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

Rigidity of pressures of Hölder potentials and the fitting of analytic functions via them

Liangang Ma and Mark Pollicott Liangang Ma, School of Mathematics and Statistics, Ludong University, Yantai 264025, Shandong, P. R. China. [email protected] Mark Pollicott, Department of Mathematics, Warwick University, Coventry, CV4 7AL, UK. [email protected]
Abstract.

The first part of this work is devoted to the study of higher differentials of pressure functions of Hölder potentials on shift spaces of finite type. By describing the differentials of pressure functions via the Central Limit Theorem for the associated random processes, we discover some rigid relationships between differentials of various orders. The rigidity imposes obstructions on fitting candidate convex analytic functions by pressure functions of Hölder potentials globally, which answers a question of Kucherenko-Quas. In the second part of the work we consider fitting candidate analytic germs by pressure functions of locally constant potentials. We prove that all 1-level candidate germs can be realised by pressures of some locally constant potentials, as long as number of the symbolic set is large enough. There are also some results on fitting 2-level germs by pressures of locally constant potentials obtained in the work.

The first author is supported by 12001056 from NSFC and ZR2019QA003 from SPNSF. The second author is partly supported by ERC-Advanced Grant 833802-Resonances and EPSRC Grant EP/T001674/1.

1. Introduction

This work deals with traditional topics in thermodynamic formalism [Bow, Rue1], which originates from theoretical physics. We focus on shift spaces of finite type here, which model dynamics of some smooth systems such as Axiom-A Diffeomorphisms through Markov partitions. Given a symbolic set Λ\Lambda of finite symbols and a continuous potential (observable) ϕ\phi on the shift space Λ\Lambda^{\mathbb{N}}, a core concept in thermodynamic formalism is the pressure P(ϕ)P(\phi). People are particularly interested in the pressure function P(tϕ)P(t\phi) with the variable t>0t>0 representing the inverse temperature. A sharp change in the pressure function (or other terms) is usually called a phase transition as tt varies, see for example [IRV, IT1, IT2, KQW, Lop1, Lop2, Lop3, Sar].

For Hölder continuous potentials, Ruelle [Rue2] proved that the pressure function P(tϕ)P(t\phi) is analytic for t(0,)t\in(0,\infty) (in fact he proved that P(ψ)P(\psi) depends analytically on ψ\psi for ψ\psi in the Hölder space C0,h(X)C^{0,h}(X) with XX being a transitive subshift space of finite type and 0<h10<h\leq 1 being the exponent [GT]). A key ingredient in his proof is the use of Ruelle (transfer) operator [BDL, GLP] acting on functions in the Hölder space. Moreover, the equilibrium measure of tϕt\phi for any t>0t>0 and Hölder potential ϕ\phi is always unique, so there are in fact no phase transitions in this case. Let

P(n)(t)=P(n)(tϕ)=dnP(tϕ)dtnP^{(n)}(t)=P^{(n)}(t\phi)=\cfrac{d^{n}P(t\phi)}{dt^{n}}

be the nn-th differential of the pressure function P(tϕ)P(t\phi) with respect to t(0,)t\in(0,\infty) for some fixed Hölder potential ϕ\phi. We also write

P(1)(t)=P(t),P(2)(t)=P′′(t),P(3)(t)=P′′′(t),P^{(1)}(t)=P^{\prime}(t),P^{(2)}(t)=P^{\prime\prime}(t),P^{(3)}(t)=P^{\prime\prime\prime}(t),\cdots

intermittently in the following. We discover that there is some rigid relationship between the differentials of the pressure function.

1.1 Theorem.

For a Hölder potential ϕ\phi on a full shift space of finite type, let P(t)=P(tϕ)P(t)=P(t\phi) be its pressure. Then there exists some positive number MϕM_{\phi} depending on ϕ\phi, such that

(1.1) 2π3(P(2)(t))3/2|P(3)(t)|9|P(3)(t)|+2|P(4)(t)|+32π3Mϕ(P(2)(t))5/2\sqrt{2\pi^{3}}(P^{(2)}(t))^{3/2}|P^{(3)}(t)|\leq 9|P^{(3)}(t)|+2|P^{(4)}(t)|+3\sqrt{2\pi^{3}}M_{\phi}(P^{(2)}(t))^{5/2}

for any t>0t>0.

A potential ϕ\phi is said to be generic (or we say it defines a non-lattice distribution, cf. [CP, Fel, PP]), if for any normalised potential ψ\psi, the spectral radius of the complex Ruelle operator ψ+itϕ\mathcal{L}_{\psi+it\phi} is less than 11 for any t0t\neq 0. For pressure functions of generic potentials, Theorem 1.1 can be strengthened to the following result.

1.2 Theorem.

For a generic Hölder potential ϕ\phi on a full shift space of finite type, let P(t)=P(tϕ)P(t)=P(t\phi) be its pressure. Then there exists some positive number MϕM_{\phi} depending on ϕ\phi, such that

(1.2) |P(3)(t)(12π(P(2)(t))3/2)|3MϕP(2)(t)|P^{(3)}(t)\big{(}1-\sqrt{2\pi}(P^{(2)}(t))^{3/2}\big{)}|\leq 3M_{\phi}P^{(2)}(t)

for any t>0t>0.

This means the second differential of the pressure function of a generic Hölder potential imposes some global subtle restriction on its third differential. It would be interesting to try to interpret the meaning of P′′(t)=12π3P^{\prime\prime}(t)=\frac{1}{\sqrt[3]{2\pi}} for the pressure function at individual parameters. Let σ:ΛΛ\sigma:\Lambda^{\mathbb{N}}\rightarrow\Lambda^{\mathbb{N}} denote the shift map. Both the proofs of Theorem 1.1 and 1.2 require use of the Ruelle operator and the Central Limit Theorem (CLT) for the process {fσn}n\{f\circ\sigma^{n}\}_{n\in\mathbb{N}}, with the latter one depending on a finer CLT in the generic case. Recall that there are some expressions on the higher differentials of the pressure function by Kotani and Sunada in [KS1] for smooth systems, and we refer the readers to [KS2] for a CLT for random walks on crystal lattices.

It is well-known that P(tϕ)P(t\phi) is convex and Lipschitz for continuous ϕ\phi, moreover, the supporting lines of its graph must intersect the vertical axis in a closed bounded interval in [0,)[0,\infty). Kucherenko and Quas have shown that any such function can be realised by the pressure function of some continuous potential on some shift space of finite type [KQ, Theorem 1], whose result fits into Katok’s flexibility programme [BKR]. However, the continuous potentials constructed in their work are not Hölder, so they ask the following question (their original problem is set in the multidimensional case).

1.3 Problem (Kucherenko-Quas).

Can a convex, Lipschitz analytic function with its supporting lines intersecting the vertical axis in a closed bounded interval in [0,)[0,\infty) be realised by the pressure function of some Hölder potential on some shift space of finite type?

Our following results are dedicated to an answer to their problem. We first point out that any convex, Lipschitz analytic function with its supporting lines intersecting the vertical axis in a closed bounded interval in [0,)[0,\infty) can be approximated by sequences of pressure functions of locally constant potentials on some shift space of finite type.

1.4 Corollary.

Let F(t)F(t) be a convex Lipschitz function on (α,)(\alpha,\infty) for some α>0\alpha>0 with Lipschitz constant L>0L>0, such that its supporting lines intersect the vertical axis in [γ¯,γ¯][\underline{\gamma},\overline{\gamma}] with 0γ¯γ¯<0\leq\underline{\gamma}\leq\overline{\gamma}<\infty. Then there exists a sequence of locally constant potentials {ϕn}n=1\{\phi_{n}\}_{n=1}^{\infty} on some shift space of finite type, such that

(1.3) limnP(tϕn)=F(t)\lim_{n\rightarrow\infty}P(t\phi_{n})=F(t)

for any t(α,)t\in(\alpha,\infty).

Proof.

This is an instant corollary of Kucherenko-Quas’ result. Let

Λ={0,1,,eγ¯}×{γ¯,,γ¯}}×{L,,L}\Lambda=\{0,1,\cdots,\lfloor e^{\overline{\gamma}}\rfloor\}\times\{\lfloor\underline{\gamma}\rfloor,\cdots,\lceil\overline{\gamma}\rceil\}\}\times\{\lfloor-L\rfloor,\cdots,\lceil L\rceil\},

in which \lfloor\ \rfloor and \lceil\ \rceil represent the floor and ceiling function respectively. According to [KQ, Theorem 1], there exists a continuous potential ϕF:Λ\phi_{F}:\Lambda^{\mathbb{Z}}\rightarrow\mathbb{R}, such that

P(tϕF)=F(t)P(t\phi_{F})=F(t)

on (α,)(\alpha,\infty). Now let

ϕn(x)ϕn,(x)=inf{ϕF(x):x[xnxn+1xn]}\phi_{n}(x)\doteq\phi_{n,-}(x)=\inf\{\phi_{F}(x):x\in[x_{-n}x_{-n+1}\cdots x_{n}]\}

for any x=x(n+1)xnxnxn+1Λx=\cdots x_{-(n+1)}x_{-n}\cdots x_{n}x_{n+1}\cdots\in\Lambda^{\mathbb{Z}} and nn\in\mathbb{N}, in which [xnxn+1xn][x_{-n}x_{-n+1}\cdots x_{n}] means the corresponding cylinder set. ϕn\phi_{n} is a locally constant potential for any fixed nn. Now fix t(α,)t\in(\alpha,\infty), by properties of the pressure function (see for example [Rue1, 6.8]),

(1.4) |P(tϕn)P(tϕF)||t|ϕnϕF.|P(t\phi_{n})-P(t\phi_{F})|\leq|t|\parallel\phi_{n}-\phi_{F}\parallel_{\infty}.

Since ϕF\phi_{F} is continuous, this implies (1.3).

One can see that in the above proof the increasing sequence of pressures {P(tϕn,)}n\{P(t\phi_{n,-})\}_{n\in\mathbb{N}} satisfies

P(tϕn,)F(t)P(t\phi_{n,-})\nearrow F(t)

as nn\rightarrow\infty since {ϕn,}n\{\phi_{n,-}\}_{n\in\mathbb{N}} is an increasing sequence tending to ϕF\phi_{F} (see [Wal1, Theorem 9.7(ii)]). Alternatively, one can take

ϕn,+(x)=sup{ϕF(x):x[xnxn+1xn]}\phi_{n,+}(x)=\sup\{\phi_{F}(x):x\in[x_{-n}x_{-n+1}\cdots x_{n}]\},

which results in a decreasing sequence of locally constant potentials approximating ϕF(x)\phi_{F}(x), or

ϕn,±(x)=ϕn,(x)+ϕn,+(x)2\phi_{n,\pm}(x)=\cfrac{\phi_{n,-}(x)+\phi_{n,+}(x)}{2},

which also results in a sequence of locally constant potentials approximating ϕF(x)\phi_{F}(x), while their pressure functions both approximate F(t)F(t). See Corollary 5.5 for an interpretation of the result from another point of view.

1.5 Remark.

The convergence in Corollary 1.4 is uniform for tt in a bounded domain since Λ\Lambda^{\mathbb{Z}} is a compact metric space by considering (1.4).

1.6 Remark.

A locally constant potential is of course Hölder, so according to Ruelle’s result, the pressure functions {P(tϕn,)}n\{P(t\phi_{n,-})\}_{n\in\mathbb{N}} are all analytic.

The following result confirms that some convex analytic functions cannot be fitted by the pressure of any Hölder potential on any shift space of finite type, which gives a negative answer to Problem 1.3.

1.7 Theorem.

For any α>0\alpha>0, there exists a strictly convex analytic function F(t)F(t) on (α,)(\alpha,\infty), with its supporting lines intersecting the vertical axis in [γ¯,γ¯][0,)[\underline{\gamma},\overline{\gamma}]\subset[0,\infty), such that there does not exist any Hölder potential ϕ\phi on any shift space of finite type satisfying

P(tϕ)=F(t)P(t\phi)=F(t)

on (α,)(\alpha,\infty).

For an explicit example of convex analytic functions in Theorem 1.7, one can simply take

F2,3,1(t)=2t2+3t+tet2+et2tF_{2,3,1}(t)=\cfrac{2t^{2}+3t+te^{-t^{2}}+e^{-t^{2}}}{t}

on (α,)(\alpha,\infty) for any α>0\alpha>0. See Proposition 4.2 for a family of such examples. Thus one can see that there are in fact elementary functions which cannot be fitted by pressures of Hölder potentials on shift spaces of finite type.

In the following we consider fitting convex analytic functions locally instead of globally, only by pressures of locally constant potentials on shift spaces of finite type. Let

Λn={1,2,,n}\Lambda_{n}=\{1,2,\cdots,n\}

be the symbolic set of nn symbols.

1.8 Theorem.

Let t>0t_{*}>0 and (a0,a1)2(a_{0},a_{1})\in\mathbb{R}^{2} satisfying

(1.5) a0t>a1.\frac{a_{0}}{t_{*}}>a_{1}.

Then for any nn\in\mathbb{N} large enough, there exist some 0mt,a0,a1,n<Mt,a0,a1,n<0\leq m_{t_{*},a_{0},a_{1},n}<M_{t_{*},a_{0},a_{1},n}<\infty depending on t,a0,a1,nt_{*},a_{0},a_{1},n, such that for any a2[mt,a0,a1,n,Mt,a0,a1,n]a_{2}\in[m_{t_{*},a_{0},a_{1},n},M_{t_{*},a_{0},a_{1},n}], there exists some sequence of reals {ci,n}i=1n\{c_{i,n}\}_{i=1}^{n}, such that the locally constant potential

ϕ(x)=cx0,n\phi(x)=c_{x_{0},n}

for x=x1x0x1[x0]x=\cdots x_{-1}x_{0}x_{1}\cdots\in[x_{0}] on the full shift space Λn\Lambda_{n}^{\mathbb{Z}} satisfies

(1.6) P(tϕ)=a0+a1(tt)+a22!(tt)2+O((tt)3)P(t\phi)=a_{0}+a_{1}(t-t_{*})+\cfrac{a_{2}}{2!}(t-t_{*})^{2}+O((t-t_{*})^{3})

on [tδn,t+δn][t_{*}-\delta_{n},t_{*}+\delta_{n}] for some δn>0\delta_{n}>0.

This means we can fit some germs at tt_{*} up to level 22 by pressures of some locally constant potentials when the number of symbols of the shift space is large enough. The values δn,{ci,n}i=1n\delta_{n},\{c_{i,n}\}_{i=1}^{n} all depend on t,a0,a1,nt_{*},a_{0},a_{1},n and a2a_{2} in fact, while we only indicate the dependence of mt,a0,a1,nm_{t_{*},a_{0},a_{1},n} and Mt,a0,a1,nM_{t_{*},a_{0},a_{1},n} as we are particularly interested in their values in the context of Theorem 1.8. There are some results on the values of

{mt,a0,a1,n,Mt,a0,a1,n}n\{m_{t_{*},a_{0},a_{1},n},M_{t_{*},a_{0},a_{1},n}\}_{n\in\mathbb{N}}

subject to t>0t_{*}>0 and (a0,a1)2(a_{0},a_{1})\in\mathbb{R}^{2} satisfying (1.5) at the end of Section 5.

We choose to present all our results in the one dimensional case, while many of these results can in fact be extended to convex Lipschitz or analytic functions F(t1,t2,,tm)F(t_{1},t_{2},\cdots,t_{m}) of mm variables naturally. Most of our results also hold on transitive subshift spaces of finite type, with some technical adjustments in their proofs involving the transition matrix.

