n+y)^{-\theta}-(n+x)^{-\theta}+\theta(n+y)^{-\theta-1}(y-x)\right\rvert\leq K(y-x)^{2} (9)

for a constant K>0K>0 which can be uniformly bounded over all x,yx,y and n1n\geq 1.

Note that

f(1n+x)(n+x)2βf(1n+y)(n+y)2β=(f(1n+x)(n+x)2βf(1n+y)(n+x)2β)+(f(1n+y)(n+x)2βf(1n+y)(n+y)2β).\frac{f\left(\frac{1}{n+x}\right)}{(n+x)^{2\beta}}-\frac{f\left(\frac{1}{n+y}\right)}{(n+y)^{2\beta}}=\left(\frac{f\left(\frac{1}{n+x}\right)}{(n+x)^{2\beta}}-\frac{f\left(\frac{1}{n+y}\right)}{(n+x)^{2\beta}}\right)+\left(\frac{f\left(\frac{1}{n+y}\right)}{(n+x)^{2\beta}}-\frac{f\left(\frac{1}{n+y}\right)}{(n+y)^{2\beta}}\right).

By first using the definition of the auxiliary function, and then applying (9) for θ=α\theta=\alpha we find that there exists some constant R>0R>0 which is independent of ff for which

|f(1n+x)f(1n+y)||f|α(1n+x)|xy(n+y)(n+x)|α|f|α(1n+x)|xy|α(n+x)2α(1+R|xy|n1+α).\begin{split}\left\lvert f\left(\frac{1}{n+x}\right)-f\left(\frac{1}{n+y}\right)\right\rvert&\leq\lvert f\rvert_{\alpha}\left(\frac{1}{n+x}\right)\left\lvert\frac{x-y}{(n+y)(n+x)}\right\rvert^{\alpha}\\ &\leq\lvert f\rvert_{\alpha}\left(\frac{1}{n+x}\right)\frac{\left\lvert x-y\right\rvert^{\alpha}}{(n+x)^{2\alpha}}\left(1+\frac{R\left\lvert x-y\right\rvert}{n^{1+\alpha}}\right).\end{split}

By estimating the second term using (9) for θ=2β\theta=2\beta, we obtain that

|f(1n+x)(n+x)2βf(1n+y)(n+y)2β||xy|α|f|α(1n+x)(n+x)2((β)+α)(1+R|xy|n1+α)+D|xy|1α|f(1n+y)|(n+y)2β+1\frac{\left\lvert\frac{f\left(\frac{1}{n+x}\right)}{(n+x)^{2\beta}}-\frac{f\left(\frac{1}{n+y}\right)}{(n+y)^{2\beta}}\right\rvert}{\left\lvert x-y\right\rvert^{\alpha}}\leq\frac{\lvert f\rvert_{\alpha}(\frac{1}{n+x})}{(n+x)^{2(\Re(\beta)+\alpha)}}\left(1+\frac{R\left\lvert x-y\right\rvert}{n^{1+\alpha}}\right)+D^{\prime}\left\lvert x-y\right\rvert^{1-\alpha}\frac{\left\lvert f\left(\frac{1}{n+y}\right)\right\rvert}{(n+y)^{2\beta+1}} (10)

for some uniformly bounded constant D>0D^{\prime}>0. Let ε>0\varepsilon>0. If |xy|ε/R\left\lvert x-y\right\rvert\leq\varepsilon/R, we sum over all nn in (10) to obtain

n=1|f(1n+x)(n+x)2βf(1n+y)(n+y)2β||xy|α𝔏(β)+α((1+ε)|f|α)(x)+DL(β)+1/2|f|,\sum_{n=1}^{\infty}\frac{\left\lvert\frac{f\left(\frac{1}{n+x}\right)}{(n+x)^{2\beta}}-\frac{f\left(\frac{1}{n+y}\right)}{(n+y)^{2\beta}}\right\rvert}{\left\lvert x-y\right\rvert^{\alpha}}\leq\mathfrak{L}_{\Re(\beta)+\alpha}((1+\varepsilon)|f|_{\alpha})(x)+D^{\prime}\|L_{\Re(\beta)+1/2}|f|\|_{\infty},

where |f||f| denotes the function x|f(x)|x\mapsto|f(x)|. Else let NN\in\mathbb{N} be large enough such that R/N1+αεR/N^{1+\alpha}\leq\varepsilon. Sum both sides of the inequality

|f(1n+x)(n+x)2βf(1n+y)(n+y)2β||xy|α2fn2(β)|xy|α\frac{\left\lvert\frac{f\left(\frac{1}{n+x}\right)}{(n+x)^{2\beta}}-\frac{f\left(\frac{1}{n+y}\right)}{(n+y)^{2\beta}}\right\rvert}{\left\lvert x-y\right\rvert^{\alpha}}\leq\frac{2\|f\|_{\infty}}{n^{2\Re(\beta)}\left\lvert x-y\right\rvert^{\alpha}}

for n{1,2,,N}n\in\{1,2,\ldots,N\} and add the respective sides in (10) over all n>Nn>N. After taking the supremum over all yy we then obtain that |Lβ(f)|α(x)𝔏(β)+α((1+ε)|f|α)(x)+EεL(β)+1/2|f|\lvert L_{\beta}(f)\rvert_{\alpha}(x)\leq\mathfrak{L}_{\Re(\beta)+\alpha}((1+\varepsilon)|f|_{\alpha})(x)+E_{\varepsilon}\|L_{\Re(\beta)+1/2}|f|\|_{\infty} for some Eε>0E_{\varepsilon}>0. The lemma follows from boundedness of the operator L(β)+1/2L_{\Re(\beta)+1/2} on L([0,1])L^{\infty}([0,1])

Remark 2.1.

To apply the usual method for obtaining quasi-compactness with essential spectral radius less than rr, we would need an estimate of the form

LβnfαRnf+rnfα,\|L_{\beta}^{n}f\|_{\alpha}\leq R_{n}\|f\|_{\infty}+r_{n}\|f\|_{\alpha},

where Rn,rn>0R_{n},r_{n}>0 and lim infn(rn)1/nr\liminf_{n}(r_{n})^{1/n}\leq r. It is indeed possible to find an upper bound for Lβnfα\|L_{\beta}^{n}f\|_{\alpha} by iteratively using Lemma 2.1. Unfortunately, this does not yield a quantitatively good upper bound, both due to the dependence on fα\|f\|_{\alpha} in the estimate (8) and when considering β\beta for which supz[0,1]|ζH(2β,z+1)|\sup_{z\in[0,1]}\left\lvert\zeta_{H}(2\beta,z+1)\right\rvert 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 UCU\subset C be a simply connected open set of \mathbb{C}. If BB is a Banach space with norm \|\cdot\|, or more generally a topological vector space, and F:UB:βfβF:U\to B:\beta\mapsto f_{\beta} is some BB-valued function, we define FF to be homeomorphic if (fβfβ0)/(ββ0)(f_{\beta}-f_{\beta_{0}})/(\beta-\beta_{0}) has a limit as ββ0\beta\to\beta_{0} for β0U\beta_{0}\in U.

If B1,B2B_{1},B_{2} are Banach spaces, let L(B1,B2)L(B_{1},B_{2}) be the set of bounded operators equipped with the usual operator norm, then a map 𝒦:UL(B1,B2):βKβ\mathcal{K}:U\mapsto L(B_{1},B_{2}):\beta\mapsto K_{\beta} 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 βLβ\beta\to L_{\beta} is a meromorphic function on (β)>(1α)/2\Re(\beta)>(1-\alpha)/2 for any α>0\alpha\in\mathbb{R}_{>0} with derivative

(dLβdβ)(f)(z)=n=0k(n+2β)dnfdzn(0)ζH(n+2β+1,z+1)n!2n=1ln(n+z)(n+z)2β(f)(1n+z).\left(\frac{dL_{\beta}}{d\beta}\right)(f)(z)=-\sum_{n=0}^{k}(n+2\beta)\frac{d^{n}f}{dz^{n}}(0)\frac{\zeta_{H}(n+2\beta+1,z+1)}{n!}-2\sum_{n=1}^{\infty}\frac{\ln(n+z)}{(n+z)^{2\beta}}(f^{\star})\left(\frac{1}{n+z}\right).