The organization of the work is as following. In Section 2 we introduce some basics in thermodynamic formalism and the Central Limit Theorem for the process generated by a potential and the shift map on the symbolic space of finite type. We give an explicit bound on the tail term in the CLT. Section 3 is devoted to the proof of Theorem 1.1 and 1.2. We formulate some expression of the derivatives of the pressure (Corollary 3.11) linking directly to the CLT, which allows us to unveil the relationship between derivatives of the pressure function of various orders. Section 4 is devoted to the proof of Theorem 1.7. In Section 5 we consider fitting 1- and 2-level candidate analytic germs locally by pressure functions of locally constant potentials (Problem 5.2) on symbolic spaces of finite type. We conjecture that any reasonable analytic germ of finite level can be fitted by the pressure function of some locally constant potential locally, as long as the number of the symbols is large enough.

2. Thermodynamic formalism and the CLT

In this section we collect some basic notions and results in thermodynamic formalism for later use. We start from the pressure. Let Λ\Lambda be some symbolic set of finite symbols, Λ\Lambda^{\mathbb{N}} be the shift space equipped with the visual metric

d(x,y)=12l(x,y)d(x,y)=\cfrac{1}{2^{l(x,y)}}

for distinct x=x0x1x2,y=y0y1y2Λx=x_{0}x_{1}x_{2}\cdots,y=y_{0}y_{1}y_{2}\cdots\in\Lambda^{\mathbb{N}}, in which

l(x,y)=min{i:xiyi}l(x,y)=\min\{i\in\mathbb{N}:x_{i}\neq y_{i}\}.

For a continuous potential ϕ:Λ\phi:\Lambda^{\mathbb{N}}\rightarrow\mathbb{R} on the compact metric space Λ\Lambda^{\mathbb{N}}, Let

Sm,ϕ(x)=i=0m1ϕσi(x)S_{m,\phi}(x)=\sum_{i=0}^{m-1}\phi\circ\sigma^{i}(x)

for mm\in\mathbb{N}, in which σ\sigma is the shift map.

2.1 Definition.

The pressure P(ϕ)P(\phi) of a continuous potential ϕ\phi on Λ\Lambda^{\mathbb{N}} is defined to be

P(ϕ)=limm1mlogσm(x)=xeSm,ϕ(x)P(\phi)=\lim_{m\rightarrow\infty}\cfrac{1}{m}\log\sum_{\sigma^{m}(x)=x}e^{S_{m,\phi}(x)}.

One can refer to [Wal1, p208] for a definition for continuous potentials on general compact metric spaces. It satisfies the well-known variational formula

P(ϕ)=sup{h(μ)+ϕ𝑑μ:μ is a σinvariant measure on Λ}P(\phi)=\sup\{h(\mu)+\int\phi d\mu:\mu\mbox{ is a }\sigma-\mbox{invariant measure on }\Lambda^{\mathbb{N}}\}.

Let C0(Λ)C^{0}(\Lambda^{\mathbb{N}}) be the collection of all the continuous potentials on Λ\Lambda^{\mathbb{N}}. Two potentials ψ,ϕC0(Λ)\psi,\phi\in C^{0}(\Lambda^{\mathbb{N}}) are said to be cohomologous [Wal2] in case there exists a continuous map φ:Λ\varphi:\Lambda^{\mathbb{N}}\rightarrow\mathbb{R} such that

ψ(x)ϕ(x)=φ(x)φσ(x)\psi(x)-\phi(x)=\varphi(x)-\varphi\circ\sigma(x).

We write ψϕ\psi\sim\phi to denote the equivalence relationship between two potentials cohomologous to each other. The maps in

{φ(x)φσ(x):φC0(Λ)}\{\varphi(x)-\varphi\circ\sigma(x):\varphi\in C^{0}(\Lambda^{\mathbb{N}})\}

are called coboundaries. The importance of the cohomologous relationship is revealed in the following result.

2.2 Proposition.

If ψϕ\psi\sim\phi, then P(ψ)=P(ϕ)P(\psi)=P(\phi). Moreover, ψ\psi and ϕ\phi share the same equilibrium state.

One can find a proof in [Rue1] or [PP]. Another important tool in thermodynamic formalism is the Ruelle operator.

2.3 Definition.

For a continuous potential ψ:Λ\psi:\Lambda^{\mathbb{N}}\rightarrow\mathbb{R}, define the Ruelle operator ψ\mathcal{L}_{\psi} acting on C0(Λ)C^{0}(\Lambda^{\mathbb{N}}) as

(ψf)(x)=y:σ(y)=xeψ(y)f(y)(\mathcal{L}_{\psi}f)(x)=\sum_{y:\sigma(y)=x}e^{\psi(y)}f(y)

for fC0(Λ)f\in C^{0}(\Lambda^{\mathbb{N}}).

One can see easily that its compositions satisfy

(2.1) (ψmf)(x)=y:σm(y)=xeSm,ψ(y)f(y)(\mathcal{L}^{m}_{\psi}f)(x)=\sum_{y:\sigma^{m}(y)=x}e^{S_{m,\psi}(y)}f(y)

for any mm\in\mathbb{N}. In case of ψ\psi being Hölder, it admits a simple maximum isolated eigenvalue λ=eP(ψ)\lambda=e^{P(\psi)} such that,

(2.2) (ψwψ)(x)=eP(ψ)wψ(x)(\mathcal{L}_{\psi}w_{\psi})(x)=e^{P(\psi)}w_{\psi}(x)

for some eigenfunction wψ(x)C0,h(Λ)w_{\psi}(x)\in C^{0,h}(\Lambda^{\mathbb{N}}), refer to [Rue1]. The unique equilibrium measure for the Hölder potential ψ\psi is denoted by μψ\mu_{\psi}. It then follows that

(2.3) (ψmwψ)(x)=emP(ψ)wψ(x)(\mathcal{L}^{m}_{\psi}w_{\psi})(x)=e^{mP(\psi)}w_{\psi}(x)

for wψ(x)C0,h(Λ)w_{\psi}(x)\in C^{0,h}(\Lambda^{\mathbb{N}}). A potential ψ\psi is said to be normalized if

P(ψ)=0P(\psi)=0 and wψ=1Λw_{\psi}=1_{\Lambda^{\mathbb{N}}},

in which 1Λ1_{\Lambda^{\mathbb{N}}} is the identity map on Λ\Lambda^{\mathbb{N}}. In case of ψ\psi being not normalized, we call

ψ¯=ψ+logwψlogwψσP(ψ)\bar{\psi}=\psi+\log w_{\psi}-\log w_{\psi}\circ\sigma-P(\psi)

the normalization of ψ\psi. It is easy to check that ψ¯\bar{\psi} is a normalized potential. Moreover, ψ¯\bar{\psi} and ψ\psi share the same equilibrium state.

Now we turn to the Central Limit Theorem for the random process {ϕσj(x)}j=0\{\phi\circ\sigma^{j}(x)\}_{j=0}^{\infty} with the equilibrium measure μψ\mu_{\psi} defined by some Hölder potential ψ\psi, while ϕ\phi is also assumed to be Hölder. It deals with the asymptotic behaviour of the distribution of Sm,ϕm\cfrac{S_{m,\phi}}{\sqrt{m}} with respect to μψ\mu_{\psi} as mm\rightarrow\infty. The Ruelle operator comes in here, see [CP, Lal, Rou]. Let

Gm(y)=μψ{xΛ:Sm,ϕ(x)m<y}G_{m}(y)=\mu_{\psi}\Big{\{}x\in\Lambda^{\mathbb{N}}:\cfrac{S_{m,\phi}(x)}{\sqrt{m}}<y\Big{\}}

for yy\in\mathbb{R}. For a,ba,b\in\mathbb{R} and b>0b>0, Let Na,b(y)N_{a,b}(y) be the normal distribution with expectation aa and standard deviation b\sqrt{b} on \mathbb{R}, that is,

dNa,b(y)dy=12πbe(ya)2/2b\cfrac{dN_{a,b}(y)}{dy}=\cfrac{1}{\sqrt{2\pi b}}e^{-(y-a)^{2}/2b}

for yy\in\mathbb{R}. For Hölder potentials ψ,ϕ\psi,\phi on a shift space of finite type, since the pressure P(ψ+tϕ)P(\psi+t\phi) is analytic in a small neighbourhood around 0, denote by

Δm=P(m)(ψ+tϕ)|t=0\Delta_{m}=P^{(m)}(\psi+t\phi)|_{t=0}

for mm\in\mathbb{N} for convenience, while the readers can understand its dependence on ψ,ϕ\psi,\phi easily from the contexts in the following. Let

P(ψ+tϕ)=m=0Δmm!tm=m=03Δmm!tm+t4κ(t)P(\psi+t\phi)=\sum_{m=0}^{\infty}\cfrac{\Delta_{m}}{m!}t^{m}=\sum_{m=0}^{3}\cfrac{\Delta_{m}}{m!}t^{m}+t^{4}\kappa(t),

in which κ(t)=m=0Δm+4(m+4)!tm\kappa(t)=\sum_{m=0}^{\infty}\cfrac{\Delta_{m+4}}{(m+4)!}t^{m}.

Central Limit Theorem.

Let ψ,ϕ\psi,\phi be Hölder potentials on a shift space of finite type with ϕ\phi being not cohomologous to a constant. If ϕ𝑑μψ=0\int\phi d\mu_{\psi}=0, we have

limmGm(y)=N0,Δ2(y)+O(1/m)\lim_{m\rightarrow\infty}G_{m}(y)=N_{0,\Delta_{2}}(y)+O(1/\sqrt{m}),

in which

(2.4) O(1/m)9|Δ3|+2|Δ4|2π3m(Δ2)3/2.O(1/\sqrt{m})\leq\cfrac{9|\Delta_{3}|+2|\Delta_{4}|}{\sqrt{2\pi^{3}m}(\Delta_{2})^{3/2}}.

The convergence is uniform with respect to yy. In case of ϕ\phi being generic, the result can be strengthened to

limmGm(y)=N0,Δ2(y)+Hm(y)+o(1/m)\lim_{m\rightarrow\infty}G_{m}(y)=N_{0,\Delta_{2}}(y)+H_{m}(y)+o(1/\sqrt{m}),

in which Hm(y)=Δ36m(1y2Δ2)ey22Δ2H_{m}(y)=\cfrac{\Delta_{3}}{6\sqrt{m}}\Big{(}1-\cfrac{y^{2}}{\Delta_{2}}\Big{)}e^{-\frac{y^{2}}{2\Delta_{2}}}.

This fits into special cases of the Berry-Esseen Theorem [Fel]. There is nothing new in the version here comparing with [CP, Theorem 2, Theorem 3] or [PP, Theorem 4.13], except the explicit bound on the tail term O(1/m)O(1/\sqrt{m}) in (2.4). In the following we justify this explicit bound. To do this, let

χm(z)=eizSm,ϕm𝑑μψ\chi_{m}(z)=\displaystyle\int e^{iz\frac{S_{m,\phi}}{\sqrt{m}}}d\mu_{\psi}

be the Fourier transformation of Gm(y)G_{m}(y). Note that the Fourier transformation of N0,Δ2(y)N_{0,\Delta_{2}}(y) is ez2Δ22e^{-\frac{z^{2}\Delta_{2}}{2}}.

2.4 Lemma.

Let ψ,ϕ\psi,\phi be Hölder potentials on a shift space of finite type with ϕ\phi being not cohomologous to a constant. For ϵ>0\epsilon>0 small enough, we have

(2.5) 12π0ϵm1z|χm(z)ez2Δ22|𝑑z2|Δ3|12πm(Δ2)3/2\cfrac{1}{2\pi}\int_{0}^{\epsilon\sqrt{m}}\cfrac{1}{z}\Big{|}\chi_{m}(z)-e^{-\frac{z^{2}\Delta_{2}}{2}}\Big{|}dz\leq\cfrac{\sqrt{2}|\Delta_{3}|}{12\sqrt{\pi m}(\Delta_{2})^{3/2}}

for any mm\in\mathbb{N} large enough.

Proof.

According to [PP, (4.6)], we have

0ϵm1z|χm(z)ez2Δ22+iz3Δ36mez2Δ22|𝑑z=O(1/m)\displaystyle\int_{0}^{\epsilon\sqrt{m}}\cfrac{1}{z}\Big{|}\chi_{m}(z)-e^{-\frac{z^{2}\Delta_{2}}{2}}+\cfrac{iz^{3}\Delta_{3}}{6\sqrt{m}}e^{-\frac{z^{2}\Delta_{2}}{2}}\Big{|}dz=O(1/m)

for ϵ>0\epsilon>0 small enough. So

(2.6) 12π0ϵm1z|χm(z)ez2Δ22|𝑑zO(1/m)+|Δ3|12πm0ϵmz2ez2Δ22𝑑z.\cfrac{1}{2\pi}\displaystyle\int_{0}^{\epsilon\sqrt{m}}\cfrac{1}{z}\Big{|}\chi_{m}(z)-e^{-\frac{z^{2}\Delta_{2}}{2}}\Big{|}dz\leq O(1/m)+\cfrac{|\Delta_{3}|}{12\pi\sqrt{m}}\displaystyle\int_{0}^{\epsilon\sqrt{m}}z^{2}e^{-\frac{z^{2}\Delta_{2}}{2}}dz.

Considering

z2ez2Δ22𝑑z=2π(Δ2)3/2\displaystyle\int_{-\infty}^{\infty}z^{2}e^{-\frac{z^{2}\Delta_{2}}{2}}dz=\cfrac{\sqrt{2\pi}}{(\Delta_{2})^{3/2}},

we obtain (2.5) from (2.6). ∎

Equipped with Lemma 2.4 we can justify the explicit bound on the tail term in the Central Limit Theorem in (2.4).

Proof of the tail term in CLT:

Proof.

Without loss of generality, suppose ψ\psi is normalized and ϕ𝑑μψ=0\int\phi d\mu_{\psi}=0. It suffices for us to justify (2.4) considering [CP, Theorem 2, Theorem 3]. Similar to the proof of [CP, Theorem 2], apply [Fel, Lemma 2] with the cumulative functions Gm(y)G_{m}(y) and N0,Δ2(y)N_{0,\Delta_{2}}(y) in our case, one gets (c.f. [CP, (20)])

(2.7) |Gm(y)N0,Δ2(y)|12π0ϵm1z|χm(z)ez2Δ22|𝑑z+24ϵ2mπ3Δ2.|G_{m}(y)-N_{0,\Delta_{2}}(y)|\leq\cfrac{1}{2\pi}\int_{0}^{\epsilon\sqrt{m}}\cfrac{1}{z}\Big{|}\chi_{m}(z)-e^{-\frac{z^{2}\Delta_{2}}{2}}\Big{|}dz+\cfrac{24}{\epsilon\sqrt{2m\pi^{3}\Delta_{2}}}.

Now let us take

1ϵ=2Δ2(|Δ3|6+|Δ4|24+δ)\cfrac{1}{\epsilon}=\cfrac{2}{\Delta_{2}}\big{(}\cfrac{|\Delta_{3}|}{6}+\cfrac{|\Delta_{4}|}{24}+\delta\big{)}

for some small δ>0\delta>0, such that it satisfies (c.f. [CP, (10)])

1ϵ>max{2Δ2(|Δ3|6+tκ(t)),2Δ2κ(t)}\cfrac{1}{\epsilon}>\max\Big{\{}\cfrac{2}{\Delta_{2}}\Big{(}\cfrac{|\Delta_{3}|}{6}+t\kappa(t)\Big{)},\cfrac{2}{\Delta_{2}}\kappa(t)\Big{\}}

for any |t|<ϵ|t|<\epsilon in (2.7). Considering (2.5), we have

(2.8) |Gm(y)N0,Δ2(y)|2|Δ3|12πm(Δ2)3/2+242mπ3Δ22Δ2(|Δ3|6+|Δ4|24+δ)=2|Δ3|12πm(Δ2)3/2+8|Δ3|2π3m(Δ2)3/2+2|Δ4|2π3m(Δ2)3/2+48δ2π3m(Δ2)3/29|Δ3|2π3m(Δ2)3/2+2|Δ4|2π3m(Δ2)3/2+48δ2π3m(Δ2)3/2.\begin{array}[]{ll}&|G_{m}(y)-N_{0,\Delta_{2}}(y)|\vspace{3mm}\\ \leq&\cfrac{\sqrt{2}|\Delta_{3}|}{12\sqrt{\pi m}(\Delta_{2})^{3/2}}+\cfrac{24}{\sqrt{2m\pi^{3}\Delta_{2}}}\cfrac{2}{\Delta_{2}}\big{(}\cfrac{|\Delta_{3}|}{6}+\cfrac{|\Delta_{4}|}{24}+\delta\big{)}\vspace{3mm}\\ =&\cfrac{\sqrt{2}|\Delta_{3}|}{12\sqrt{\pi m}(\Delta_{2})^{3/2}}+\cfrac{8|\Delta_{3}|}{\sqrt{2\pi^{3}m}(\Delta_{2})^{3/2}}+\cfrac{2|\Delta_{4}|}{\sqrt{2\pi^{3}m}(\Delta_{2})^{3/2}}+\cfrac{48\delta}{\sqrt{2\pi^{3}m}(\Delta_{2})^{3/2}}\vspace{3mm}\\ \leq&\cfrac{9|\Delta_{3}|}{\sqrt{2\pi^{3}m}(\Delta_{2})^{3/2}}+\cfrac{2|\Delta_{4}|}{\sqrt{2\pi^{3}m}(\Delta_{2})^{3/2}}+\cfrac{48\delta}{\sqrt{2\pi^{3}m}(\Delta_{2})^{3/2}}.\end{array}

Finally, let δ0\delta\rightarrow 0 in (2.8), we get (2.4).

We will deal with the pressure function P(ψ+tϕ)P(\psi+t\phi) for t0t\geq 0 and ψ,ϕC0,h(Λ)\psi,\phi\in C^{0,h}(\Lambda^{\mathbb{N}}) for some 0<h10<h\leq 1 in the following sections. By [Rue2], P(ψ+tϕ)P(\psi+t\phi) depends analytically on tt in case that ψ,ϕ\psi,\phi are Hölder. We will often assume that

ϕ𝑑μψ=0\int\phi d\mu_{\psi}=0

in the following when dealing with the higher differentials of P(ψ+tϕ)P(\psi+t\phi) because if ϕ𝑑μψ=c0\int\phi d\mu_{\psi}=c\neq 0, we have

P(ψ+t(ϕc))=P(ψ+tϕ)ctP(\psi+t(\phi-c))=P(\psi+t\phi)-ct,

then

(2.9) dnP(ψ+t(ϕc))dtn=dnP(ψ+tϕ)dtn\cfrac{d^{n}P(\psi+t(\phi-c))}{dt^{n}}=\cfrac{d^{n}P(\psi+t\phi)}{dt^{n}}

for any n2n\geq 2 while (ϕc)𝑑μψ=0\int(\phi-c)d\mu_{\psi}=0. We can also assume that ψ\psi is normalized when dealing with the differentials of P(ψ+tϕ)P(\psi+t\phi). If this is not the case we can simply change ψ\psi to its normalization ψ¯\bar{\psi} while

(2.10) dnP(ψ+tϕ)dtn=dnP(ψ¯+tϕ)dtn\cfrac{d^{n}P(\psi+t\phi)}{dt^{n}}=\cfrac{d^{n}P(\bar{\psi}+t\phi)}{dt^{n}}

for n1n\geq 1 because

P(ψ¯+tϕ)=P(ψ+tϕ)P(ψ)P(\bar{\psi}+t\phi)=P(\psi+t\phi)-P(\psi)

for any tt\in\mathbb{R}.

3. Derivatives of the pressures of Hölder potentials

In this section we formulate some explicit expressions for the derivatives of the pressure P(tϕ)=P(t)P(t\phi)=P(t) in terms of the derivatives of the eigenfunction of tϕ\mathcal{L}_{t\phi} for ϕC0,h(Λ)\phi\in C^{0,h}(\Lambda^{\mathbb{N}}) with respect to tt. We give basically two expressions of the derivatives, one of which allows the introduction of the random stochastic process {ϕσj(x)}j=0m\{\phi\circ\sigma^{j}(x)\}_{j=0}^{m} for mm\in\mathbb{N}. Upon the expression we prove Theorem 1.1 and 1.2 in virtue of the CLT for the random process {ϕσj(x)}j=0\{\phi\circ\sigma^{j}(x)\}_{j=0}^{\infty}.

First we define some basics to deal with the higher derivatives of compositional functions by the Faà di Bruno’s formula. For an integer jj\in\mathbb{N}, we say

τ=τ1τ2τq\tau=\tau_{1}\tau_{2}\cdots\tau_{q}

with qq\in\mathbb{N} is a partition of jj if the non-increasing sequence of positive integers jτ1τ2τq1j\geq\tau_{1}\geq\tau_{2}\geq\cdots\geq\tau_{q}\geq 1 satisfies i=1qτi=j\sum_{i=1}^{q}\tau_{i}=j. Denote the collection of all the possible partitions of jj by 𝔓(j)\mathfrak{P}(j). For example, Table 1 lists all the partitions in 𝔓(5)\mathfrak{P}(5).

Table 1. Partitions of 55
5 q=1
4,1 q=2
3,2 q=2
3,1,1 q=3
2,2,1 q=3
2,1,1,1 q=4
1,1,1,1,1 q=5

We sometimes simply write τ\tau to denote the set {τ1,τ2,,τq}\{\tau_{1},\tau_{2},\cdots,\tau_{q}\} for convenience in the following, so #τ=q\#\tau=q. Now for τ\tau being a partition of j1j\geq 1, let {Bjτ}\{B_{j}^{\tau}\} be the number of different choices of dividing a set of jj different elements into #τ=q\#\tau=q sets of sizes {τi}i=1q\{\tau_{i}\}_{i=1}^{q} respectively (with no order on the sets of partitions). Set B00=1B_{0}^{0}=1 for convenience. For example, consider the cases j=5j=5 and τ=3,1,1\tau=3,1,1, the number of different choices of dividing a set of 55 different elements into q=3q=3 sets of sizes 3,1,13,1,1 respectively is

C53=10=B53,1,1C_{5}^{3}=10=B_{5}^{3,1,1}.

Table 2 lists all the numbers {B5τ}τ𝔓(5)\{B_{5}^{\tau}\}_{\tau\in\mathfrak{P}(5)}.

Table 2. The coefficients B5τB_{5}^{\tau}
B55=1B_{5}^{5}=1
B54,1=5B_{5}^{4,1}=5
B53,2=10B_{5}^{3,2}=10
B53,1,1=10B_{5}^{3,1,1}=10
B52,2,1=15B_{5}^{2,2,1}=15
B52,1,1,1=10B_{5}^{2,1,1,1}=10
B51,1,1,1,1=1B_{5}^{1,1,1,1,1}=1

For a smooth map f:XYf:X\rightarrow Y between two metric spaces X,YX,Y and some partition τ=τ1,τ2,,τq𝔓(j)\tau=\tau_{1},\tau_{2},\cdots,\tau_{q}\in\mathfrak{P}(j) with j1j\geq 1, let

f(τ)(x)=f(τ1)(x)f(τ2)(x)f(τq)(x)f^{(\tau)}(x)=f^{(\tau_{1})}(x)f^{(\tau_{2})}(x)\cdots f^{(\tau_{q})}(x)

be the product of the derivatives. For j=0j=0 and τ=0𝔓(0)\tau=0\in\mathfrak{P}(0), set f(0)(x)=1f^{(0)}(x)=1. Then for two smooth functions f:XYf:X\rightarrow Y and g:YZg:Y\rightarrow Z between metric spaces X,Y,ZX,Y,Z, we have

(3.1) dj(gf(x))dxj=τ𝔓(j)Bjτg(#τ)(f(x))f(τ)(x)\cfrac{d^{j}(g\circ f(x))}{dx^{j}}=\sum_{\tau\in\mathfrak{P}(j)}B_{j}^{\tau}g^{(\#\tau)}(f(x))f^{(\tau)}(x)

in virtue of Faà di Bruno’s formula.

Now we turn to the higher differentials of the pressure function. We start by considering some standard case, then extend the result to the general case.

3.1 Theorem.

Let ψ,ϕC0,h(Λ)\psi,\phi\in C^{0,h}(\Lambda^{\mathbb{N}}) with ψ\psi being normalized for some finite symbolic set Λ\Lambda. Assume ϕ𝑑μψ=0\int\phi d\mu_{\psi}=0, in which μψ\mu_{\psi} is the equilibrium state of ψ\psi. Let w(t,x)w(t,x) be the eigenfunction of the maximum isolated eigenvalue eP(ψ+tϕ)e^{P(\psi+t\phi)} of ψ+tϕ\mathcal{L}_{\psi+t\phi}, which depends analytically on tt in a small neighbourhood of 0. Considering the differentials of the pressure function P(ψ+tϕ)P(\psi+t\phi) at t=0t=0, we have

(3.2) P(n)(ψ+tϕ)|t=0=j=1nCnjΛ(ϕ(x))jw(nj)(0,x)𝑑μψ(x)j=2n2Cnjτ𝔓(j),1τBjτP(τ)(ψ+tϕ)|t=0Λw(nj)(0,x)𝑑μψ(x)τ𝔓(n),{1,n}τ=BnτP(τ)(ψ+tϕ)|t=0\begin{array}[]{ll}P^{(n)}(\psi+t\phi)|_{t=0}=&\sum_{j=1}^{n}C_{n}^{j}\int_{\Lambda^{\mathbb{N}}}(\phi(x))^{j}w^{(n-j)}(0,x)d\mu_{\psi}(x)\\ &-\sum_{j=2}^{n-2}C_{n}^{j}\sum_{\tau\in\mathfrak{P}(j),1\notin\tau}B_{j}^{\tau}P^{(\tau)}(\psi+t\phi)|_{t=0}\int_{\Lambda^{\mathbb{N}}}w^{(n-j)}(0,x)d\mu_{\psi}(x)\\ &-\sum_{\tau\in\mathfrak{P}(n),\{1,n\}\cap\tau=\emptyset}B_{n}^{\tau}P^{(\tau)}(\psi+t\phi)|_{t=0}\end{array}

for any n2n\geq 2.

Proof.

According to the above notations, note that

(3.3) (ψ+tϕw(t,))(x)=eP(ψ+tϕ)w(t,x).(\mathcal{L}_{\psi+t\phi}w(t,\cdot))(x)=e^{P(\psi+t\phi)}w(t,x).

The nn-th derivative of (ψ+tϕw(t,))(x)=y:σ(y)=xeψ(y)+tϕ(y)w(t,y)(\mathcal{L}_{\psi+t\phi}w(t,\cdot))(x)=\sum_{y:\sigma(y)=x}e^{\psi(y)+t\phi(y)}w(t,y) gives

(3.4) dnψ+tϕw(t,)(x)dtn=y:σ(y)=xj=0nCnjdje(ψ+tϕ)(y)dtjw(nj)(t,y)=y:σ(y)=xj=0nCnje(ψ+tϕ)(y)(ϕ(y))jw(nj)(t,y)=j=0nCnjψ+tϕ((ϕ())jw(nj)(t,)).\begin{array}[]{ll}&\cfrac{d^{n}\mathcal{L}_{\psi+t\phi}w(t,\cdot)(x)}{dt^{n}}\vspace{1mm}\\ =&\sum_{y:\sigma(y)=x}\sum_{j=0}^{n}C_{n}^{j}\cfrac{d^{j}e^{(\psi+t\phi)(y)}}{dt^{j}}w^{(n-j)}(t,y)\vspace{3mm}\\ =&\sum_{y:\sigma(y)=x}\sum_{j=0}^{n}C_{n}^{j}e^{(\psi+t\phi)(y)}(\phi(y))^{j}w^{(n-j)}(t,y)\vspace{3mm}\\ =&\sum_{j=0}^{n}C_{n}^{j}\mathcal{L}_{\psi+t\phi}\big{(}(\phi(\cdot))^{j}w^{(n-j)}(t,\cdot)\big{)}.\end{array}

All differentials are with respect to tt. In case of t=0t=0 this means

(3.5) dnψ+tϕw(t,)(x)dtn|t=0=j=0nCnjψ((ϕ())jw(nj)(0,)).\cfrac{d^{n}\mathcal{L}_{\psi+t\phi}w(t,\cdot)(x)}{dt^{n}}|_{t=0}=\sum_{j=0}^{n}C_{n}^{j}\mathcal{L}_{\psi}\big{(}(\phi(\cdot))^{j}w^{(n-j)}(0,\cdot)\big{)}.

Note that the dual operator ψ\mathcal{L}_{\psi}^{*} fixes μψ\mu_{\psi}, so integration of both sides of (3.5) gives

(3.6) dnψ+tϕw(t,)(x)dtn|t=0dμψ(x)=j=0nCnj(ϕ(x))jw(nj)(0,x)μψ(x).\int\cfrac{d^{n}\mathcal{L}_{\psi+t\phi}w(t,\cdot)(x)}{dt^{n}}|_{t=0}d\mu_{\psi}(x)=\sum_{j=0}^{n}C_{n}^{j}\int(\phi(x))^{j}w^{(n-j)}(0,x)\mu_{\psi}(x).