3 Uniformly Bounded Essential Spectrum

By (7), we can estimate the essential spectral radius by finding for each ll\in\mathbb{N} a compact operator TT^{\prime} for which LβlT\|L_{\beta}^{l}-T^{\prime}\| is small. A first naive approach would be to consider the operator PP which sends ff to a piecwise linear approximation that interpolates ff on the points 0,1N,2N,,10,\frac{1}{N},\frac{2}{N},\cdots,1 and considering LβlLβlP\|L_{\beta}^{l}-L_{\beta}^{l}\circ P\| as NN\to\infty. For (β)>0\Re(\beta)>0, we indeed have by Lemma 2.1 that for any ε>0\varepsilon>0 there exists some NN for which |(LβLβP)(f)|α(x)𝔏(β)+α((1+ε)|fP(f)|α)(x)+ε\lvert(L_{\beta}-L_{\beta}\circ P)(f)\rvert_{\alpha}(x)\leq\mathfrak{L}_{\Re(\beta)+\alpha}((1+\varepsilon)|f-P(f)|_{\alpha})(x)+\varepsilon. However, we do not have sufficient control over the \|\cdot\|_{\infty}-norm. We remedy this by considering a countable interpolation which ensures (LβlP)f(L_{\beta}^{l}\circ P)f remains a (nonlinear) interpolation of LβlfL_{\beta}^{l}f.

To this end, let 𝒫0,N\mathcal{P}_{0,N} be the set of points {0,1N,2N,,k1N}\left\{0,\frac{1}{N},\frac{2}{N},\cdots,\frac{k-1}{N}\right\}. For ll\in\mathbb{N}, we let

𝒫l,N:=n1,,nlψn1,,nl(𝒫0,N), and 𝒫l,N~={1}l=0l𝒫l,N,\mathcal{P}_{l,N}:=\bigcup_{n_{1},\ldots,n_{l}\in\mathbb{N}}\psi_{n_{1},\ldots,n_{l}}(\mathcal{P}_{0,N})\text{, and }\widetilde{\mathcal{P}_{l,N}}=\{1\}\cup\bigcup_{l^{\prime}=0}^{l}\mathcal{P}_{l^{\prime},N},

i.e. 𝒫l,N\mathcal{P}_{l^{\prime},N} is the set of all points of the form [n1,,nl+k/N][n_{1},\ldots,n_{l^{\prime}}+k/N] and 𝒫l,N~\widetilde{\mathcal{P}_{l,N}} is the union of the set of all such points for lll^{\prime}\leq l with 𝒫0,N\mathcal{P}_{0,N}.

We remark that the set 𝒫l,N~\widetilde{\mathcal{P}_{l,N}} is closed. Indeed, the set of accumulation points are the set of all fractions that can be written as [n1,,nl][n_{1},\ldots,n_{l^{\prime}}] for l<ll^{\prime}<l and these are in 𝒫l,N~\widetilde{\mathcal{P}_{l,N}} by definition. Hence its complement is a countable disjoint collection of open intervals and we can define the following.

Definition 3.1.

For α(0,1)\alpha\in(0,1), we define the operator Pl,N:Cα([0,1])Cα([0,1])P_{l,N}:C^{\alpha}([0,1])\to C^{\alpha}([0,1]) by letting Pl,N(f)(x)=f(x)P_{l,N}(f)(x)=f(x) for x𝒫l,N~x\in\widetilde{\mathcal{P}_{l,N}}. If x𝒫l,N~x\neq\widetilde{\mathcal{P}_{l,N}}, let aa and bb the respective largest and smallest element of 𝒫l,N~\widetilde{\mathcal{P}_{l,N}} such that a<x<ba<x<b and interpolate the function ff between the two points to obtain

Pl,N(f)(x)=bxbaf(a)+xabaf(b).P_{l,N}(f)(x)=\frac{b-x}{b-a}f(a)+\frac{x-a}{b-a}f(b).

We prove it is well-defined in Lemma A.2 in the appendix. It is fairly easy to show that Pl,NP_{l,N} is not a compact operator. However, we shall show that βlPl,N\mathcal{L}_{\beta}^{l}\circ P_{l,N} is compact. As an intermediate step, we prove the following proposition.

Proposition 3.1.

Let α(0,1)\alpha\in(0,1) and let ll\in\mathbb{N}. Let K>0K>0. For each kk\in\mathbb{N}, let kCα([0,1])\mathcal{F}_{k}\subset C^{\alpha}([0,1]) be a family of functions where all fkf\in\mathcal{F}_{k} satisfy fK\|f\|\leq K. Assume furthermore that for all n1,n2,,nln_{1},n_{2},\ldots,n_{l}\in\mathbb{N} we have that

limksupfkfψn1,,nl=0.\lim_{k\to\infty}\sup_{f\in\mathcal{F}_{k}}\|f\circ\psi_{n_{1},\ldots,n_{l}}\|=0. (11)

Then

limksupfkLβ(f)ψn2,,nl=0\lim_{k\to\infty}\sup_{f\in\mathcal{F}_{k}}\|L_{\beta}(f)\circ\psi_{n_{2},\ldots,n_{l}}\|=0

for all n2,,nln_{2},\ldots,n_{l} if l>1l>1 and limksupfkLβ(f)=0\lim_{k\to\infty}\sup_{f\in\mathcal{F}_{k}}\|L_{\beta}(f)\|=0 if l=1l=1.

Applying the above proposition ll times, we obtain the following corollary.

Corollary 3.1.

Let ll\in\mathbb{N} and let k\mathcal{F}_{k} be a collection of subsets of Cα([0,1])C^{\alpha}([0,1]) as in Proposition 3.1. Then limksupfkLβl(f)=0\lim_{k\to\infty}\sup_{f\in\mathcal{F}_{k}}\|L^{l}_{\beta}(f)\|=0.

The corollary is thus a way of converting a statement of the convergence of the family k\mathcal{F}_{k} of functions on individual sets of the form {[n1,,nl+x]:x[0,1]}\{[n_{1},\ldots,n_{l}+x]:x\in[0,1]\} to convergence of Lβl(k)L_{\beta}^{l}(\mathcal{F}_{k}) with respect to the \|\cdot\|-norm.

Proof of Proposition 3.1..

Our first claim is that (11) implies that limnsupfnf=0\lim_{n\to\infty}\sup_{f\in\mathcal{F}_{n}}\|f\|_{\infty}=0. Let ε>0\varepsilon>0. Let 𝒥l\mathcal{J}\subset\mathbb{N}^{l} be a finite collection of digits such that the set

A:=(n1,,nl)Jn1,,nlA:=\bigcup_{(n_{1},\ldots,n_{l})\in J}\langle n_{1},\ldots,n_{l}\rangle

satisfies d(x,A)εd(x,A)\leq\varepsilon for all x[0,1]x\in[0,1]. By (11) , there exists some NN\in\mathbb{N} for which |f(y)|ε\left\lvert f(y)\right\rvert\leq\varepsilon for all yAy\in A and fkf\in\mathcal{F}_{k} with kNk\geq N. For any x[0,1]x\in[0,1], there exists a yAy\in A with d(y,x)2εd(y,x)\leq 2\varepsilon. Hence by uniform boundedness of the Hölder norm,

f(x)ε+(2ε)αK.f(x)\leq\varepsilon+(2\varepsilon)^{\alpha}K.

Letting ε0\varepsilon\to 0, we obtain the claim. It suffices therefore to show that limksupfkfψn2,,nlα=0\lim_{k\to\infty}\sup_{f\in\mathcal{F}_{k}}\|f\circ\psi_{n_{2},\ldots,n_{l}}\|_{\alpha}=0. Assume for the moment that l>1l>1.

We need uniform estimates on the Hölder seminorm of the function.