In order to get the nn-th derivative of P(ψ+tϕ)P(\psi+t\phi), differentiate eP(ψ+tϕ)w(t,x)e^{P(\psi+t\phi)}w(t,x) for nn times by (3.1), we get

(3.7) dn(eP(ψ+tϕ)w(t,x))dtn=j=0nCnjdjeP(ψ+tϕ)dtjw(nj)(t,x)=j=0n1CnjdjeP(ψ+tϕ)dtjw(nj)(t,x)+dneP(ψ+tϕ)dtnw(t,x)=j=0n1Cnjτ𝔓(j)BjτP(τ)(ψ+tϕ)eP(ψ+tϕ)w(nj)(t,x)+τ𝔓(n)BnτP(τ)(ψ+tϕ)eP(ψ+tϕ)w(t,x)=j=0n1Cnj(τ𝔓(j),1τBjτP(τ)(ψ+tϕ)+τ𝔓(j),1τBjτP(τ)(ψ+tϕ))eP(ψ+tϕ)w(nj)(t,x)+τ𝔓(n),nτBnτP(τ)(ψ+tϕ)eP(ψ+tϕ)w(t,x)+P(n)(ψ+tϕ)eP(ψ+tϕ)w(t,x).\begin{array}[]{ll}&\cfrac{d^{n}\Big{(}e^{P(\psi+t\phi)}w(t,x)\Big{)}}{dt^{n}}\vspace{1mm}\\ =&\sum_{j=0}^{n}C_{n}^{j}\cfrac{d^{j}e^{P(\psi+t\phi)}}{dt^{j}}w^{(n-j)}(t,x)\vspace{3mm}\\ =&\sum_{j=0}^{n-1}C_{n}^{j}\cfrac{d^{j}e^{P(\psi+t\phi)}}{dt^{j}}w^{(n-j)}(t,x)+\cfrac{d^{n}e^{P(\psi+t\phi)}}{dt^{n}}w(t,x)\vspace{3mm}\\ =&\sum_{j=0}^{n-1}C_{n}^{j}\sum_{\tau\in\mathfrak{P}(j)}B_{j}^{\tau}P^{(\tau)}(\psi+t\phi)e^{P(\psi+t\phi)}w^{(n-j)}(t,x)\vspace{3mm}\\ &+\sum_{\tau\in\mathfrak{P}(n)}B_{n}^{\tau}P^{(\tau)}(\psi+t\phi)e^{P(\psi+t\phi)}w(t,x)\vspace{3mm}\\ =&\sum_{j=0}^{n-1}C_{n}^{j}\Big{(}\sum_{\tau\in\mathfrak{P}(j),1\notin\tau}B_{j}^{\tau}P^{(\tau)}(\psi+t\phi)+\sum_{\tau\in\mathfrak{P}(j),1\in\tau}B_{j}^{\tau}P^{(\tau)}(\psi+t\phi)\Big{)}e^{P(\psi+t\phi)}w^{(n-j)}(t,x)\vspace{3mm}\\ &+\sum_{\tau\in\mathfrak{P}(n),n\notin\tau}B_{n}^{\tau}P^{(\tau)}(\psi+t\phi)e^{P(\psi+t\phi)}w(t,x)+P^{(n)}(\psi+t\phi)e^{P(\psi+t\phi)}w(t,x).\vspace{3mm}\\ \end{array}

Remember P(ψ)=0P(\psi)=0 and w(0,x)=1w(0,x)=1 as ψ\psi is normalized ([PP, p66]). Take t=0t=0 in (3.7) we get

(3.8) dn(eP(ψ+tϕ)w(t,x))dtn|t=0=j=0n1Cnj(τ𝔓(j),1τBjτP(τ)(ψ+tϕ)|t=0+τ𝔓(j),1τBjτP(τ)(ψ+tϕ)|t=0)w(nj)(0,x)+τ𝔓(n),nτBnτP(τ)(ψ+tϕ)|t=0+P(n)(ψ+tϕ)|t=0\begin{array}[]{ll}&\cfrac{d^{n}\Big{(}e^{P(\psi+t\phi)}w(t,x)\Big{)}}{dt^{n}}|_{t=0}\vspace{1mm}\\ =&\sum_{j=0}^{n-1}C_{n}^{j}\Big{(}\sum_{\tau\in\mathfrak{P}(j),1\notin\tau}B_{j}^{\tau}P^{(\tau)}(\psi+t\phi)|_{t=0}+\sum_{\tau\in\mathfrak{P}(j),1\in\tau}B_{j}^{\tau}P^{(\tau)}(\psi+t\phi)|_{t=0}\Big{)}w^{(n-j)}(0,x)\vspace{3mm}\\ &+\sum_{\tau\in\mathfrak{P}(n),n\notin\tau}B_{n}^{\tau}P^{(\tau)}(\psi+t\phi)|_{t=0}+P^{(n)}(\psi+t\phi)|_{t=0}\vspace{3mm}\\ \end{array}

Since ϕ𝑑μψ=P(ψ+tϕ)|t=0=0\int\phi d\mu_{\psi}=P^{\prime}(\psi+t\phi)|_{t=0}=0 and w(0,x)𝑑μψ=0\int w^{\prime}(0,x)d\mu_{\psi}=0 ([PP, p66]), integrate both sides of (3.8) with respect to μψ\mu_{\psi}, we get

(3.9) dn(eP(ψ+tϕ)w(t,x))dtn|t=0dμψ=j=0n1Cnjτ𝔓(j),1τBjτP(τ)(ψ+tϕ)|t=0w(nj)(0,x)𝑑μψ+τ𝔓(n),{1,n}τ=BnτP(τ)(ψ+tϕ)|t=0+P(n)(ψ+tϕ)|t=0.\begin{array}[]{ll}&\displaystyle\int\cfrac{d^{n}\Big{(}e^{P(\psi+t\phi)}w(t,x)\Big{)}}{dt^{n}}|_{t=0}d\mu_{\psi}\vspace{1mm}\\ =&\sum_{j=0}^{n-1}C_{n}^{j}\sum_{\tau\in\mathfrak{P}(j),1\notin\tau}B_{j}^{\tau}P^{(\tau)}(\psi+t\phi)|_{t=0}\int w^{(n-j)}(0,x)d\mu_{\psi}\vspace{3mm}\\ &+\sum_{\tau\in\mathfrak{P}(n),\{1,n\}\cap\tau=\emptyset}B_{n}^{\tau}P^{(\tau)}(\psi+t\phi)|_{t=0}+P^{(n)}(\psi+t\phi)|_{t=0}.\vspace{3mm}\\ \end{array}

Finally, combining (3.6) and (3.9) together we get (3.2).

3.2 Remark.

The terms

j=2n2Cnjτ𝔓(j),1τBjτP(τ)(ψ+tϕ)|t=0Λw(nj)(0,x)𝑑μψ(x)-\sum_{j=2}^{n-2}C_{n}^{j}\sum_{\tau\in\mathfrak{P}(j),1\notin\tau}B_{j}^{\tau}P^{(\tau)}(\psi+t\phi)|_{t=0}\int_{\Lambda^{\mathbb{N}}}w^{(n-j)}(0,x)d\mu_{\psi}(x)

and

τ𝔓(n),{1,n}τ=BnτP(τ)(ψ+tϕ)|t=0-\sum_{\tau\in\mathfrak{P}(n),\{1,n\}\cap\tau=\emptyset}B_{n}^{\tau}P^{(\tau)}(\psi+t\phi)|_{t=0}

in (3.2) are null in case of n3n\leq 3. This also applies to the corresponding terms later.

3.3 Remark.

These appear to be inductive formulas, while one can always get non-inductive ones via substituting the lower differentials P(τ)(ψ+tϕ)|t=0P^{(\tau)}(\psi+t\phi)|_{t=0} by their non-inductive versions depending only on ϕ(x),{w(j)(0,x)}j=1n\phi(x),\{w^{(j)}(0,x)\}_{j=1}^{n} and μψ(x)\mu_{\psi}(x). This also applies to Theorem 3.7.

One can find some description of derivatives of the pressure function by covariance of the sequence of functions {ϕσj}j\{\phi\circ\sigma^{j}\}_{j\in\mathbb{N}} in [KS1, Corollary 1] for smooth ϕ\phi. Without the assumptions of ψ\psi being normalized and ϕ𝑑μψ=0\int\phi d\mu_{\psi}=0, Theorem 3.1 evolves into the following form.

3.4 Corollary.

Let ψ,ϕC0,h(Λ)\psi,\phi\in C^{0,h}(\Lambda^{\mathbb{N}}) with some finite symbolic set Λ\Lambda. ψ+tϕ\mathcal{L}_{\psi+t\phi} admits a maximum isolated eigenvalue eP(ψ+tϕ)e^{P(\psi+t\phi)} close to eP(ψ)e^{P(\psi)} with eigenfunction w(t,x)w(t,x) whose projection depends analytically on tt in a small neighbourhood of 0. Considering the differentials of the pressure P(ψ+tϕ)P(\psi+t\phi) at t=0t=0, we have

(3.10) P(n)(ψ+tϕ)|t=0=j=1nCnjΛ(ϕ(x)ϕ𝑑μψ)jw(nj)(0,x)𝑑μψ(x)j=2n2Cnjτ𝔓(j),1τBjτP(τ)(ψ+tϕ)|t=0Λw(nj)(0,x)𝑑μψ(x)τ𝔓(n),{1,n}τ=BnτP(τ)(ψ+tϕ)|t=0\begin{array}[]{ll}P^{(n)}(\psi+t\phi)|_{t=0}=&\sum_{j=1}^{n}C_{n}^{j}\displaystyle\int_{\Lambda^{\mathbb{N}}}\Big{(}\phi(x)-\int\phi d\mu_{\psi}\Big{)}^{j}w^{(n-j)}(0,x)d\mu_{\psi}(x)\vspace{3mm}\\ &-\sum_{j=2}^{n-2}C_{n}^{j}\sum_{\tau\in\mathfrak{P}(j),1\notin\tau}B_{j}^{\tau}P^{(\tau)}(\psi+t\phi)|_{t=0}\displaystyle\int_{\Lambda^{\mathbb{N}}}w^{(n-j)}(0,x)d\mu_{\psi}(x)\vspace{3mm}\\ &-\sum_{\tau\in\mathfrak{P}(n),\{1,n\}\cap\tau=\emptyset}B_{n}^{\tau}P^{(\tau)}(\psi+t\phi)|_{t=0}\end{array}

for any n2n\geq 2.

Proof.

Let

ψ¯=ψ+logwψ(x)logwψσP(ψ)\bar{\psi}=\psi+\log w_{\psi}(x)-\log w_{\psi}\circ\sigma-P(\psi)

in which wψ(x)w_{\psi}(x) is the eigenfunction of ψ\mathcal{L}_{\psi} corresponding to the eigenvalue eP(ψ)e^{P(\psi)}. Take pressure in the following equation

ψ¯+tϕ=ψ+tϕ+logwψ(x)logwψσP(ψ)\bar{\psi}+t\phi=\psi+t\phi+\log w_{\psi}(x)-\log w_{\psi}\circ\sigma-P(\psi),

then apply Proposition 2.2, we see that

P(ψ¯+tϕ)=P(ψ+tϕ)P(ψ)P(\bar{\psi}+t\phi)=P(\psi+t\phi)-P(\psi).

This implies

(3.11) dnP(ψ¯+tϕ)dtn=dnP(ψ+tϕ)dtn\cfrac{d^{n}P(\bar{\psi}+t\phi)}{dt^{n}}=\cfrac{d^{n}P(\psi+t\phi)}{dt^{n}}

for any n1n\geq 1. Now apply Theorem 3.1 to the normalized potential ψ¯\bar{\psi} and ϕϕ𝑑μψ\phi-\int\phi d\mu_{\psi}, (note that (ϕϕ𝑑μψ)𝑑μψ=0\int\big{(}\phi-\int\phi d\mu_{\psi}\big{)}d\mu_{\psi}=0 and μψ=μψ¯\mu_{\psi}=\mu_{\bar{\psi}}), we justify the corollary considering (3.11). ∎

In the following we present some concrete formulas of some special order nn in virtue of Theorem 3.1 for later use.

3.5 Corollary.

Let ψ,ϕC0,h(Λ)\psi,\phi\in C^{0,h}(\Lambda^{\mathbb{N}}) with ψ\psi being normalized. Let μψ\mu_{\psi} be the equilibrium state of ψ\psi and ϕ𝑑μψ=0\int\phi d\mu_{\psi}=0. Let eP(ψ+tϕ)e^{P(\psi+t\phi)} be the maximum eigenvalue of ψ+tϕ\mathcal{L}_{\psi+t\phi} with eigenfunction w(t,x)w(t,x) for small tt. Then we have

(3.12) P′′′(ψ+tϕ)|t=0=3ϕw′′(0,x)𝑑μψ+3ϕ2w(0,x)𝑑μψ+ϕ3𝑑μψ.P^{{}^{\prime\prime\prime}}(\psi+t\phi)|_{t=0}=3\int\phi w^{{}^{\prime\prime}}(0,x)d\mu_{\psi}+3\int\phi^{2}w^{{}^{\prime}}(0,x)d\mu_{\psi}+\int\phi^{3}d\mu_{\psi}.
Proof.

This follows instantly from Theorem 3.1 with n=3n=3, along with some direct computations on the Faà di Bruno’s coefficients {B3τ}τ𝔓(3)\{B_{3}^{\tau}\}_{\tau\in\mathfrak{P}(3)}. ∎

3.6 Corollary.

Let ψ,ϕC0,h(Λ)\psi,\phi\in C^{0,h}(\Lambda^{\mathbb{N}}) with ψ\psi being normalized. Let μψ\mu_{\psi} be the equilibrium state of ψ\psi and ϕ𝑑μψ=0\int\phi d\mu_{\psi}=0. Let eP(ψ+tϕ)e^{P(\psi+t\phi)} be the maximum eigenvalue of ψ+tϕ\mathcal{L}_{\psi+t\phi} with eigenfunction w(t,x)w(t,x) for small tt. Then we have

(3.13) P′′′′(ψ+tϕ)|t=0=4ϕw′′′(0,x)𝑑μψ+6ϕ2w′′(0,x)𝑑μψ+4ϕ3w(0,x)𝑑μψ+ϕ4𝑑μψ6P′′(ψ+tϕ)|t=0w′′(0,x)𝑑μψ3(P′′(ψ+tϕ)|t=0)2=4ϕw′′′(0,x)𝑑μψ+6ϕ2w′′(0,x)𝑑μψ+4ϕ3w(0,x)𝑑μψ+ϕ4𝑑μψ6(ϕ2𝑑μψ+2ϕw(0,x)𝑑μψ)w′′(0,x)𝑑μψ3(ϕ2𝑑μψ+2ϕw(0,x)𝑑μψ)2.\begin{array}[]{ll}&P^{{}^{\prime\prime\prime\prime}}(\psi+t\phi)|_{t=0}\vspace{3mm}\\ =&4\int\phi w^{{}^{\prime\prime\prime}}(0,x)d\mu_{\psi}+6\int\phi^{2}w^{{}^{\prime\prime}}(0,x)d\mu_{\psi}+4\int\phi^{3}w^{{}^{\prime}}(0,x)d\mu_{\psi}+\int\phi^{4}d\mu_{\psi}\vspace{3mm}\\ &-6P^{{}^{\prime\prime}}(\psi+t\phi)|_{t=0}\int w^{{}^{\prime\prime}}(0,x)d\mu_{\psi}-3(P^{{}^{\prime\prime}}(\psi+t\phi)|_{t=0})^{2}\vspace{3mm}\\ =&4\int\phi w^{{}^{\prime\prime\prime}}(0,x)d\mu_{\psi}+6\int\phi^{2}w^{{}^{\prime\prime}}(0,x)d\mu_{\psi}+4\int\phi^{3}w^{{}^{\prime}}(0,x)d\mu_{\psi}+\int\phi^{4}d\mu_{\psi}\vspace{3mm}\\ &-6(\int\phi^{2}d\mu_{\psi}+2\int\phi w^{\prime}(0,x)d\mu_{\psi})\int w^{{}^{\prime\prime}}(0,x)d\mu_{\psi}-3(\int\phi^{2}d\mu_{\psi}+2\int\phi w^{\prime}(0,x)d\mu_{\psi})^{2}.\end{array}
Proof.

The first equality follows instantly from Theorem 3.1 with n=4n=4 along with some direct computations on the Faà di Bruno’s coefficients {B4τ}τ𝔓(4)\{B_{4}^{\tau}\}_{\tau\in\mathfrak{P}(4)}. The second one is true as

P′′(ψ+tϕ)|t=0=ϕ2𝑑μψ+2ϕw(0,x)𝑑μψ.P^{{}^{\prime\prime}}(\psi+t\phi)|_{t=0}=\int\phi^{2}d\mu_{\psi}+2\int\phi w^{\prime}(0,x)d\mu_{\psi}.

The latter description depends only on ϕ(x),{w(j)(0,x)}j=13\phi(x),\{w^{(j)}(0,x)\}_{j=1}^{3} and μψ(x)\mu_{\psi}(x). ∎

One can also get some precise formulas for some particular nn in Corollary 3.4, and some non-inductive ones as we indicate in Remark 3.3. While the formulas (3.2, 3.10, 3.12, 3.13) all give interesting descriptions of the differentials of the pressure function P(ψ+tϕ)P(\psi+t\phi), it seems to us difficult to discover any essential rigid restriction on them, or relationships between them. In the following we turn to the description of them by the random stochastic process {ϕσj(x)}j=0\{\phi\circ\sigma^{j}(x)\}_{j=0}^{\infty}. This is not a new idea on exploring the regularity of the pressure function P(ψ+tϕ)P(\psi+t\phi), as one can recall it from many others’ work in thermodynamic formalism. Again we first consider some standard case, then extend to the general case.

3.7 Theorem.

Let ψ,ϕC0,h(Λ)\psi,\phi\in C^{0,h}(\Lambda^{\mathbb{N}}) with ψ\psi being normalized. Let μψ\mu_{\psi} be the equilibrium state of ψ\psi and ϕ𝑑μψ=0\int\phi d\mu_{\psi}=0. Let eP(ψ+tϕ)e^{P(\psi+t\phi)} be the maximum isolated eigenvalue of eP(ψ+tϕ)e^{P(\psi+t\phi)} with eigenfunction w(t,x)w(t,x) whose projection depends analytically on tt. Considering the differentials of the pressure P(ψ+tϕ)P(\psi+t\phi) at t=0t=0, we have

(3.14) P(n)(ψ+tϕ)|t=0=limm1m(j=2nCnjΛ(Sm,ϕ(x))jw(nj)(0,x)dμψ(x)j=2n2Cnjτ𝔓(j),1τm#τBjτP(τ)(ψ+tϕ)|t=0Λw(nj)(0,x)𝑑μψ(x)τ𝔓(n),{1,n}τ=m#τBnτP(τ)(ψ+tϕ)|t=0)\begin{array}[]{ll}&P^{(n)}(\psi+t\phi)|_{t=0}\\ =&\lim_{m\rightarrow\infty}\cfrac{1}{m}\Big{(}\sum_{j=2}^{n}C_{n}^{j}\displaystyle\int_{\Lambda^{\mathbb{N}}}(S_{m,\phi}(x))^{j}w^{(n-j)}(0,x)d\mu_{\psi}(x)\vspace{3mm}\\ &-\sum_{j=2}^{n-2}C_{n}^{j}\sum_{\tau\in\mathfrak{P}(j),1\notin\tau}m^{\#\tau}B_{j}^{\tau}P^{(\tau)}(\psi+t\phi)|_{t=0}\displaystyle\int_{\Lambda^{\mathbb{N}}}w^{(n-j)}(0,x)d\mu_{\psi}(x)\vspace{3mm}\\ &-\sum_{\tau\in\mathfrak{P}(n),\{1,n\}\cap\tau=\emptyset}m^{\#\tau}B_{n}^{\tau}P^{(\tau)}(\psi+t\phi)|_{t=0}\Big{)}\end{array}

for any n2n\geq 2.

Proof.

The proof follows the routine of Proof of Theorem 3.1. Considering (2.1), we do nn-differentials on both sides of (2.3), take t=0t=0, then integrate both sides with respect to μψ(x)\mu_{\psi}(x), divided by mm, we get

(3.15) P(n)(ψ+tϕ)|t=0=1m(j=1nCnjΛ(Sm,ϕ(x))jw(nj)(0,x)dμψ(x)j=2n2Cnjτ𝔓(j),1τm#τBjτP(τ)(ψ+tϕ)|t=0Λw(nj)(0,x)𝑑μψ(x)τ𝔓(n),{1,n}τ=m#τBnτP(τ)(ψ+tϕ)|t=0)\begin{array}[]{ll}&P^{(n)}(\psi+t\phi)|_{t=0}\\ =&\cfrac{1}{m}\Big{(}\sum_{j=1}^{n}C_{n}^{j}\displaystyle\int_{\Lambda^{\mathbb{N}}}(S_{m,\phi}(x))^{j}w^{(n-j)}(0,x)d\mu_{\psi}(x)\vspace{3mm}\\ &-\sum_{j=2}^{n-2}C_{n}^{j}\sum_{\tau\in\mathfrak{P}(j),1\notin\tau}m^{\#\tau}B_{j}^{\tau}P^{(\tau)}(\psi+t\phi)|_{t=0}\displaystyle\int_{\Lambda^{\mathbb{N}}}w^{(n-j)}(0,x)d\mu_{\psi}(x)\vspace{3mm}\\ &-\sum_{\tau\in\mathfrak{P}(n),\{1,n\}\cap\tau=\emptyset}m^{\#\tau}B_{n}^{\tau}P^{(\tau)}(\psi+t\phi)|_{t=0}\Big{)}\end{array}

as (3.2). Now since w(n1)(0,x)w^{(n-1)}(0,x) is bounded on XX, the ergodic theorem guarantees

(3.16) limm1mΛSm,ϕ(x)w(n1)(0,x)𝑑μψ(x)=0.\lim_{m\rightarrow\infty}\cfrac{1}{m}\int_{\Lambda^{\mathbb{N}}}S_{m,\phi}(x)w^{(n-1)}(0,x)d\mu_{\psi}(x)=0.

Then (3.14) follows from (3.15) as mm\rightarrow\infty considering (3.16). ∎

Theorem 3.7 establishes some link between the differentials of the pressure function and the the process {ϕσj(x)}j=0\{\phi\circ\sigma^{j}(x)\}_{j=0}^{\infty} through Sm,ϕS_{m,\phi} with respect to the equilibrium state μψ\mu_{\psi}. We also formulate a general version of the result.

3.8 Corollary.

Let ψ,ϕC0,h(Λ)\psi,\phi\in C^{0,h}(\Lambda^{\mathbb{N}}) with μψ\mu_{\psi} be the equilibrium state of ψ\psi. ψ+tϕ\mathcal{L}_{\psi+t\phi} admits a maximum isolated eigenvalue eP(ψ+tϕ)e^{P(\psi+t\phi)} close to eP(ψ)e^{P(\psi)} with eigenfunction w(t,x)w(t,x) whose projection depends analytically on tt in a small neighbourhood of 0. Considering the differentials of the pressure function P(ψ+tϕ)P(\psi+t\phi) at t=0t=0, we have

(3.17) P(n)(ψ+tϕ)|t=0=limm1m(j=2nCnjΛ(Sm,ϕmϕdμψ)jw(nj)(0,x)dμψ(x)j=2n2Cnjτ𝔓(j),1τm#τBjτP(τ)(ψ+tϕ)|t=0Λw(nj)(0,x)𝑑μψ(x)τ𝔓(n),{1,n}τ=m#τBnτP(τ)(ψ+tϕ)|t=0)\begin{array}[]{ll}&P^{(n)}(\psi+t\phi)|_{t=0}\\ =&\lim_{m\rightarrow\infty}\cfrac{1}{m}\Big{(}\sum_{j=2}^{n}C_{n}^{j}\displaystyle\int_{\Lambda^{\mathbb{N}}}\big{(}S_{m,\phi}-m\int\phi d\mu_{\psi}\big{)}^{j}w^{(n-j)}(0,x)d\mu_{\psi}(x)\vspace{3mm}\\ &-\sum_{j=2}^{n-2}C_{n}^{j}\sum_{\tau\in\mathfrak{P}(j),1\notin\tau}m^{\#\tau}B_{j}^{\tau}P^{(\tau)}(\psi+t\phi)|_{t=0}\displaystyle\int_{\Lambda^{\mathbb{N}}}w^{(n-j)}(0,x)d\mu_{\psi}(x)\vspace{3mm}\\ &-\sum_{\tau\in\mathfrak{P}(n),\{1,n\}\cap\tau=\emptyset}m^{\#\tau}B_{n}^{\tau}P^{(\tau)}(\psi+t\phi)|_{t=0}\Big{)}\end{array}

for any n2n\geq 2.

Proof.

Equipped with Theorem 3.7, the proof follows in line with the Proof of Corollary 3.4. ∎

Now we give some precise descriptions of the third and fourth differentials of P(ψ+tϕ)P(\psi+t\phi) in virtue of Theorem 3.7.

3.9 Corollary.

Let ψ,ϕC0,h(Λ)\psi,\phi\in C^{0,h}(\Lambda^{\mathbb{N}}) with ψ\psi being normalized. Let μψ\mu_{\psi} be the equilibrium state of ψ\psi and ϕ𝑑μψ=0\int\phi d\mu_{\psi}=0. Let eP(ψ+tϕ)e^{P(\psi+t\phi)} be the maximum eigenvalue of ψ+tϕ\mathcal{L}_{\psi+t\phi} with eigenfunction w(t,x)w(t,x) for small tt. Then we have

(3.18) P′′(ψ+tϕ)|t=0=limm1mSm,ϕ2𝑑μψ.P^{{}^{\prime\prime}}(\psi+t\phi)|_{t=0}=\lim_{m\rightarrow\infty}\cfrac{1}{m}\int S_{m,\phi}^{2}d\mu_{\psi}.
3.10 Remark.

P′′(ψ+tϕ)|t=0P^{{}^{\prime\prime}}(\psi+t\phi)|_{t=0} is called variance of the random process {ϕσj(x)}j=0\{\phi\circ\sigma^{j}(x)\}_{j=0}^{\infty}, whose name can be interpreted from the Central Limit Theorem. See [Rue1, PP].

3.11 Corollary.

Let ψ,ϕC0,h(Λ)\psi,\phi\in C^{0,h}(\Lambda^{\mathbb{N}}) with ψ\psi being normalized. Let μψ\mu_{\psi} be the equilibrium state of ψ\psi and ϕ𝑑μψ=0\int\phi d\mu_{\psi}=0. Let eP(ψ+tϕ)e^{P(\psi+t\phi)} be the maximum eigenvalue of ψ+tϕ\mathcal{L}_{\psi+t\phi} with eigenfunction w(t,x)w(t,x) for small tt. Then we have

(3.19) P′′′(ψ+tϕ)|t=0=limm3mSm,ϕ2w(0,x)𝑑μψ+limm1mSm,ϕ3𝑑μψ.P^{{}^{\prime\prime\prime}}(\psi+t\phi)|_{t=0}=\lim_{m\rightarrow\infty}\cfrac{3}{m}\int S_{m,\phi}^{2}w^{{}^{\prime}}(0,x)d\mu_{\psi}+\lim_{m\rightarrow\infty}\cfrac{1}{m}\int S_{m,\phi}^{3}d\mu_{\psi}.
Proof.

This follows instantly from Theorem 3.7 with n=3n=3. ∎

3.12 Corollary.

Let ψ,ϕC0,h(Λ)\psi,\phi\in C^{0,h}(\Lambda^{\mathbb{N}}) with ψ\psi being normalized. Let μψ\mu_{\psi} be the equilibrium state of ψ\psi and ϕ𝑑μψ=0\int\phi d\mu_{\psi}=0. Let eP(ψ+tϕ)e^{P(\psi+t\phi)} be the maximum eigenvalue of ψ+tϕ\mathcal{L}_{\psi+t\phi} with eigenfunction w(t,x)w(t,x) for small tt. Then we have

(3.20) P(4)(ψ+tϕ)|t=0=limm(6mSm,ϕ2w′′(0,x)dμψ+4mSm,ϕ3w(0,x)dμψ+1mSm,ϕ4dμψ6P′′(ψ+tϕ)|t=0w′′(0,x)dμψ3m(P′′(ψ+tϕ)|t=0)2)=limm(6mSm,ϕ2w′′(0,x)dμψ+4mSm,ϕ3w(0,x)dμψ+1mSm,ϕ4dμψ6mSm,ϕ2dμψw′′(0,x)dμψ3m(Sm,ϕ2dμψ)2).\begin{array}[]{ll}&P^{(4)}(\psi+t\phi)|_{t=0}\vspace{3mm}\\ =&\lim_{m\rightarrow\infty}\Big{(}\cfrac{6}{m}\displaystyle\int S_{m,\phi}^{2}w^{{}^{\prime\prime}}(0,x)d\mu_{\psi}+\cfrac{4}{m}\int S_{m,\phi}^{3}w^{{}^{\prime}}(0,x)d\mu_{\psi}+\cfrac{1}{m}\int S_{m,\phi}^{4}d\mu_{\psi}\vspace{3mm}\\ &-6P^{{}^{\prime\prime}}(\psi+t\phi)|_{t=0}\displaystyle\int w^{{}^{\prime\prime}}(0,x)d\mu_{\psi}-3m(P^{{}^{\prime\prime}}(\psi+t\phi)|_{t=0})^{2}\Big{)}\vspace{3mm}\\ =&\lim_{m\rightarrow\infty}\Big{(}\cfrac{6}{m}\displaystyle\int S_{m,\phi}^{2}w^{{}^{\prime\prime}}(0,x)d\mu_{\psi}+\cfrac{4}{m}\int S_{m,\phi}^{3}w^{{}^{\prime}}(0,x)d\mu_{\psi}+\cfrac{1}{m}\int S_{m,\phi}^{4}d\mu_{\psi}\vspace{3mm}\\ &-\cfrac{6}{m}\displaystyle\int S_{m,\phi}^{2}d\mu_{\psi}\displaystyle\int w^{{}^{\prime\prime}}(0,x)d\mu_{\psi}-\cfrac{3}{m}(\int S_{m,\phi}^{2}d\mu_{\psi})^{2}\Big{)}\vspace{3mm}\\ .\end{array}
Proof.

The first equality follows instantly from Theorem 3.7 with n=4n=4, while the second one is true considering (3.18). The last description depends only on ϕ(x),{w(j)(0,x)}j=12\phi(x),\{w^{(j)}(0,x)\}_{j=1}^{2} and μψ(x)\mu_{\psi}(x). ∎

Through the above formulas we see the importance of the asymptotic distribution of the random variable Sm,ϕS_{m,\phi} with respect to μψ\mu_{\psi}, which is describe by the Central Limit Theorem for the process {ϕσj(x)}j=0\{\phi\circ\sigma^{j}(x)\}_{j=0}^{\infty}. Equipped with all the above results, now we are in a position to prove the rigidity results on the third differentials of P(ψ+tϕ)P(\psi+t\phi) upon Corollary 3.11. We first show Theorem 1.2.

Proof of Theorem 1.2.

From now on we fix t(0,)t_{*}\in(0,\infty). Let ψ=tϕ\psi=t_{*}\phi. Simply by making a change of variable we can see that

P(n)(t)=P(n)(tϕ)|t=t=P(n)(ψ+tϕ)|t=0P^{(n)}(t_{*})=P^{(n)}(t\phi)|_{t=t_{*}}=P^{(n)}(\psi+t\phi)|_{t=0}

for any n0n\geq 0. So (1.2) is equivalent to

(3.21) |P′′′(ψ+tϕ)|t=0(12π(P′′(ψ+tϕ)|t=0)3/2)|3MϕP′′(ψ+tϕ)|t=0.|P^{\prime\prime\prime}(\psi+t\phi)|_{t=0}\big{(}1-\sqrt{2\pi}(P^{\prime\prime}(\psi+t\phi)|_{t=0})^{3/2}\big{)}|\leq 3M_{\phi}P^{\prime\prime}(\psi+t\phi)|_{t=0}.

We can assume ψ\psi is normalized as otherwise we can change it to its normalization considering (2.10). Moreover, it suffices for us to prove it under the assumption ϕ𝑑μψ=0\int\phi d\mu_{\psi}=0 in virtue of (2.9). If P′′(ψ+tϕ)|t=0=0P^{\prime\prime}(\psi+t\phi)|_{t=0}=0, then ϕ\phi is cohomologous to a constant according to [PP, Proposition 4.12]. This forces P′′′(ψ+tϕ)|t=0=0P^{\prime\prime\prime}(\psi+t\phi)|_{t=0}=0, so (3.21) is satisfied in this case. In the following we assume P′′(ψ+tϕ)|t=0>0P^{\prime\prime}(\psi+t\phi)|_{t=0}>0. We resort to Corollary 3.11 to justify (3.21) under the above assumptions. We first estimate the term 1mSm,ϕ3𝑑μψ\cfrac{1}{m}\int S_{m,\phi}^{3}d\mu_{\psi} in (3.19). Now the Central Limit Theorem comes in.