Lβ(f)([n2,n3,,nl+x])=f(0)ζH(2β,1+[n2,n3,,nl+x])+n1=1[n1,n2,,nl+x]2βf([n1,n2,,nl+x]),\begin{split}L_{\beta}(f)([n_{2},n_{3},\ldots,n_{l}+x])&=f(0)\zeta_{H}(2\beta,1+[n_{2},n_{3},\ldots,n_{l}+x])\\ &+\sum_{n_{1}=1}^{\infty}[n_{1},n_{2},\ldots,n_{l}+x]^{2\beta}f^{\star}\left([n_{1},n_{2},\ldots,n_{l}+x]\right),\end{split} (12)

where f=ff(0)f^{\star}=f-f(0). Let ε>0\varepsilon>0 and note that there exists some N1N_{1}\in\mathbb{N} depending only on K,βK,\beta and α\alpha such that

supxy|f(0)||ζH(2β,1+[n2,n3,,nl+y])ζH(2β,1+[n2,n3,,nl+x])||xy|αε\sup_{x\neq y}\left\lvert f(0)\right\rvert\frac{\left\lvert\zeta_{H}(2\beta,1+[n_{2},n_{3},\ldots,n_{l}+y])-\zeta_{H}(2\beta,1+[n_{2},n_{3},\ldots,n_{l}+x])\right\rvert}{\left\lvert x-y\right\rvert^{\alpha}}\leq\varepsilon

for all fkf\in\mathcal{F}_{k} with kN1k\geq N_{1}, since ζH\zeta_{H} is analytic and supfk|f(0)|0\sup_{f\in\mathcal{F}_{k}}\left\lvert f(0)\right\rvert\to 0 as kk\to\infty. We now estimate the contribution to the Hölder seminorm from the second term in (12). For x,y[0,1]x,y\in[0,1], denote x^=[n2,,nl+x]\hat{x}=[n_{2},\ldots,n_{l}+x] and y^=[n2,,nl+y]\hat{y}=[n_{2},\ldots,n_{l}+y]. Let MM\in\mathbb{N} such that for all fCα([0,1])f\in C^{\alpha}([0,1]) with fK\|f\|\leq K:

n=M|f(1n+x^)(n+x^)2βf(1n+y^)(n+y^)2β|ε|x^y^|α\sum_{n=M}^{\infty}\left\lvert\frac{f\left(\frac{1}{n+\hat{x}}\right)}{(n+\hat{x})^{2\beta}}-\frac{f\left(\frac{1}{n+\hat{y}}\right)}{(n+\hat{y})^{2\beta}}\right\rvert\leq\varepsilon\left\lvert\hat{x}-\hat{y}\right\rvert^{\alpha}

for all xyx\neq y. The existence of such an MM 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

supfkn=1M1|f(1n+x^)(n+x^)2βf(1n+y^)(n+y^)2β|ε|x^y^|α\sup_{f\in\mathcal{F}_{k}}\sum_{n=1}^{M-1}\left\lvert\frac{f\left(\frac{1}{n+\hat{x}}\right)}{(n+\hat{x})^{2\beta}}-\frac{f\left(\frac{1}{n+\hat{y}}\right)}{(n+\hat{y})^{2\beta}}\right\rvert\leq\varepsilon\left\lvert\hat{x}-\hat{y}\right\rvert^{\alpha}

for all fkf\in\mathcal{F}_{k} with kk greater than some N2N_{2}\in\mathbb{N} and x,y[0,1]x,y\in[0,1] with xyx\neq y. We obtain by (12) and the above three equations that

|L(f)([n2,n3,,nl+y])L(f)([n2,n3,,nl+x])||[n2,n3,,nl+y][n2,n3,,nl+x]|α3ε\frac{\left\lvert L(f)([n_{2},n_{3},\ldots,n_{l}+y])-L(f)([n_{2},n_{3},\ldots,n_{l}+x])\right\rvert}{\left\lvert[n_{2},n_{3},\ldots,n_{l}+y]-[n_{2},n_{3},\ldots,n_{l}+x]\right\rvert^{\alpha}}\leq 3\varepsilon

for all fkf\in\mathcal{F}_{k} with k>max{N1,N2}k>\max\{N_{1},N_{2}\}. The lemma for l>1l>1 follows from letting ε0\varepsilon\to 0. For the case l=1l=1, replace [n2,n3,,nl+x][n_{2},n_{3},\ldots,n_{l}+x] , [n2,n3,,nl+y][n_{2},n_{3},\ldots,n_{l}+y] and [n1,n2,,nl+x][n_{1},n_{2},\ldots,n_{l}+x] with x,yx,y and 1/(n1+x)1/(n_{1}+x) respectively.

Proposition 3.2.

The operator LβlPl,NL_{\beta}^{l}\circ P_{l,N} is compact for each ll\in\mathbb{N} and NN\in\mathbb{N}.

Proof.

Let BCαB_{C^{\alpha}} denote the set of functions with f<1\|f\|<1. We shall prove (LβlPl,N)(BCα)(L_{\beta}^{l}\circ P_{l,N})(B_{C^{\alpha}}) is precompact by showing that any sequence in the set has a Cauchy subsequence.

Let gng_{n} be a sequence of functions in (LβlPl,N)(BCα)(L_{\beta}^{l}\circ P_{l,N})(B_{C^{\alpha}}) and let fnf_{n} be a sequence of functions in BCαB_{C^{\alpha}} with (LβlPl,N)(fn)=gn(L_{\beta}^{l}\circ P_{l,N})(f_{n})=g_{n} for all nn. For any x𝒫l,N~x\in\widetilde{\mathcal{P}_{l,N}} there exists a subsequence fmkf_{m_{k}} such that fmk(x)f_{m_{k}}(x) is a Cauchy sequence. By a diagonal argument, we may assume fmk(x)f_{m_{k}}(x) is Cauchy for all x𝒫l,N~x\in\widetilde{\mathcal{P}_{l,N}}.

For any n1,,nln_{1},\ldots,n_{l}\in\mathbb{N}, let II be the closed interval spanned by [n1,,nl][n_{1},\ldots,n_{l}] and [n1,,nl+1][n_{1},\ldots,n_{l}+1]. The intersection I𝒫l,N~I\cap\widetilde{\mathcal{P}_{l,N}} is finite. Indeed, it consists of the endpoints, the points [n1,n2,,nl+k/N][n_{1},n_{2},\ldots,n_{l}+k/N] with k{0,,N1}k\in\{0,\ldots,N-1\} and the points [n1,n2,,nl+p/N][n_{1},n_{2},\ldots,n_{l}^{\prime}+p/N] with l<ll^{\prime}<l and p/Nnl+1,,nlp/N\in\langle n_{l^{\prime}+1},\ldots,n_{l}\rangle.

Since the functions Pl,N(fmk)P_{l,N}(f_{m_{k}}) restricted to II are piecewise linear functions interpolated at a finite number of points xx for which fmk(x)f_{m_{k}}(x) is Cauchy, the functions Pl,N(fmk)|I\left.P_{l,N}(f_{m_{k}})\right\rvert_{I} form a Cauchy sequence in Cα(I)C^{\alpha}(I). Since ψn1,,nl\psi_{n_{1},\ldots,n_{l}} is analytic on a neighbourhood of [0,1][0,1], it follows that Pl,N(fmk)ψn1,,nlP_{l,N}(f_{m_{k}})\circ\psi_{n_{1},\ldots,n_{l}} is Cauchy on Cα([0,1])C^{\alpha}([0,1]). In particular, the families of functions defined by

k={Pl,N(fmr)Pl,N(fms):r,sk}\mathcal{F}_{k}=\{P_{l,N}(f_{m_{r}})-P_{l,N}(f_{m_{s}}):r,s\geq k\}

satisfy the conditions of Proposition 3.1. By Corollary 3.1, we thus obtain that

limksupr,sk(LβlPl,N)(fmr)(LβlPl,N)(fms)=0\lim_{k\to\infty}\sup_{r,s\geq k}\left\|(L_{\beta}^{l}\circ P_{l,N})(f_{m_{r}})-(L_{\beta}^{l}\circ P_{l,N})(f_{m_{s}})\right\|=0

which shows gmkg_{m_{k}} is a Cauchy subsequence. ∎

Let us now prove most of Theorem 1.1.

Theorem 3.1.

Let 0<α<10<\alpha<1. Then for all β\beta with (β)>(1α)/2\Re(\beta)>(1-\alpha)/2 the transfer operator LβL_{\beta} satisfies

ρe(Lβ)ρ(𝔏(β)+α).\rho_{e}(L_{\beta})\leq\rho(\mathfrak{L}_{\Re(\beta)+\alpha}).
Proof.