1mSm,ϕ3𝑑μψ=m(Sm,ϕm)3𝑑μψ=my3𝑑Gm(y)=my3𝑑N0,P′′(ψ+tϕ)|t=0(y)+my3𝑑Hm(y)+mo(1/m)=m0+y3d(P′′′(ψ+tϕ)|t=06(1y2P′′(ψ+tϕ)|t=0)ey2/2P′′(ψ+tϕ)|t=0)+mo(1/m)=P′′′(ψ+tϕ)|t=02π(P′′(ψ+tϕ)|t=0)3/2+mo(1/m).\begin{array}[]{ll}&\cfrac{1}{m}\displaystyle\int S_{m,\phi}^{3}d\mu_{\psi}\vspace{3mm}\\ =&\sqrt{m}\displaystyle\int\Big{(}\cfrac{S_{m,\phi}}{\sqrt{m}}\Big{)}^{3}d\mu_{\psi}\vspace{3mm}\\ =&\sqrt{m}\displaystyle\int y^{3}dG_{m}(y)\vspace{3mm}\\ =&\sqrt{m}\displaystyle\int y^{3}dN_{0,P^{\prime\prime}(\psi+t\phi)|_{t=0}}(y)+\sqrt{m}\int y^{3}dH_{m}(y)+\sqrt{m}\cdot o(1/\sqrt{m})\vspace{3mm}\\ =&\sqrt{m}\cdot 0+\displaystyle\int y^{3}d\Big{(}\cfrac{P^{\prime\prime\prime}(\psi+t\phi)|_{t=0}}{6}\big{(}1-\cfrac{y^{2}}{P^{\prime\prime}(\psi+t\phi)|_{t=0}}\big{)}e^{-y^{2}/2P^{\prime\prime}(\psi+t\phi)|_{t=0}}\Big{)}\\ &+\sqrt{m}\cdot o(1/\sqrt{m})\vspace{3mm}\\ =&P^{\prime\prime\prime}(\psi+t\phi)|_{t=0}\sqrt{2\pi}(P^{\prime\prime}(\psi+t\phi)|_{t=0})^{3/2}+\sqrt{m}\cdot o(1/\sqrt{m}).\vspace{3mm}\\ \end{array}

By taking mm\rightarrow\infty we get

(3.22) limm1mSm,ϕ3𝑑μψ=P′′′(ψ+tϕ)|t=02π(P′′(ψ+tϕ)|t=0)3/2.\lim_{m\rightarrow\infty}\cfrac{1}{m}\displaystyle\int S_{m,\phi}^{3}d\mu_{\psi}=P^{\prime\prime\prime}(\psi+t\phi)|_{t=0}\sqrt{2\pi}(P^{\prime\prime}(\psi+t\phi)|_{t=0})^{3/2}.

Considering (3.19) we have

(3.23) P′′′(ψ+tϕ)|t=0(12π(P′′(ψ+tϕ)|t=0)3/2)=limm3mSm,ϕ2w(0,x)𝑑μψ.P^{\prime\prime\prime}(\psi+t\phi)|_{t=0}\big{(}1-\sqrt{2\pi}(P^{\prime\prime}(\psi+t\phi)|_{t=0})^{3/2}\big{)}=\lim_{m\rightarrow\infty}\cfrac{3}{m}\int S_{m,\phi}^{2}w^{{}^{\prime}}(0,x)d\mu_{\psi}.

Since w(0,x)w^{{}^{\prime}}(0,x) depends continuously on xXx\in X, there exists some MϕM_{\phi} depending on ϕ\phi, such that

(3.24) |w(0,x)|Mϕ.|w^{{}^{\prime}}(0,x)|\leq M_{\phi}.

Now taking absolute values on both sides of (3.23) we justify (3.21), considering (3.24) and (3.18).

The proof of Theorem 1.1 on the pressure functions of non-generic Hölder potentials follows a similar way.

Proof of Theorem 1.1.

Fix t(0,)t_{*}\in(0,\infty), we can simply assume ψ=tϕ\psi=t_{*}\phi is normalised and ϕ𝑑μψ=0\int\phi d\mu_{\psi}=0. In case that P′′(ψ+tϕ)|t=0=0P^{\prime\prime}(\psi+t\phi)|_{t=0}=0, so ϕ\phi is cohomologous to a constant, (1.1) holds obviously. In the following we assume ϕ\phi is not cohomologous to a constant, so P′′(ψ+tϕ)|t=0>0P^{\prime\prime}(\psi+t\phi)|_{t=0}>0. We again resort to Corollary 3.11 to justify (1.1) under these assumptions. Now for the term 1mSm,ϕ3𝑑μψ\cfrac{1}{m}\int S_{m,\phi}^{3}d\mu_{\psi} in (3.19), in virtue of the Central Limit Theorem,

(3.25) 1mSm,ϕ3𝑑μψ=m(Sm,ϕm)3𝑑μψ=my3𝑑Gm(y)my3𝑑N0,P′′(ψ+tϕ)|t=0(y)+m9|P′′′(ψ+tϕ)|t=0|+2|P(4)(ψ+tϕ)|t=0|2π3m(P′′(ψ+tϕ)|t=0)3/2=m0+9|P′′′(ψ+tϕ)|t=0|+2|P(4)(ψ+tϕ)|t=0|2π3(P′′(ψ+tϕ)|t=0)3/2=9|P′′′(ψ+tϕ)|t=0|+2|P(4)(ψ+tϕ)|t=0|2π3(P′′(ψ+tϕ)|t=0)3/2\begin{array}[]{ll}&\cfrac{1}{m}\displaystyle\int S_{m,\phi}^{3}d\mu_{\psi}\vspace{3mm}\\ =&\sqrt{m}\displaystyle\int\Big{(}\cfrac{S_{m,\phi}}{\sqrt{m}}\Big{)}^{3}d\mu_{\psi}\vspace{3mm}\\ =&\sqrt{m}\displaystyle\int y^{3}dG_{m}(y)\vspace{3mm}\\ \leq&\sqrt{m}\displaystyle\int y^{3}dN_{0,P^{\prime\prime}(\psi+t\phi)|_{t=0}}(y)+\sqrt{m}\cfrac{9|P^{\prime\prime\prime}(\psi+t\phi)|_{t=0}|+2|P^{(4)}(\psi+t\phi)|_{t=0}|}{\sqrt{2\pi^{3}m}(P^{\prime\prime}(\psi+t\phi)|_{t=0})^{3/2}}\vspace{3mm}\\ =&\sqrt{m}\cdot 0+\cfrac{9|P^{\prime\prime\prime}(\psi+t\phi)|_{t=0}|+2|P^{(4)}(\psi+t\phi)|_{t=0}|}{\sqrt{2\pi^{3}}(P^{\prime\prime}(\psi+t\phi)|_{t=0})^{3/2}}\vspace{3mm}\\ =&\cfrac{9|P^{\prime\prime\prime}(\psi+t\phi)|_{t=0}|+2|P^{(4)}(\psi+t\phi)|_{t=0}|}{\sqrt{2\pi^{3}}(P^{\prime\prime}(\psi+t\phi)|_{t=0})^{3/2}}\\ \end{array}

for mm large enough. By taking mm\rightarrow\infty in (3.25), we get

(3.26) limm1mSm,ϕ3𝑑μψ9|P′′′(ψ+tϕ)|t=0|+2|P(4)(ψ+tϕ)|t=0|2π3(P′′(ψ+tϕ)|t=0)3/2.\lim_{m\rightarrow\infty}\cfrac{1}{m}\displaystyle\int S_{m,\phi}^{3}d\mu_{\psi}\leq\cfrac{9|P^{\prime\prime\prime}(\psi+t\phi)|_{t=0}|+2|P^{(4)}(\psi+t\phi)|_{t=0}|}{\sqrt{2\pi^{3}}(P^{\prime\prime}(\psi+t\phi)|_{t=0})^{3/2}}.

Taking modulus on both sides of (3.26) we get

(3.27) |P′′′(ψ+tϕ)|t=0||limm1mSm,ϕ3𝑑μψ|+|limm3mSm,ϕ2w(0,x)𝑑μψ|9|P′′′(ψ+tϕ)|t=0|+2|P(4)(ψ+tϕ)|t=0|2π3(P′′(ψ+tϕ)|t=0)3/2+3Mϕ|limm3mSm,ϕ2𝑑μψ|=9|P′′′(ψ+tϕ)|t=0|+2|P(4)(ψ+tϕ)|t=0|2π3(P′′(ψ+tϕ)|t=0)3/2+3MϕP′′(ψ+tϕ)|t=0\begin{array}[]{ll}&|P^{\prime\prime\prime}(\psi+t\phi)|_{t=0}|\vspace{3mm}\\ \leq&\Big{|}\lim_{m\rightarrow\infty}\cfrac{1}{m}\displaystyle\int S_{m,\phi}^{3}d\mu_{\psi}\Big{|}+\Big{|}\lim_{m\rightarrow\infty}\cfrac{3}{m}\int S_{m,\phi}^{2}w^{{}^{\prime}}(0,x)d\mu_{\psi}\Big{|}\vspace{3mm}\\ \leq&\cfrac{9|P^{\prime\prime\prime}(\psi+t\phi)|_{t=0}|+2|P^{(4)}(\psi+t\phi)|_{t=0}|}{\sqrt{2\pi^{3}}(P^{\prime\prime}(\psi+t\phi)|_{t=0})^{3/2}}+3M_{\phi}\Big{|}\lim_{m\rightarrow\infty}\cfrac{3}{m}\displaystyle\int S_{m,\phi}^{2}d\mu_{\psi}\Big{|}\vspace{3mm}\\ =&\cfrac{9|P^{\prime\prime\prime}(\psi+t\phi)|_{t=0}|+2|P^{(4)}(\psi+t\phi)|_{t=0}|}{\sqrt{2\pi^{3}}(P^{\prime\prime}(\psi+t\phi)|_{t=0})^{3/2}}+3M_{\phi}P^{\prime\prime}(\psi+t\phi)|_{t=0}\end{array}

for some |w(0,x)|Mϕ|w^{{}^{\prime}}(0,x)|\leq M_{\phi}, which results in (1.1). ∎

One can predict from Corollary 3.12, Theorem 3.7 and the proof of Theorem 1.1, Theorem 1.2 that some more rigid relationships between higher differentials of the pressure function {P(n)(tϕ)}n\{P^{(n)}(t\phi)\}_{n\in\mathbb{N}} are possible. These rigidity relationships impose restrictions on fitting convex analytic functions whose supporting lines intersecting the vertical axis in some bounded set in [0,)[0,\infty) by pressures of Hölder potentials.

4. Global Fitting of convex analytic functions via pressures of Hölder potentials

This section is dedicated to the proof of Theorem 1.7. We start from the following result on some global behaviour of the pressure functions of generic Hölder potentials.

4.1 Theorem.

Let α>0\alpha>0. If a strictly convex analytic function F(t)F(t) on (α,)(\alpha,\infty), with its supporting lines intersecting the vertical axis in [γ¯,γ¯][0,)[\underline{\gamma},\overline{\gamma}]\subset[0,\infty), such that

(4.1) supt(α,){|F′′′(t)F′′(t)2πF′′(t)|}=,\sup_{t\in(\alpha,\infty)}\Bigg{\{}\Big{|}\cfrac{F^{{}^{\prime\prime\prime}}(t)}{F^{{}^{\prime\prime}}(t)}-\sqrt{2\pi F^{{}^{\prime\prime}}(t)}\Big{|}\Bigg{\}}=\infty,

then there does not exist any generic Hölder potential ϕ\phi on any shift space XX of finite type satisfying

P(tϕ)=F(t)P(t\phi)=F(t)

on (α,)(\alpha,\infty).

Proof.

This follows directly from Theorem 1.2 in fact. Suppose on the contrary that there exist some shift space XX of finite type and some generic Hölder potential ϕC0,h(X)\phi\in C^{0,h}(X) satisfying P(tϕ)=F(t)P(t\phi)=F(t) on (α,)(\alpha,\infty), then according to Theorem 1.2, we have

supt(α,){|F′′′(t)F′′(t)2πF′′(t)|}3Mϕ\sup_{t\in(\alpha,\infty)}\Bigg{\{}\Big{|}\cfrac{F^{{}^{\prime\prime\prime}}(t)}{F^{{}^{\prime\prime}}(t)}-\sqrt{2\pi F^{{}^{\prime\prime}}(t)}\Big{|}\Bigg{\}}\leq 3M_{\phi}

for some finite Mϕ>0M_{\phi}>0. This contradicts (4.1). ∎

Be careful that we cannot exclude the possibility that one can locally fit some convex analytic function through the pressure of some generic Hölder potential on some shift space of finite type by Theorem 1.2. This is because for any strictly convex analytic function F(t)F(t) on (α,)(\alpha,\infty) and αα¯α¯\alpha\leq\underline{\alpha}\leq\overline{\alpha}, we always have

supα¯tα¯{|F′′′(t)F′′(t)2πF′′(t)|}<\sup_{\underline{\alpha}\leq t\leq\overline{\alpha}}\Bigg{\{}\Big{|}\cfrac{F^{{}^{\prime\prime\prime}}(t)}{F^{{}^{\prime\prime}}(t)}-\sqrt{2\pi F^{{}^{\prime\prime}}(t)}\Big{|}\Bigg{\}}<\infty.

So one cannot exclude the possibility that there exists some generic Hölder potential ϕ\phi on some shift space of finite type satisfying

P(tϕ)=F(t)P(t\phi)=F(t)

on [α¯,α¯][\underline{\alpha},\overline{\alpha}] through Theorem 1.2. See Section 5 for more results on the problem of local fitting of some convex analytic functions through the pressures of Hölder potentials.

Now for α>0\alpha>0, let

α={F(t):F(t) is a strictly convex analytic function on (α,) satisfying (4.1), its supporting lines intersect the vertical axis in a bounded interval in [0,)}.\begin{array}[]{ll}\mathcal{F}_{\alpha}=&\{F(t):F(t)\mbox{ is a strictly convex analytic function on }(\alpha,\infty)\mbox{ satisfying }(\ref{eq26}),\vspace{3mm}\\ &\mbox{ its supporting lines intersect the vertical axis in a bounded interval in }[0,\infty)\}.\vspace{3mm}\end{array}

We will show that α\mathcal{F}_{\alpha}\neq\emptyset for any α>0\alpha>0 in the following.

4.2 Proposition.

For any α>0\alpha>0, we have

~α={Fa,b,c(t)=at2+bt+tect2+ect2t|(α,)}a,b>0,c>1/22α\tilde{\mathcal{F}}_{\alpha}=\Big{\{}F_{a,b,c}(t)=\cfrac{at^{2}+bt+te^{-ct^{2}}+e^{-ct^{2}}}{t}\Big{|}_{(\alpha,\infty)}\Big{\}}_{a,b>0,c>1/2\sqrt{2}}\subset\mathcal{F}_{\alpha}.

Proof.