By Nussbaum’s formula and Proposition 3.2, it suffices to prove that there is some constant C′′>0C^{\prime\prime}>0 for which

lim supNLβlLβlPl,NC′′𝔏(β)+αl for all l.\limsup_{N\to\infty}\|L_{\beta}^{l}-L_{\beta}^{l}\circ P_{l,N}\|\leq C^{\prime\prime}\left\|\mathfrak{L}_{\Re(\beta)+\alpha}^{l}\right\|\text{ for all }l.

We show in Lemma A.2 in the appendix that there exists a constant E>E> for which Pl,NE\|P_{l,N}\|\leq E for all l,Nl,N.

For fixed ll, we first claim that

LβlLβlPl,N0 as n\|L_{\beta}^{l^{\prime}}-L_{\beta}^{l^{\prime}}\circ P_{l,N}\|_{\infty}\to 0\text{ as }n\to\infty

for all lll^{\prime}\leq l. Let fCα([0,1])f\in C^{\alpha}([0,1]). For notational simplicity, let gl,N:=fPl,Nfg_{l,N}:=f-P_{l,N}f, which by the aforementioned lemma satisfies gl,N(E+1)f\|g_{l,N}\|\leq(E+1)\|f\| for all l,Nl,N. Since gl,N(0)=0g_{l,N}(0)=0 we have that

Lβ(gl,N)(x)=n=11(n+x)2β(gl,N)(1n+x).L_{\beta}(g_{l,N})(x)=\sum_{n=1}^{\infty}\frac{1}{(n+x)^{2\beta}}(g_{l,N})\left(\frac{1}{n+x}\right).

It follows from the above formula that since gl,N(x)=0g_{l,N}(x)=0 for all x𝒫k,Nx\in\mathcal{P}_{k,N} with klk\leq l, the equality Lβ(gl,N)(x)=0L_{\beta}(g_{l,N})(x)=0 holds for all x𝒫l1,Nx\in\mathcal{P}_{l-1,N}. By repeating this argument, we obtain that Lβl(gl,N)(x)=0L_{\beta}^{l^{\prime}}(g_{l,N})(x)=0 for all x𝒫ll,N~x\in\widetilde{\mathcal{P}_{l-l^{\prime},N}}. By definition of the α\|\cdot\|_{\alpha}-seminorm, we obtain the inequality Lβl(gl,N)λαLβl(gl,N)α\|L_{\beta}^{l^{\prime}}(g_{l,N})\|_{\infty}\leq\lambda^{\alpha}\|L_{\beta}^{l^{\prime}}(g_{l,N})\|_{\alpha}, where λ\lambda is the maximal distance of a point xx to 𝒫ll,N~\widetilde{\mathcal{P}_{l-l^{\prime},N}}. We thus obtain

(LβlLβlPl,N)(f)=Lβl(gl,N)(E+1)λαLβlf.\|(L_{\beta}^{l^{\prime}}-L_{\beta}^{l^{\prime}}\circ P_{l,N})(f)\|_{\infty}=\|L_{\beta}^{l^{\prime}}(g_{l,N})\|_{\infty}\leq(E+1)\lambda^{\alpha}\|L_{\beta}^{l^{\prime}}\|\|f\|.

Letting NN\to\infty proves the claim.

To show the theorem use the claim for l=ll=l^{\prime} and Lemma 2.1 to obtain for any ε>0\varepsilon>0 that

lim supNLβlLβlPl,Nlim supNsupfCα([0,1])f1𝔏(β)+α((1+ε)|Lβl1(gl,N)|α).\limsup_{N\to\infty}\left\|L_{\beta}^{l}-L_{\beta}^{l}\circ P_{l,N}\right\|\leq\limsup_{N\to\infty}\sup_{\begin{subarray}{c}f\in C^{\alpha}([0,1])\\ \|f\|\leq 1\end{subarray}}\left\|\mathfrak{L}_{\Re(\beta)+\alpha}\left((1+\varepsilon)\left\lvert L_{\beta}^{l-1}(g_{l,N})\right\rvert_{\alpha}\right)\right\|_{\infty}.

Since ε\varepsilon is arbitrary, the above also holds for ε=0\varepsilon=0. Applying the claim for l=l1,l2,,0l^{\prime}=l-1,l-2,\ldots,0 we obtain once again by Lemma 2.1 that

lim supNLβlLβlPl,Nlim supNsupfCα([0,1])f1𝔏(β)+αl(|gl,N|α)(E+1)𝔏(β)+αl.\limsup_{N\to\infty}\left\|L_{\beta}^{l}-L_{\beta}^{l}\circ P_{l,N}\right\|\leq\limsup_{N\to\infty}\sup_{\begin{subarray}{c}f\in C^{\alpha}([0,1])\\ \|f\|\leq 1\end{subarray}}\left\|\mathfrak{L}_{\Re(\beta)+\alpha}^{l}\left(\left\lvert g_{l,N}\right\rvert_{\alpha}\right)\right\|_{\infty}\leq(E+1)\left\|\mathfrak{L}_{\Re(\beta)+\alpha}^{l}\right\|.

Letting C′′=E+1C^{\prime\prime}=E+1 proves the theorem.

Remark 3.1.

By using the spectral radius formula ρ(𝔏(β)+α)=lim supl𝔏(β)+αl1/l\rho(\mathfrak{L}_{\Re(\beta)+\alpha})=\limsup_{l\to\infty}\|\mathfrak{L}_{\Re(\beta)+\alpha}^{l}\|^{1/l} and bounding every function fL([0,1])f\in L^{\infty}([0,1]) by a constant, we obtain that ρ(𝔏(β)+α)ρ((β)+α)\rho(\mathfrak{L}_{\Re(\beta)+\alpha})\leq\rho(\mathcal{L}_{\Re(\beta)+\alpha}) and hence that

ρ(𝔏(β)+α)=λ1((β)+α),\rho(\mathfrak{L}_{\Re(\beta)+\alpha})=\lambda_{1}(\Re(\beta)+\alpha),

where λ1(t)\lambda_{1}(t) for t>1/2t>1/2 is the eigenvalue of maximum modulus of t\mathcal{L}_{t}. The eigenvalue is guaranteed to be simple and positive by the Perron-Frobenius theorem. In fact λ1(t)\lambda_{1}(t) is a real-analytic and decreasing function of tt with λ(1)=1\lambda(1)=1, see [15, 9].

Theorem 1.1 for α(0,1)\alpha\in(0,1) then follows from Remark 3.1 and from the following corollary of Theorem 3.1.

Corollary 3.2.

Let λ\lambda\in\mathbb{C} satisfy |λ|>ρ(𝔏(β)+α)\left\lvert\lambda\right\rvert>\rho(\mathfrak{L}_{\Re(\beta)+\alpha}). Then λ\lambda is an eigenvalue of LβL_{\beta} if and only if it is an eigenvalue of β\mathcal{L}_{\beta}. In this case the respective eigenspaces EλαE_{\lambda}^{\alpha} and EλωE_{\lambda}^{\omega} 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 k:=αk:=\lfloor\alpha\rfloor and let η=αα\eta=\alpha-\lfloor\alpha\rfloor. We first assume η>0\eta>0 and adapt a standard trick (see e.g. [6] III) to reduce to the problem on Cη([0,1])C^{\eta}([0,1]). We note that for (β)>>1\Re(\beta)>>1 we obtain by first applying the product rule for higher derivatives and then the chain rule that there exists coefficients crc_{r} with c0=1c_{0}=1 for which

dkLβ(f)dzk(z)=r=0kcrLβ+r/2+(kr)(dkrfdzkr)(z).\begin{split}\frac{d^{k}L_{\beta}(f)}{dz^{k}}(z)=\sum_{r=0}^{k}c_{r}L_{\beta+r/2+(k-r)}\left(\frac{d^{k-r}f}{dz^{k-r}}\right)(z).\end{split}

This holds using analytic continuation to (β)>(1k)/2\Re(\beta)>(1-k)/2. Recall that the any composition of operators is compact if one of them is compact and that the operator Cα([0,1])Cη([0,1])C^{\alpha}([0,1])\to C^{\eta}([0,1]) which sends ff to dkrf/dzkrd^{k-r}f/dz^{k-r} is compact for r>0r>0. Hence we can write

dkLβ(f)dzk=Lβ+k(dkLβ(f)dzk)+𝒦(f),\frac{d^{k}L_{\beta}(f)}{dz^{k}}=L_{\beta+k}\left(\frac{d^{k}L_{\beta}(f)}{dz^{k}}\right)+\mathcal{K}(f),

where 𝒦\mathcal{K} is some compact operator from Cα([0,1])Cη([0,1])C^{\alpha}([0,1])\to C^{\eta}([0,1]). Since the essential spectral radius does not depend on compact perturbations, we have that the essential spectral radius of LβL_{\beta} as an operator on Cα([0,1])C^{\alpha}([0,1]) is the same as the essential spectral radius of Lβ+kL_{\beta+k} as an operator on Cη([0,1])C^{\eta}([0,1]). This follows from (7), but is perhaps easier to see using the Banach space isomorphism Cα([0,1])×Cη([0,1]):f((dkfdzr(0))r=0k1,dkfdzk)C^{\alpha}([0,1])\to\mathbb{R}\times C^{\eta}([0,1]):f\mapsto\left(\left(\frac{d^{k}f}{dz^{r}}(0)\right)_{r=0}^{k-1},\frac{d^{k}f}{dz^{k}}\right). The case η=0\eta=0 follows mutatis mutandis, replacing Cη([0,1])C^{\eta}([0,1]) with C([0,1])C([0,1]). Hence Theorem 1.1 for α>1\alpha>1 follows from the case α(0,1)\alpha\in(0,1). ∎

Remark 3.2.

The zeroes of Z(s)Z(s) on the line s=1/2+its=1/2+it are those for which there are eigenvalues of Δ\Delta called ‘Maass cusp forms’ with eigenvalue s(1s)s(1-s). Furthermore the multiplicity of the zeroes and the eigenspaces of Δ\Delta match. The other zeroes of Z(s)Z(s) are a simple zero at s=1s=1 and at the values of ss for which 2s2s is a nontrivial zero of the Riemann zeta function ζ\zeta. Since ρ(a)<1\rho(\mathcal{L}_{a})<1 when a>1a>1, we see that the eigenspaces of β\mathcal{L}_{\beta} and LβL_{\beta} for the eigenvalues λ=±1\lambda=\pm 1 are identical when 1>α>1(β)1>\alpha>1-\Re(\beta). In particular, the eigenspaces associated to the Maass cusp forms are preserved when α>1/2\alpha>1/2, and the eigenspaces associated to the nontrivial zeroes of the Riemann zeta function111In fact, the zeroes of Z(s)Z(s) associated with the nontrivial zeroes of ζ\zeta always occur due to 11 being an eigenvalue of β\mathcal{L}_{\beta}. to the right of the critical line are preserved when α=3/4\alpha=3/4.

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 LβL_{\beta} acting on spaces other than A(D)A_{\infty}(D) or Cα([0,1])C^{\alpha}([0,1]). For example, it is clear that the estimates on the essential spectrum still hold for LβL_{\beta} acting on Cα([a,M])C^{\alpha}([a,M]) with 0<a10<a\leq 1 and M(1+a)1M\geq(1+a)^{-1}. As we will remark upon Section 4, eigenfunctions of β\mathcal{L}_{\beta} extend to functions on (1,)\mathbb{C}\setminus(-1,\infty), so the point spectrum beyond the estimate of Theorem 1.1 does not depend on a,Ma,M. In fact by letting a0a\to 0 and MM\to\infty, we see that a natural space for LβL_{\beta} is the space Cα((1,))C^{\alpha}((-1,\infty)), which we define to be the Fréchet space of functions ff such that on every compact K(1,)K\subset(-1,\infty) we have that ff is α\lfloor\alpha\rfloor-times differentiable such that the α\lfloor\alpha\rfloor-th derivative is continuous if α=α\alpha=\lfloor\alpha\rfloor and is (αα)(\alpha-\lfloor\alpha\rfloor)-Hölder continuous if not.

4 Relation with the three-term functional equation.

In this section, we consider the following functional equation.

λf(z)λf(z+1)=(1z+1)2βf(1z+1),\lambda f(z)-\lambda f(z+1)=\left(\frac{1}{z+1}\right)^{2\beta}f\left(\frac{1}{z+1}\right), (13)

where λ\{0}\lambda\in\mathbb{C}\backslash\{0\}. This equation was first introduced (in slightly modified form) for λ=1\lambda=1 in [14] and for λ=±1\lambda=\pm 1 in [13]. The formulation in this document can be found in [4]. In particular, we consider this functional equation for functions on (1,)(-1,\infty) of varying regularity and for holomorphic functions on \(,1]\mathbb{C}\backslash(-\infty,1].

It is fairly easy to prove that any eigenfunction of β\mathcal{L}_{\beta} with eigenvalue λ0\lambda\neq 0 satisfies (13). Indeed, this follows from the identity ζ(s,z)ζ(s,z+1)=zs\zeta(s,z)-\zeta(s,z+1)=z^{-s}, see e.g. Proposition 7 in [4]. Furthermore, from the identity

f=λ1(n=0kdnfdzn(0)ζH(n+2β,z+1)n!+n=11(n+z)2β(f)(1n+z),),f=\lambda^{-1}\left(\sum_{n=0}^{k}\frac{d^{n}f}{dz^{n}}(0)\frac{\zeta_{H}(n+2\beta,z+1)}{n!}+\sum_{n=1}^{\infty}\frac{1}{(n+z)^{2\beta}}(f^{\star})\left(\frac{1}{n+z}\right),\right),

we can use the contraction properties of the inverse branches ψn\psi_{n} inductively to extend ff to a function which is holomorphic on (,1]\mathbb{C}\setminus(-\infty,1].

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 ff. Let us first describe the asymptotic behaviour of a generic solution in the non-analytic case (e.g. at Hölder regularity).

Define Cα(/)C^{\alpha}(\mathbb{R}/\mathbb{Z}) be the space of all α\lfloor\alpha\rfloor-times continuously differentiable functions on the torus with (αα)(\alpha-\lfloor\alpha\rfloor)-Hölder derivative if α\alpha is not an integer. We identify these functions with 11-periodic functions on Cα((1,))C^{\alpha}((-1,\infty)) by setting Q(z)=Q(zmod1)Q(z)=Q(z\mod 1) for z(1,)z\in(-1,\infty).

Proposition 4.1.

Let fCα((1,))f\in C^{\alpha}((-1,\infty)) solve (13) for some λ0\lambda\neq 0. Assume furthermore that (β)>(1α)/2\Re(\beta)>(1-\alpha)/2 and β{1/2,0,1/2,3/2,}\beta\notin\{1/2,0,-1/2,-3/2,\ldots\}. Then there exist a function QCα((1,))Q\in C^{\alpha}((-1,\infty)) and coefficients C0,,Cα,C0,CαC_{0},\ldots,C_{\lfloor\alpha\rfloor},C_{0}^{*},\ldots C_{\lfloor\alpha\rfloor}^{*} depending only on λ,β\lambda,\beta and the derivatives {f(0),dfdz(0),dαfdzα(0)}\left\{f(0),\frac{df}{dz}(0),\ldots\frac{d^{\lfloor\alpha\rfloor}f}{dz^{\lfloor\alpha\rfloor}}(0)\right\} such that

f(z)=Q(z)+n=0αCnz12βn+O(x12(β)α) as zf(z)=Q(z)+\sum_{n=0}^{\lfloor\alpha\rfloor}C_{n}z^{{1-2\beta-n}}+O(x^{1-2\Re(\beta)-\alpha})\text{ as }z\to\infty (14)

and

f(1z)=1λz2βQ(1/z)+n=0αCnzn1+O(xα1) as z0,f(1-z)=\frac{1}{\lambda z^{2\beta}}Q(1/z)+\sum_{n=0}^{\lfloor\alpha\rfloor}C_{n}^{*}z^{{n-1}}+O(x^{\alpha-1})\text{ as }z\to 0, (15)

where we may replace the O(x12(β)α),O(xα1)O(x^{1-2\Re(\beta)-\alpha}),O(x^{\alpha-1}) with o(x12(β)α),o(xα1)o(x^{1-2\Re(\beta)-\alpha}),o(x^{\alpha-1}) respectively if α=α\alpha=\lfloor\alpha\rfloor.

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 λ±1\lambda\neq\pm 1.