The restricted functions on (α,)(\alpha,\infty) are of course analytic. Considering the second derivative of a function Fa,b,c(t)~αF_{a,b,c}(t)\in\tilde{\mathcal{F}}_{\alpha}, we have

Fa,b,c′′(t)=4c2t2ect2+4c2tect22cect2+2ct1ect2+2t3ect2F_{a,b,c}^{{}^{\prime\prime}}(t)=4c^{2}t^{2}e^{-ct^{2}}+4c^{2}te^{-ct^{2}}-2ce^{-ct^{2}}+2ct^{-1}e^{-ct^{2}}+2t^{-3}e^{-ct^{2}}

for t(0,)t\in(0,\infty). Now since

4c2t+2ct128c3>2c4c^{2}t+2ct^{-1}\geq 2\sqrt{8c^{3}}>2c

considering c>1/22c>1/2\sqrt{2}, we can see that Fa,b,c′′(t)>0F_{a,b,c}^{{}^{\prime\prime}}(t)>0 on (0,)(0,\infty). This shows that for any α>0\alpha>0, Fa,b,c(t)~αF_{a,b,c}(t)\in\tilde{\mathcal{F}}_{\alpha} is a convex function. Considering the third differential of a function Fa,b,c(t)~αF_{a,b,c}(t)\in\tilde{\mathcal{F}}_{\alpha}, we have

Fa,b,c′′′(t)=8c3t3ect28c3t2ect2+12c2tect26ct2ect26t4ect2F_{a,b,c}^{{}^{\prime\prime\prime}}(t)=-8c^{3}t^{3}e^{-ct^{2}}-8c^{3}t^{2}e^{-ct^{2}}+12c^{2}te^{-ct^{2}}-6ct^{-2}e^{-ct^{2}}-6t^{-4}e^{-ct^{2}}

for t(0,)t\in(0,\infty). Then we have

limt(Fa,b,c′′′(t)Fa,b,c′′(t)2πFa,b,c′′(t))=limt8c3t3ect24c2t2ect2=\lim_{t\rightarrow\infty}\Bigg{(}\cfrac{F^{{}^{\prime\prime\prime}}_{a,b,c}(t)}{F^{{}^{\prime\prime}}_{a,b,c}(t)}-\sqrt{2\pi F^{{}^{\prime\prime}}_{a,b,c}(t)}\Bigg{)}=\lim_{t\rightarrow\infty}\cfrac{-8c^{3}t^{3}e^{-ct^{2}}}{4c^{2}t^{2}e^{-ct^{2}}}=-\infty.

This means that Fa,b,c(t)~αF_{a,b,c}(t)\in\tilde{\mathcal{F}}_{\alpha} satisfies (4.1). To see that the supporting lines of a function Fa,b,c(t)~αF_{a,b,c}(t)\in\tilde{\mathcal{F}}_{\alpha} intersect the vertical axis in a bounded domain in [0,)[0,\infty), write the function as

Fa,b,c(t)=at+b+ect2+t1ect2F_{a,b,c}(t)=at+b+e^{-ct^{2}}+t^{-1}e^{-ct^{2}}.

Its graph on (0,)(0,\infty) is a strictly convex smooth curve with asymptotes y=at+by=at+b and t=0t=0. ∎

In Figure 1 we provide the readers with the graph of the function

F2,3,1(t)=2t2+3t+tet2+et2tF_{2,3,1}(t)=\cfrac{2t^{2}+3t+te^{-t^{2}}+e^{-t^{2}}}{t}

on (0,)(0,\infty).

Refer to caption
Figure 1. Graph of F2,3,1(t)F_{2,3,1}(t)

This means that any function in the family ~α\tilde{\mathcal{F}}_{\alpha} cannot be fitted by any generic Hölder potential on any shift space of finite type globally, considering Theorem 4.1. In the following we deny the possibility that they can be fitted by non-generic Hölder potentials on shift spaces of finite type.

4.3 Definition.

A continuous potential ϕ:X\phi:X\to\mathbb{R} on a shift space XX of finite type is said to be non-generic if for some normalised potential ψ\psi, the spectral radius of the complex Ruelle operator ψ+itϕ\mathcal{L}_{\psi+it\phi} equals 11 for some t0t\neq 0.

One can show that if ϕ\phi is non-generic then there exists a continuous function u:Xu:X\to\mathbb{R}, cϕc_{\phi}\in\mathbb{R} and a locally constant potential ϕ~:X\tilde{\phi}:X\to\mathbb{R}, such that

(4.2) ϕ(x)=uσ(x)u(x)+cϕ+ϕ~(x).\phi(x)=u\circ\sigma(x)-u(x)+c_{\phi}+\tilde{\phi}(x).
4.4 Proposition.

For any α>0\alpha>0 and any Fa,b,c(t)~αF_{a,b,c}(t)\in\tilde{\mathcal{F}}_{\alpha} with a,b>0,c>122a,b>0,c>\frac{1}{2\sqrt{2}}, there does not exist any non-generic Hölder potential ϕ\phi on any shift space of finite type such that

P(tϕ)=F(t)P(t\phi)=F(t)

on (α,)(\alpha,\infty).

Proof.

Note that for a non-generic Hölder potential ϕ\phi on a shift space of finite type, according to (4.2), we have

P(tϕ)=tcϕ+P(tϕ~)P(t\phi)=tc_{\phi}+P(t\tilde{\phi}),

in which ϕ~\tilde{\phi} is some locally constant potential. By the explicit formula (see Lemma 5.3) for the pressure functions of locally constant potentials on shift spaces of finite type, we see that any Fa,b,c(t)F_{a,b,c}(t) cannot be fitted by pressure of any non-generic Hölder potential ϕ\phi globally. ∎

Equipped with all the above results, Theorem 1.7 follows instantly from Proposition 4.2 and 4.4.

5. Local fitting of analytic germs via pressures of locally constant potentials

In this section we deal with the local fitting of analytic functions by the pressures of Hölder potentials, especially the pressures of piecewise constant ones. Firstly we borrow some notion originating from analytic continuation.

5.1 Definition.

A germ at tt_{*} is the sum of infinite power series

g(t)=a0+a1(tt)+a22!(tt)2+a33!(tt)3+g(t)=a_{0}+a_{1}(t-t_{*})+\cfrac{a_{2}}{2!}(t-t_{*})^{2}+\cfrac{a_{3}}{3!}(t-t_{*})^{3}+\cdots

for some (a0,a1,)(a_{0},a_{1},\cdots)\in\mathbb{R}^{\infty}.

The convergent radius (the superior of values δ0\delta\geq 0 on [tδ,t+δ][t_{*}-\delta,t_{*}+\delta] the germ converges) of the power series is called the radius of the germ. We are only interested in germs of radius δ>0\delta>0. The following problem will be our concern in this section.

5.2 Problem.

For a germ

g(t)=a0+a1(tt)+a22!(tt)2+g(t)=a_{0}+a_{1}(t-t_{*})+\cfrac{a_{2}}{2!}(t-t_{*})^{2}+\cdots

at tt_{*} with some strictly positive radius, does there exist some Hölder potential ϕ\phi on some shift space of finite type and some δ>0\delta>0, such that

P(tϕ)=g(t)P(t\phi)=g(t)

on [tδ,t+δ][t_{*}-\delta,t_{*}+\delta]?

The question can still be understood in Katok’s flexibility program in the class of symbolic dynamical systems, or even in some smooth systems. Obvious conditions to guarantee a positive answer to the problem are (1.5) and

(5.1) a2>0.a_{2}>0.

Condition (5.1) guarantees convexity of the germ (in some neighbourhood of tt_{*}) while (1.5) guarantees the supporting lines of the germ intersect the vertical axis in a bounded set in [0,)[0,\infty) (also in some neighbourhood of tt_{*}). We are especially interested in its answer when the Hölder potential in Problem 5.2 is required to be a piecewise constant one. We have seen the importance of the family of locally constant potentials in approximating convex analytic functions in Corollary 1.4. In fact Corollary 1.4 has some interesting interpretation in approximation theory [Tim], when we consider the explicit expressions of the pressures of locally constant potentials on the shift space of finite type. For nn\in\mathbb{N}, recall that

Λn={1,2,,n}\Lambda_{n}=\{1,2,\cdots,n\}.

5.3 Lemma.

For an integer k0k\geq 0, consider some locally constant potential

ϕ(x)=cxkxk+1x0xk1xk\phi(x)=c_{x_{-k}x_{-k+1}\cdots x_{0}\cdots x_{k-1}x_{k}}

for x=x1x0x1[xkxk]x=\cdots x_{-1}x_{0}x_{1}\cdots\in[x_{-k}\cdots x_{k}] on the shift space Λn\Lambda_{n}^{\mathbb{Z}}, we have

P(tϕ)=log(xk,,xk)Λn2k+1etcxkxkP(t\phi)=\log\sum_{(x_{-k},\cdots,x_{k})\in\Lambda_{n}^{2k+1}}e^{tc_{x_{-k}\cdots x_{k}}}

for any t(,)t\in(-\infty,\infty).

Proof.

This follows from [Wal1, Theorem 9.6] by some direct calculations through Definition 2.1 of the pressure. See also [Wal1, p214]. ∎

5.4 Remark.

The result can be extended to transitive subshifts of finite type. In this case the pressure is the logarithm of the maximal eigenvalue of some appropriate matrix.

Now combining Corollary 1.4 and Lemma 5.3, we have the following result.

5.5 Corollary.

Let F(t)F(t) be a convex Lipschitz function on (α,)(\alpha,\infty) for some α>0\alpha>0, such that its supporting lines intersect the vertical axis in [γ¯,γ¯][\underline{\gamma},\overline{\gamma}] with 0γ¯γ¯<0\leq\underline{\gamma}\leq\overline{\gamma}<\infty. Then there exists some KK\in\mathbb{N} and some sequences of constants

{cn,j}j=1Kn\{c_{n,j}\}_{j=1}^{K^{n}},

such that

(5.2) limnlogj=1Knetcn,j=F(t)\lim_{n\rightarrow\infty}\log\sum_{j=1}^{K^{n}}e^{tc_{n,j}}=F(t)

for any t(α,)t\in(\alpha,\infty).

Proof.

Take K=#ΛK=\#\Lambda for the symbolic set in the proof of Corollary 1.4, then the locally constant potential ϕn(x)=ϕn,(x)\phi_{n}(x)=\phi_{n,-}(x) admits KnK^{n} constant values respectively on corresponding level-nn cylinder sets. Denote these values by {cn,j}j=1Kn\{c_{n,j}\}_{j=1}^{K^{n}} for nn\in\mathbb{N}. According to Lemma 5.3,

P(tϕn,)=logj=1Knetcn,jP(t\phi_{n,-})=\log\sum_{j=1}^{K^{n}}e^{tc_{n,j}}

for any n1n\geq 1. This gives (5.2) by virtue of (1.3).

Corollary 5.5 indicates that logarithm of the finite sums of the exponential maps in the family {etc}c\{e^{tc}\}_{c\in\mathbb{R}} are dense in the space of certain convex Lipschitz maps on (α,)(\alpha,\infty). The above approximation is uniform with respect to tt in a bounded set. This makes the family {etc}c\{e^{tc}\}_{c\in\mathbb{R}} (family of locally constant potentials) important in detecting the properties of certain convex Lipschitz maps (among continuous or Hölder potentials).

From now on we turn our attention to Problem 5.2, but with restriction on locally constant potentials. We focus on locally constant potentials defined on the level-0 cylinder sets, whose theory is presumably parallel to the ones defined on the deeper cylinder sets. On the shift space Λn\Lambda_{n}^{\mathbb{Z}} with n2n\geq 2, consider the locally constant potential

ϕ(x)=zx0\phi(x)=z_{x_{0}}

for x=x1x0x1[x0]x=\cdots x_{-1}x_{0}x_{1}\cdots\in[x_{0}], in which {zi}1in\{z_{i}\}_{1\leq i\leq n} are all constants. Let

Q0(t,z1,z2,,zn)=i=1netziQ_{0}(t,z_{1},z_{2},\cdots,z_{n})=\sum_{i=1}^{n}e^{tz_{i}},

so

P(tϕ)=logQ0(t,z1,,zn)P(t\phi)=\log Q_{0}(t,z_{1},\cdots,z_{n})

by Lemma 5.3. Let

Q1(t,z1,z2,,zn)=i=1nzietziQ_{1}(t,z_{1},z_{2},\cdots,z_{n})=\sum_{i=1}^{n}z_{i}e^{tz_{i}}

and

Q2(t,z1,z2,,zn)=1i<jn(zizj)2et(zi+zj)Q_{2}(t,z_{1},z_{2},\cdots,z_{n})=\sum_{1\leq i<j\leq n}(z_{i}-z_{j})^{2}e^{t(z_{i}+z_{j})}.

Through some elementary calculations one can check that

P(tϕ)=dP(tϕ)dt=Q1(t,z1,,zn)Q0(t,z1,,zn)P^{\prime}(t\phi)=\cfrac{dP(t\phi)}{dt}=\cfrac{Q_{1}(t,z_{1},\cdots,z_{n})}{Q_{0}(t,z_{1},\cdots,z_{n})}

while

(5.3) P′′(tϕ)=d2P(tϕ)dt2=Q2(t,z1,,zn)Q02(t,z1,,zn).P^{\prime\prime}(t\phi)=\cfrac{d^{2}P(t\phi)}{dt^{2}}=\cfrac{Q_{2}(t,z_{1},\cdots,z_{n})}{Q_{0}^{2}(t,z_{1},\cdots,z_{n})}.

Let

R2(t,z1,z2,,zn)=i=1nzi2etziR_{2}(t,z_{1},z_{2},\cdots,z_{n})=\sum_{i=1}^{n}z_{i}^{2}e^{tz_{i}},

one can check that

Q2(t,z1,,zn)=Q0(t,z1,,zn)R2(t,z1,,zn)Q12(t,z1,,zn)Q_{2}(t,z_{1},\cdots,z_{n})=Q_{0}(t,z_{1},\cdots,z_{n})R_{2}(t,z_{1},\cdots,z_{n})-Q_{1}^{2}(t,z_{1},\cdots,z_{n}).

In the following we will often fix t=t>0t=t_{*}>0, so we will frequently write

Q0(t,z1,z2,,zn)=Q0(z1,z2,,zn)Q_{0}(t_{*},z_{1},z_{2},\cdots,z_{n})=Q_{0}(z_{1},z_{2},\cdots,z_{n})

with tt_{*} omitted for convenience. Similar notations apply to other terms above. Let

(5.4) Q0(z1,,zn)=i=1netzi=ea0,Q_{0}(z_{1},\cdots,z_{n})=\sum_{i=1}^{n}e^{t_{*}z_{i}}=e^{a_{0}},
(5.5) Q1(z1,,zn)=i=1nzietzi=a1ea0Q_{1}(z_{1},\cdots,z_{n})=\sum_{i=1}^{n}z_{i}e^{t_{*}z_{i}}=a_{1}e^{a_{0}}

be two equations with unknowns {z1,z2,,zn}\{z_{1},z_{2},\cdots,z_{n}\} for fixed t>0,(a0,a1)2t_{*}>0,(a_{0},a_{1})\in\mathbb{R}^{2} and some n2n\geq 2. Let

Γ5.4n={(z1,z2,,zn)n:z1,z2,,zn satisfy (5.4)}\Gamma_{\ref{eq35}}^{n}=\{(z_{1},z_{2},\cdots,z_{n})\in\mathbb{R}^{n}:z_{1},z_{2},\cdots,z_{n}\mbox{ satisfy }(\ref{eq35})\}

and

Γ5.5n={(z1,z2,,zn)n:z1,z2,,zn satisfy (5.5)}\Gamma_{\ref{eq36}}^{n}=\{(z_{1},z_{2},\cdots,z_{n})\in\mathbb{R}^{n}:z_{1},z_{2},\cdots,z_{n}\mbox{ satisfy }(\ref{eq36})\}.

They are both n1n-1 dimensional smooth hypersurfaces. We first present readers with the following result on fitting an analytic function

a0+a1(tt)+O((tt)2)a_{0}+a_{1}(t-t_{*})+O((t-t_{*})^{2})

with t,a0,a1t_{*},a_{0},a_{1} subject to (1.5) around some fixed t>0t_{*}>0 by pressures of locally constant potentials on general shift spaces of finite type.