Define the function Q:=fλ1Lβ(f)Q:=f-\lambda^{-1}L_{\beta}(f), where we extend LβL_{\beta} to an operator on Cα((1,))C^{\alpha}((-1,\infty)) as in Remark 3.3. For (β)>1\Re(\beta)>1 it is clear that

Q(z)Q(z+1)=f(z)f(z+1)λ11(1+z)2βf(11+z)=0Q(z)-Q(z+1)=f(z)-f(z+1)-\lambda^{-1}\frac{1}{\left(1+z\right)^{2\beta}}f\left(\frac{1}{1+z}\right)=0

and this formula continues to hold for all (β)>(1α)/2\Re(\beta)>(1-\alpha)/2 (with β1/2\beta\neq 1/2) by analytic continuation. It is also clear from the definition that QCα(/)Q\in C^{\alpha}(\mathbb{R}/\mathbb{Z}).

Assume α=k+η\alpha=k+\eta with (k,η)×[0,1)(k,\eta)\in\mathbb{N}\times[0,1) We now consider the asymptotic behaviour of Lβ(f)(z)L_{\beta}(f)(z). We note that if η>0\eta>0 and z[0,1]z\in[0,1],

|f(z)n=0kdnfdzn(0)znn!|dkfdzkηzαk!,\left\lvert f(z)-\sum_{n=0}^{k}\frac{d^{n}f}{dz^{n}}(0)\frac{z^{n}}{n!}\right\rvert\leq\left\|\frac{d^{k}f}{dz^{k}}\right\|_{\eta}\frac{z^{\alpha}}{k!}, (16)

where gη\|g\|_{\eta} is the η\eta-Hölder coefficient of gg on [0,1][0,1]. Hence we obtain by an application of the integral test that if f(z)=f(z)n=0kdnfdzn(0)znn!f^{\star}(z)=f(z)-\sum_{n=0}^{k}\frac{d^{n}f}{dz^{n}}(0)\frac{z^{n}}{n!},

n=11(n+z)2β(f)(1n+z)=O(z12(β)α)\sum_{n=1}^{\infty}\frac{1}{(n+z)^{2\beta}}(f^{\star})\left(\frac{1}{n+z}\right)=O(z^{1-2\Re(\beta)-\alpha})

as zz\to\infty with the implied constant easily computable and depending on dkfdzkη,β\left\|\frac{d^{k}f}{dz^{k}}\right\|_{\eta},\beta and α\alpha. We therefore obtain by (4) that

Lβ(f)(z)=n=0kdnfdzn(0)ζH(n+2β,z+1)n!+O(z12(β)α) as zL_{\beta}(f)(z)=\sum_{n=0}^{k}\frac{d^{n}f}{dz^{n}}(0)\frac{\zeta_{H}(n+2\beta,z+1)}{n!}+O(z^{1-2\Re(\beta)-\alpha})\text{ as }z\to\infty (17)

We obtain (14) for η>0\eta>0 by writing f=Q+λ1Lβ(f)f=Q+\lambda^{-1}L_{\beta}(f) using the identity ζH(s,z)=ζH(s,z+1)+zs\zeta_{H}(s,z)=\zeta_{H}(s,z+1)+z^{-s} and the well-known expansion (see [18] §25.11)

ζH(s,z)z1ss1+12zs+k=1B2k2k!Γ(s+2k1)Γ(s)z1s2k as z,\zeta_{H}(s,z)\sim\frac{z^{1-s}}{s-1}+\frac{1}{2}z^{-s}+\sum_{k=1}^{\infty}\frac{B_{2k}}{2k!}\frac{\Gamma(s+2k-1)}{\Gamma(s)}z^{1-s-2k}\text{ as }z\to\infty, (18)

where B2kB_{2k} are the Bernoulli numbers.

For η=0\eta=0, we have that the left hand side of (16) is o(1)o(1) as z0z\to 0 and we follow our previous argument, mutatis mutandis. The asymptotic behavior of f(1z)f(1-z) for small zz follows from the identity f(1z)=λ1z2βf(1/z)+f(z)f(1-z)=\frac{\lambda^{-1}}{z^{2\beta}}f(1/z)+f(z), using the asymptotics (14) to obtain the expansion for f(1/z)f(1/z) as z0z\to 0 and the expansion f(z)=n=0αdnfdzn(0)zkk!+O(zα)f(z)=\sum_{n=0}^{\lfloor\alpha\rfloor}\frac{d^{n}f}{dz^{n}}(0)\frac{z^{k}}{k!}+O(z^{\alpha}).

We see by Proposition 4.14.1 that a generic solution of (13) in Cα((1,))C^{\alpha}((-1,\infty)) satisfies f=O(z12(β))f=O(z^{1-2\Re(\beta)}) if (β)>(1α)/2\Re(\beta)>(1-\alpha)/2. On the other hand, we see that the condition Q=0Q=0 in (14) or in (15) is equivalent to the condition Lβf=λfL_{\beta}f=\lambda f, which by Theorem 1.1 implies ff extends to a holomorphic function in (,1]\mathbb{C}\setminus(-\infty,-1] if |λ|ρ((β)+α)\left\lvert\lambda\right\rvert\leq\rho(\mathcal{L}_{\Re(\beta)+\alpha}). 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 (β)=1/2\Re(\beta)=1/2. Suppose f:(1,)f:(-1,\infty)\to\mathbb{C} is some solution of (13) for |λ|=1\left\lvert\lambda\right\rvert=1. Assume furthermore it satisfies the mild growth condition f(z)=C0x12β+o(1)f(z)=C_{0}x^{1-2\beta}+o(1) as zz\to\infty. Then if ff is 1/2+ε1/2+\varepsilon-Hölder for some ε>0\varepsilon>0, it is automatically holomorphically extendable to the cut plane (,1]\mathbb{C}\setminus(-\infty,-1].

This corollary immediately follows from Theorem 1.1 and the fact that ρ(a)<1\rho(\mathcal{L}_{a})<1 if a>1a>1. We remark that when (β)=1/2\Re(\beta)=1/2 and λ=1\lambda=1 or 1-1, the space of eigenvalues of β\mathcal{L}_{\beta} is in bijection with the space of even, respectively uneven Maass cusp forms on the modular surface with eigenvalue β(1β)\beta(1-\beta) and we have that C0=0C_{0}=0, i.e. f(z)=o(1)f(z)=o(1) as zz\to\infty, see [14, 13].

Proposition 4.1 showed that a solution of the three term equation has the associated periodic function (1Lβ)f(1-L_{\beta})f. By applying the inverse, we obtain the following proposition.

Proposition 4.2.

Let α>0\alpha>0, λ{0}\lambda\in\mathbb{C}\setminus\{0\} and let QCα(/)Q\in C^{\alpha}(\mathbb{R}/\mathbb{Z}). For all β\beta for which (β)>(1α)/2\Re(\beta)>(1-\alpha)/2 and

r((β)+α)<|λ|,r(\mathcal{L}_{\Re(\beta)+\alpha})<\left\lvert\lambda\right\rvert,

the function

f=(1λ1Lβ)1Q,f=(1-\lambda^{-1}L_{\beta})^{-1}Q,

is a meromorphic expression over all β\beta satisfying the conditions in this proposition. Furthermore ff satisfies (13), (14) and (15).

Proof.

We interpret (λLβ)1(\lambda-L_{\beta})^{-1} in the sense of Remark 3.3. More specifically, for LβL_{\beta} acting on Cr([a,(1a)1])C^{r}([a,(1-a)^{-1}]), the expression (1λ1Lβ)1(Q|[a,(1a)1])(1-\lambda^{-1}L_{\beta})^{-1}\left(\left.Q\right\rvert_{[a,(1-a)^{-1}]}\right) is meromorphic and satisfies

(1λ1Lβ)1(Q|[a,(1a)1])(z)=n=0λnLβn(Q|[a,(1a)1])(z).=Q(z)+n=1λna1,,anQ([a1,,an+z])Πk=1n[ak,,an+z]2β\begin{split}&(1-\lambda^{-1}L_{\beta})^{-1}\left(\left.Q\right\rvert_{[a,(1-a)^{-1}]}\right)(z)=\sum_{n=0}^{\infty}\lambda^{-n}L_{\beta}^{n}\left(\left.Q\right\rvert_{[a,(1-a)^{-1}]}\right)(z).\\ &=Q(z)+\sum_{n=1}^{\infty}\lambda^{-n}\sum_{a_{1},\ldots,a_{n}\in\mathbb{N}}Q([a_{1},\ldots,a_{n}+z])\Pi_{k=1}^{n}[a_{k},\ldots,a_{n}+z]^{2\beta}\end{split} (19)

if (β)\Re(\beta) is large enough. By Theorem 1.1, we see that (1λ1Lβ)1(1-\lambda^{-1}L_{\beta})^{-1} is a meromorphic operator-valued function for all a1a\leq 1. Hence the functions (1λ1Lβ)1(Q|[a,(1a)1])(1-\lambda^{-1}L_{\beta})^{-1}\left(\left.Q\right\rvert_{[a,(1-a)^{-1}]}\right) are meromorphic functions over the half-plane consisting of all β\beta 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 aa on the intersection of their domains. Hence we may define (1λ1Lβ)1Q(z)(1-\lambda^{-1}L_{\beta})^{-1}Q(z) for any z(1,)z\in(-1,\infty) by choosing aa in (19) small enough.

We may thus define

f:=(1λ1Lβ)1Q(z)=Q(z)+n=1λna1,,anQ([a1,,an+z])Πk=1n[ak,,an+z]2βf:=(1-\lambda^{-1}L_{\beta})^{-1}Q(z)=Q(z)+\sum_{n=1}^{\infty}\lambda^{-n}\sum_{a_{1},\ldots,a_{n}\in\mathbb{N}}Q([a_{1},\ldots,a_{n}+z])\Pi_{k=1}^{n}[a_{k},\ldots,a_{n}+z]^{2\beta}

For (β)\Re(\beta) large enough and use (19) to obtain the analytic continuation. It is clear from the above formula that ff satisfies (13) for large (β)\Re(\beta) and hence by analytic continuation for all β\beta 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.

Let β{1/2,0,1/2,},λ0\beta\notin\{1/2,0,-1/2,\ldots\},\lambda\neq 0 and let FEλ,β,α\operatorname{FE}^{\lambda,\beta,\alpha} denote the set of solutions of (13) belonging to Cα((1,))C^{\alpha}((-1,\infty)). Assume furthermore that (β)>(1α)/2\Re(\beta)>(1-\alpha)/2 and that ρ((β)+α)<λ\rho(\mathcal{L}_{\Re(\beta)+\alpha})<\lambda. Then the map

𝒜λ,β,α:FEλ,β,αCα(/)\mathcal{A}^{\lambda,\beta,\alpha}:\operatorname{FE}^{\lambda,\beta,\alpha}\to C^{\alpha}(\mathbb{R}/\mathbb{Z})

which sends fFEλ,β,αf\in\operatorname{FE}^{\lambda,\beta,\alpha} to its associated periodic function QQ defined in Proposition 4.1 is bijective if and only if λσ(β)\lambda\neq\sigma(\mathcal{L}_{\beta}).

Proof.

The map 𝒜λ,β,α\mathcal{A}^{\lambda,\beta,\alpha} is explicitly given by the restriction of f(1λ1Lβ)ff\mapsto(1-\lambda^{-1}L_{\beta})f to FEλ,β,α\operatorname{FE}^{\lambda,\beta,\alpha}, so theorem follows immediately from Proposition 4.1 and Proposition 4.2. ∎

Remark 4.1.

Proposition 4.2 does not give us a construction of solutions of (13) when λ\lambda is an eigenvalue of LβL_{\beta}. We remark that in the case λ=±1\lambda=\pm 1, explicit examples are constructed in [13], showing that the spaces FE±1,β,α\operatorname{FE}^{\pm 1,\beta,\alpha} are still uncountable-dimensional. For example, the function zQ(z)+(z+1)2β1Q(1/z)z\mapsto Q(z)+(z+1)^{2\beta-1}Q(-1/z), where Q(z)=Q(z)Q(z)=\mp Q(-z) and is 11-periodic. These constructions seem to rely on an additional symmetry for solutions of (13) with λ=±1\lambda=\pm 1. We conjecture these might not necessarily exist for other values of λσ(Lβ)\lambda\in\sigma(L_{\beta}).

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 0<α<10<\alpha<1. Let a<bc<da<b\leq c<d and let g:{a,b,c,d}g:\{a,b,c,d\}\to\mathbb{R} be a function. Then there exists a constant E>0E>0 independent of a,b,c,da,b,c,d for which

|g(d)g(a)|(da)αEmax{|f(b)f(a)|(ba)α,|f(c)f(b)|(cb)α,|f(d)f(c)|(dc)α}\frac{\left\lvert g(d)-g(a)\right\rvert}{(d-a)^{\alpha}}\leq E\max\left\{\frac{\left\lvert f(b)-f(a)\right\rvert}{(b-a)^{\alpha}},\frac{\left\lvert f(c)-f(b)\right\rvert}{(c-b)^{\alpha}},\frac{\left\lvert f(d)-f(c)\right\rvert}{(d-c)^{\alpha}}\right\}
Proof.

We first start with the following basic inequality. Let λ1,λ2,λ3>0\lambda_{1},\lambda_{2},\lambda_{3}>0. Then

λ12α+λ22α+λ32α(λ1+λ2+λ3)αmax{31/2α,1}.\frac{\sqrt{\lambda_{1}^{2\alpha}+\lambda_{2}^{2\alpha}+\lambda_{3}^{2\alpha}}}{(\lambda_{1}+\lambda_{2}+\lambda_{3})^{\alpha}}\leq\max\{3^{1/2-\alpha},1\}.

This can be proved by noting that if α<1/2\alpha<1/2 the left hand side is maximised for λ1=λ2=λ3\lambda_{1}=\lambda_{2}=\lambda_{3} and if α>1/2\alpha>1/2 then it is maximised when two of the parameters are zero. Assume first that bcb\neq c. Then we can write (g(d)g(a))/(da)α(g(d)-g(a))/(d-a)^{\alpha} as the dot product

g(d)g(a)(da)α=1(da)α[(ba)α(cb)α(dc)α][M1M2M3], whereM1=f(b)f(a)(ba)α,M2=f(c)f(b)(cb)α,M3=f(d)f(c)(dc)α.\begin{split}\frac{g(d)-g(a)}{(d-a)^{\alpha}}&=\frac{1}{(d-a)^{\alpha}}\begin{bmatrix}(b-a)^{\alpha}&(c-b)^{\alpha}&(d-c)^{\alpha}\end{bmatrix}\begin{bmatrix}M_{1}\\ M_{2}\\ M_{3}\end{bmatrix}\text{, where}\\ M_{1}&=\frac{f(b)-f(a)}{(b-a)^{\alpha}},M_{2}=\frac{f(c)-f(b)}{(c-b)^{\alpha}},M_{3}=\frac{f(d)-f(c)}{(d-c)^{\alpha}}.\end{split}

In particular, by submultiplicativity of the dot product with respect to the usual vector norm and the aforementioned inequality we have

|g(d)g(a)(da)α|max{31/2α,1}M12+M22+M33max{31α,3}max{|M1|,|M2|,|M3|},\left\lvert\frac{g(d)-g(a)}{(d-a)^{\alpha}}\right\rvert\leq\max\{3^{1/2-\alpha},1\}\sqrt{M_{1}^{2}+M_{2}^{2}+M_{3}^{3}}\leq\max\{3^{1-\alpha},\sqrt{3}\}\max\{\left\lvert M_{1}\right\rvert,\left\lvert M_{2}\right\rvert,\left\lvert M_{3}\right\rvert\},

which proves the lemma for b<cb<c. For b=cb=c the proof is analogous . ∎

Lemma A.2.

The operator Pl,N:Cα([0,1])Cα([0,1])P_{l,N}:C^{\alpha}([0,1])\to C^{\alpha}([0,1]) is well-defined, continuous and has uniformly bounded operator norm over all l,Nl,N.

Proof.

It is fairly straightforward to show that for fCα([0,1])f\in C^{\alpha}([0,1]) the assignment fPl,N(f)f\mapsto P_{l,N}(f) is linear. We therefore just need to show continuity and Hölder continuity of fPl,N(f)f\mapsto P_{l,N}(f) and bound the operator norm.

We note that if (a,b)[0,1]𝒫l,N~(a,b)\subset[0,1]\setminus\widetilde{\mathcal{P}_{l,N}} is a connected component, then Pl,N(f)((a,b))f((a,b))P_{l,N}(f)((a,b))\subset f((a,b)) by the intermediate value theorem, which proves continuity at the accumulation points and hence continuity of ff everywhere.

We now estimate the Hölder norm. Let a<da<d be two points in [0,1][0,1]. We estimate the quantity H:=|Pl,N(f)(d)Pl,N(f)(a)|/|da|αH:=\left\lvert P_{l,N}(f)(d)-P_{l,N}(f)(a)\right\rvert/{\left\lvert d-a\right\rvert^{\alpha}}. By perturbing the two points by an arbitrarily small amount, we may assume aa and dd are not accumulation points of 𝒫l,N(f)\mathcal{P}_{l,N}(f). Let (x,b)(x,b) and (c,y)(c,y) be the connected components of [0,1]𝒫l,N~[0,1]\setminus\widetilde{\mathcal{P}_{l,N}} such that a(x,b)a\in(x,b) and d(c,y)d\in(c,y). In case the two intervals are the same, it is clear that HfαH\leq\|f\|_{\alpha}. In case they are different, we let g=Pl,N(f)g=P_{l,N}(f) on {a,b,c,d}\{a,b,c,d\} as in Lemma A.1, whence we obtain that

Hmax{31α,3}max{|g(b)g(a)(ba)α|,|g(c)g(b)(cb)α|,|g(d)g(c)(dc)α|}.H\leq\max\{3^{1-\alpha},\sqrt{3}\}\max\left\{\left\lvert\frac{g(b)-g(a)}{(b-a)^{\alpha}}\right\rvert,\left\lvert\frac{g(c)-g(b)}{(c-b)^{\alpha}}\right\rvert,\left\lvert\frac{g(d)-g(c)}{(d-c)^{\alpha}}\right\rvert\right\}.

The lemma follows after noting that Pl,N(f)P_{l,N}(f) coincides with ff on x,b,c,yx,b,c,y and using that

|Pl,N(f)(b)Pl,N(f)(a)|(ba)α|f(b)f(x)|(bx)α and |Pl,N(f)(d)Pl,N(f)(c)|(dc)α|f(y)f(c)|(yc)α.\frac{\left\lvert P_{l,N}(f)(b)-P_{l,N}(f)(a)\right\rvert}{(b-a)^{\alpha}}\leq\frac{\left\lvert f(b)-f(x)\right\rvert}{(b-x)^{\alpha}}\text{ and }\frac{\left\lvert P_{l,N}(f)(d)-P_{l,N}(f)(c)\right\rvert}{(d-c)^{\alpha}}\leq\frac{\left\lvert f(y)-f(c)\right\rvert}{(y-c)^{\alpha}}.

A.2 Ruelle’s Corollary

If T:B1B2T:B_{1}\to B_{2} is a bounded linear operator between Banach spaces, denote its kernel by ker(T)\ker(T). If TT has closed image Im(T)\operatorname{Im}(T), denote its cokernel by coker(T):=B2/Im(T)\operatorname{coker}(T):=B_{2}/\operatorname{Im}(T). We say TT is Fredholm of index kk if TT has closed image, coker(T):=B2/Im(T)\operatorname{coker}(T):=B_{2}/\operatorname{Im}(T) and ker(T)\ker(T) are finite and dimker(T)dimcoker(T)=k\mathrm{dim}\ker(T)-\mathrm{dim}\operatorname{coker}(T)=k. The composition of two Fredholm operators with respective indices k,lk,l is Fredholm with index k+lk+l. For a Banach space BB, we denote its dual by BB^{*} and denote by TT^{*} the induced adjoint operator T:B2B1:GGTT^{*}:B_{2}^{*}\to B_{1}^{*}:G\mapsto G\circ T. The following is a straightforward exercise in functional analysis.

Lemma A.3.

If T:B1B2T:B_{1}\to B_{2} has closed image and ker(T):={HB2:HT=0}\ker(T^{*}):=\{H\in B_{2}^{*}:H\circ T=0\} is finite-dimensional, then dimker(T)=dimcoker(T)\mathrm{dim}\ker(T^{*})=\mathrm{dim}\operatorname{coker}(T).

Suppose now B=B1=B2B=B_{1}=B_{2}, so T:BBT:B\to B. We recall the following facts about the spectral theory of operators on Banach spaces

  • If λσe(T)\lambda\neq\sigma_{e}(T), then λIT\lambda I-T is Fredholm of index 0.

  • If TT is a compact operator, so is TT^{*}.

We refer the reader to [7] for more details on the spectral theory of Banach operators, noting that σe(F)\sigma_{e}(F) in this preprint corresponds to σe5(F)\sigma_{e5}(F) 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 λ\lambda of LβL_{\beta} with λ>ρ(L(β)+α)ρe(Lβ)\lambda>\rho(L_{\Re(\beta)+\alpha})\geq\rho_{e}(L_{\beta}) is an eigenvalue of λβ\mathcal{\lambda}_{\beta} and that the dimensions of the respective generalised eigenspaces ELβ(λ)=r=1+ker(LβλI)rE_{L_{\beta}}(\lambda)=\bigcup_{r=1}^{+\infty}\ker(L_{\beta}-\lambda I)^{r} and Eβ(λ)=r=1+ker(βλI)rE_{\mathcal{L}_{\beta}}(\lambda)=\bigcup_{r=1}^{+\infty}\ker(\mathcal{L}_{\beta}-\lambda I)^{r} 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 β\mathcal{L}_{\beta}^{*} is compact operator and that the relevant operators are Fredholm of index 0, it suffices to prove that the generalised eigenspaces ELβ(λ)E_{L_{\beta}}(\lambda)^{*} and Eβ(λ)E_{\mathcal{L}_{\beta}}(\lambda)^{*} belonging respectively to LβL_{\beta}^{*} and β\mathcal{L}_{\beta}^{*} have equal dimension.

There is a natural (injective) inclusion map ι:A(D)Cα([0,1])\iota:A_{\infty}(D)\to C^{\alpha}([0,1]) defined by restriction to [0,1][0,1]. The dual operator ι:CαA(D)\iota^{*}:{C^{\alpha}}^{*}\to A_{\infty}(D)^{*} is defined by restricting distributions on Cα([0,1]){C^{\alpha}([0,1])} to distributions on A(D)A_{\infty}(D). This restricts to a natural linear map from ELβ(λ)E_{L_{\beta}}(\lambda)^{*} to Eβ(λ)E_{\mathcal{L}_{\beta}}(\lambda)^{*}. Surjectivity of this map follows from the Hahn-Banach theorem. We now show injectivity. Suppose that for some σELβ(λ)\sigma\in E_{L_{\beta}}(\lambda)^{*}, its restriction σA:A(D)\sigma^{A}:A_{\infty}(D)\to\mathbb{C} is identically zero. We show σ=0\sigma=0.

In order to do so, let α<α\alpha^{\prime}<\alpha such that λ>ρ(L(β)+α)\lambda>\rho(L_{\Re(\beta)+\alpha^{\prime}}). Use Hahn-Banach to extend σ\sigma to σα:Cα([0,1])\sigma^{\alpha^{\prime}}:C^{\alpha^{\prime}}([0,1])\to\mathbb{C}. For any fCα([0,1])f\in C^{\alpha}([0,1]), let fnf_{n} be a series of smooth approximations which converge in the Cα([0,1])C^{\alpha^{\prime}}([0,1])-norm. By cutting of fourier expansions outside of balls of increasing radii in Fourier space, we may assume the maps fnf_{n} are analytically extendable to elements of A(D)A_{\infty}(D). Hence we obtain

0=limnσA(fn)=limnσα(fn)=σα(f)=σ(f),0=\lim_{n\to\infty}\sigma^{A}(f_{n})=\lim_{n\to\infty}\sigma^{\alpha^{\prime}}(f_{n})=\sigma^{\alpha^{\prime}}(f)=\sigma(f),

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 PSL(2,𝐙){\rm PSL}(2,{\bf Z}) 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 PSL(2,𝐙){\rm PSL}(2,{\bf Z}) 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 PSL(2,𝐙){\rm PSL}(2,{\bf Z}) 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