5.6 Theorem.

Let t>0,(a0,a1)2,n2t_{*}>0,(a_{0},a_{1})\in\mathbb{R}^{2},n\geq 2 satisfying (1.5) and

(5.6) a0lognt<a1.\cfrac{a_{0}-\log n}{t_{*}}<a_{1}.

Then there exists some δn>0\delta_{n}>0 and some sequence {ri,n}i=1n\{r_{i,n}\}_{i=1}^{n}\subset\mathbb{R}, such that the locally constant potential

ϕ(x)=rx0\phi(x)=r_{x_{0}}

for x=x1x0x1[x0]x=\cdots x_{-1}x_{0}x_{1}\cdots\in[x_{0}] on the full shift space Λn\Lambda_{n}^{\mathbb{Z}} satisfies

P(tϕ)=a0+a1(tt)+O((tt)2)P(t\phi)=a_{0}+a_{1}(t-t_{*})+O((t-t_{*})^{2})

on [tδn,t+δn][t_{*}-\delta_{n},t_{*}+\delta_{n}].

Proof.

In fact it suffices for us to show that the system of equations

{(5.4),(5.5)\left\{\begin{array}[]{ll}(\ref{eq35}),\\ (\ref{eq36})\end{array}\right.

with unknowns {z1,z2,,zn}\{z_{1},z_{2},\cdots,z_{n}\} admits a solution under conditions of the theorem. Without loss of generality we assume

(5.7) z1z2zn.z_{1}\leq z_{2}\leq\cdots\leq z_{n}.

Under this assumption, it is easy to see that

a0logntzn<a0t\cfrac{a_{0}-\log n}{t_{*}}\leq z_{n}<\cfrac{a_{0}}{t_{*}}.

Now we estimate the values of Q1(z1,,zn)Q_{1}(z_{1},\cdots,z_{n}) with znz_{n} approaching the terminals. When znz_{n} approaches the right terminal from below, we have

lim(z1,z2,,zn)Γ5.4n,zna0tQ1(z1,,zn)=a0tea0>a1ea0\lim_{(z_{1},z_{2},\cdots,z_{n})\in\Gamma_{\ref{eq35}}^{n},\ z_{n}\nearrow\frac{a_{0}}{t_{*}}}Q_{1}(z_{1},\cdots,z_{n})=\cfrac{a_{0}}{t_{*}}e^{a_{0}}>a_{1}e^{a_{0}}

in virtue of (1.5). When znz_{n} approaches the left terminal from above, we have

lim(z1,z2,,zn)Γ5.4n,zna0tQ1(z1,,zn)=a0logntea0<a1ea0\lim_{(z_{1},z_{2},\cdots,z_{n})\in\Gamma_{\ref{eq35}}^{n},\ z_{n}\searrow\frac{a_{0}}{t_{*}}}Q_{1}(z_{1},\cdots,z_{n})=\cfrac{a_{0}-\log n}{t_{*}}e^{a_{0}}<a_{1}e^{a_{0}}

in virtue of (5.6). Since Γ5.4n\Gamma_{\ref{eq35}}^{n} is a smooth hypersurface, by the mean value theorem, there exists some (r1,n,r2,n,,rn,n)Γ5.4n(r_{1,n},r_{2,n},\cdots,r_{n,n})\in\Gamma_{\ref{eq35}}^{n} satisfying (5.4) and (5.5) simultaneously. At last, for x=x1x0x1[x0]x=\cdots x_{-1}x_{0}x_{1}\cdots\in[x_{0}] on the full shift space Λn\Lambda_{n}^{\mathbb{Z}}, let

ϕ(x)=rx0,n\phi(x)=r_{x_{0},n}

be the locally constant potential. As P(tϕ)P(t\phi) is analytic, there exists some δn>0\delta_{n}>0 such that

P(tϕ)=a0+a1(tt)+O((tt)2)P(t\phi)=a_{0}+a_{1}(t-t_{*})+O((t-t_{*})^{2})

for t[tδn,t+δn]t\in[t_{*}-\delta_{n},t_{*}+\delta_{n}]. ∎

5.7 Remark.

The core step in the proof of Theorem 5.6 is in fact finding the extremes of the function Q1(z1,,zn)Q_{1}(z_{1},\cdots,z_{n}) subject to (5.4), (1.5) and (5.6). One can detect the points of extremes by the Karush-Kuhn-Tucker (KKT) conditions [Kar, KT], which generalizes the method of Lagrange multipliers by allowing inequality subjections.

Be careful that those {ri,n}i=1n\{r_{i,n}\}_{i=1}^{n} all depend on nn in fact. Theorem 5.6 induces the following interesting flexibility result on fitting certain analytic functions locally by pressures of locally constant potentials on general shift space of finite type.

5.8 Corollary.

Let t>0t_{*}>0 and (a0,a1)2(a_{0},a_{1})\in\mathbb{R}^{2} satisfy (1.5). Then there exists some NN\in\mathbb{N}, such that for any nNn\geq N, there exist some some δn>0\delta_{n}>0 and some sequence {ri,n}i=1n\{r_{i,n}\}_{i=1}^{n}\subset\mathbb{R}, such that the locally constant potential

ϕ(x)=rx0,n\phi(x)=r_{x_{0},n}

for x=x1x0x1[x0]x=\cdots x_{-1}x_{0}x_{1}\cdots\in[x_{0}] on the full shift space Λn\Lambda_{n}^{\mathbb{Z}} satisfies

P(tϕ)=a0+a1(tt)+O((tt)2)P(t\phi)=a_{0}+a_{1}(t-t_{*})+O((t-t_{*})^{2})

on [tδn,t+δn][t_{*}-\delta_{n},t_{*}+\delta_{n}].

Proof.

Under conditions of the corollary, for the given values t,a0,a1t_{*},a_{0},a_{1} satisfying (1.5), choose NN\in\mathbb{N} large enough such that

a0logNt<a1\cfrac{a_{0}-\log N}{t_{*}}<a_{1}.

This means that for any n>Nn>N condition (5.6) is satisfied for t,a0,a1,nt_{*},a_{0},a_{1},n. Then the conclusion follows from Theorem 5.6. ∎

Note that on some particular symbolic spaces Theorem 5.6 and 5.8 may be trivial. For example, for given (t,a0,a1)3(t_{*},a_{0},a_{1})\in\mathbb{R}^{3} without any subjections, by choosing β=ea0ta1\beta=e^{a_{0}-t_{*}a_{1}}, consider the constant potential

ϕ(x)=a1\phi(x)=a_{1}

on the β\beta-shift space with symbols {0,1,,β}\{0,1,\cdots,\lfloor\beta\rfloor\}. It is easy to see that

P(tϕ)=a0ta1+a1t=a0+a1(tt)P(t\phi)=a_{0}-t_{*}a_{1}+a_{1}t=a_{0}+a_{1}(t-t_{*})

on (,)(-\infty,\infty). However, our results guarantee conclusions on general shift spaces.

From now on we go towards the proof of Theorem 1.8. For fixed t>0,(a0,a1)2t_{*}>0,(a_{0},a_{1})\in\mathbb{R}^{2} and n3n\geq 3, let

Γ5.4,5.5n=Γ5.4nΓ5.5n={(z1,z2,,zn)n:z1,z2,,zn satisfy (5.4) and (5.5)}\Gamma_{\ref{eq35},\ref{eq36}}^{n}=\Gamma_{\ref{eq35}}^{n}\cap\Gamma_{\ref{eq36}}^{n}=\{(z_{1},z_{2},\cdots,z_{n})\in\mathbb{R}^{n}:z_{1},z_{2},\cdots,z_{n}\mbox{ satisfy }(\ref{eq35})\mbox{ and }(\ref{eq36})\}.

We describe some topological properties of the set Γ5.4,5.5n\Gamma_{\ref{eq35},\ref{eq36}}^{n} in the following result.

5.9 Lemma.

For fixed t>0,(a0,a1)2t_{*}>0,(a_{0},a_{1})\in\mathbb{R}^{2} subject to (1.5) and n3n\geq 3, in case Γ5.4,5.5n\Gamma_{\ref{eq35},\ref{eq36}}^{n}\neq\emptyset and a1a0lognta_{1}\neq\cfrac{a_{0}-\log n}{t_{*}}, it is a compact (n2)(n-2)-dimension smooth manifold.

Proof.

The Jacobian of the functions Q0(z1,,zn)ea0Q_{0}(z_{1},\cdots,z_{n})-e^{a_{0}} and Q1(z1,,zn)a1ea0Q_{1}(z_{1},\cdots,z_{n})-a_{1}e^{a_{0}} with respect to z1,z2,,znz_{1},z_{2},\cdots,z_{n} is

J=(tetz1tetz2tetznetz1+tz1etz1etz2+tz2etz2etzn+tznetzn).J=\begin{pmatrix}t_{*}e^{t_{*}z_{1}}&t_{*}e^{t_{*}z_{2}}&\cdots&t_{*}e^{t_{*}z_{n}}\\ e^{t_{*}z_{1}}+t_{*}z_{1}e^{t_{*}z_{1}}&e^{t_{*}z_{2}}+t_{*}z_{2}e^{t_{*}z_{2}}&\cdots&e^{t_{*}z_{n}}+t_{*}z_{n}e^{t_{*}z_{n}}\end{pmatrix}.

Its rank is strictly less than 22 if and only if

z1=z2==znz_{1}=z_{2}=\cdots=z_{n}.

Since a1a0lognta_{1}\neq\cfrac{a_{0}-\log n}{t_{*}}, this is excluded from points in Γ5.4,5.5\Gamma_{\ref{eq35},\ref{eq36}}. By the implicit function theorem [Lan], if Γ5.4,5.5n\Gamma_{\ref{eq35},\ref{eq36}}^{n} is not empty, it is an (n2)(n-2)-dimension smooth manifold locally. The gradient of the function Q0(z1,,zn)ea0Q_{0}(z_{1},\cdots,z_{n})-e^{a_{0}} is

(Q0(z1,,zn)ea0)=(tetz1,tetz2,,tetzn)\bigtriangledown(Q_{0}(z_{1},\cdots,z_{n})-e^{a_{0}})=(t_{*}e^{t_{*}z_{1}},t_{*}e^{t_{*}z_{2}},\cdots,t_{*}e^{t_{*}z_{n}}),

whose individual components will always be strictly positive. The gradient of the function Q1(z1,,zn)a1ea0Q_{1}(z_{1},\cdots,z_{n})-a_{1}e^{a_{0}} is

(Q1(z1,,zn)a1ea0)=(etz1+tz1etz1,etz2+tz2etz2,,etzn+tznetzn)\bigtriangledown(Q_{1}(z_{1},\cdots,z_{n})-a_{1}e^{a_{0}})=(e^{t_{*}z_{1}}+t_{*}z_{1}e^{t_{*}z_{1}},e^{t_{*}z_{2}}+t_{*}z_{2}e^{t_{*}z_{2}},\cdots,e^{t_{*}z_{n}}+t_{*}z_{n}e^{t_{*}z_{n}}),

with the ii-th individual component vanishes if and only if zi=1tz_{i}=-\cfrac{1}{t_{*}} for 1in1\leq i\leq n. So Γ5.4n\Gamma_{\ref{eq35}}^{n} and Γ5.5n\Gamma_{\ref{eq36}}^{n} cannot be tangent to each other. Moreover, note that

etzi+tzietzi>0e^{t_{*}z_{i}}+t_{*}z_{i}e^{t_{*}z_{i}}>0

if zi>1tz_{i}>-\cfrac{1}{t_{*}} while

etzi+tzietzi<0e^{t_{*}z_{i}}+t_{*}z_{i}e^{t_{*}z_{i}}<0

if zi<1tz_{i}<-\cfrac{1}{t_{*}} for any 1in1\leq i\leq n. These force the intersection of zeros of the two functions Q0(z1,,zn)ea0Q_{0}(z_{1},\cdots,z_{n})-e^{a_{0}} and Q1(z1,,zn)a1ea0Q_{1}(z_{1},\cdots,z_{n})-a_{1}e^{a_{0}} to be connected, if the intersection is not empty. This implies Γ5.4,5.5\Gamma_{\ref{eq35},\ref{eq36}} is a manifold globally in case of being nonempty. Γ5.4,5.5n\Gamma_{\ref{eq35},\ref{eq36}}^{n} is compact since it is a bounded set. ∎

Let

Γ5.4,1,2,13={(z1,z2,z3)3:z1,z2,z3 satisfy ez1+ez2+ez3=e2}\Gamma_{\ref{eq35},1,2,1}^{3}=\{(z_{1},z_{2},z_{3})\in\mathbb{R}^{3}:z_{1},z_{2},z_{3}\mbox{ satisfy }e^{z_{1}}+e^{z_{2}}+e^{z_{3}}=e^{2}\}

and

Γ5.5,1,2,13={(z1,z2,z3)3:z1,z2,z3 satisfy z1ez1+z2ez2+z3ez3=e2}\Gamma_{\ref{eq36},1,2,1}^{3}=\{(z_{1},z_{2},z_{3})\in\mathbb{R}^{3}:z_{1},z_{2},z_{3}\mbox{ satisfy }z_{1}e^{z_{1}}+z_{2}e^{z_{2}}+z_{3}e^{z_{3}}=e^{2}\}

be the corresponding surfaces with t=1,a0=2,a1=1t_{*}=1,a_{0}=2,a_{1}=1. Figure 2 depicts parts of the two 22-dimension surfaces, whose intersection will be a 11-dimension smooth curve.

Refer to caption
Figure 2. Γ5.4,1,2,13\Gamma_{\ref{eq35},1,2,1}^{3} (green) and Γ5.5,1,2,13\Gamma_{\ref{eq36},1,2,1}^{3} (red)

Equipped with all the above results, now we are ready to prove Theorem 1.8.

Proof of Theorem 1.8.

First, for the given t>0t_{*}>0 and (a0,a1)2(a_{0},a_{1})\in\mathbb{R}^{2} satisfying (1.5), if nn is large enough, Γ5.4,5.5n\Gamma_{\ref{eq35},\ref{eq36}}^{n} is not empty according to Corollary 5.8. So Γ5.4,5.5n\Gamma_{\ref{eq35},\ref{eq36}}^{n} is a compact (n2)(n-2)-dimension smooth manifold for nn large enough. In the following we always assume nn is large enough. Now let

mt,a0,a1,n=min{R2(t,z1,z2,,zn)ea0a12:(z1,z2,,zn)Γ5.4,5.5n}m_{t_{*},a_{0},a_{1},n}=\min\Big{\{}\cfrac{R_{2}(t_{*},z_{1},z_{2},\cdots,z_{n})}{e^{a_{0}}}-a_{1}^{2}:(z_{1},z_{2},\cdots,z_{n})\in\Gamma_{\ref{eq35},\ref{eq36}}^{n}\Big{\}}

while

(5.8) Mt,a0,a1,n=max{R2(t,z1,z2,,zn)ea0a12:(z1,z2,,zn)Γ5.4,5.5n}.M_{t_{*},a_{0},a_{1},n}=\max\Big{\{}\cfrac{R_{2}(t_{*},z_{1},z_{2},\cdots,z_{n})}{e^{a_{0}}}-a_{1}^{2}:(z_{1},z_{2},\cdots,z_{n})\in\Gamma_{\ref{eq35},\ref{eq36}}^{n}\Big{\}}.

For any mt,a0,a1,na2Mt,a0,a1,nm_{t_{*},a_{0},a_{1},n}\leq a_{2}\leq M_{t_{*},a_{0},a_{1},n}, since Γ5.4,5.5n\Gamma_{\ref{eq35},\ref{eq36}}^{n} is a smooth manifold, there exist {ci,n}i=1n\{c_{i,n}\}_{i=1}^{n}\subset\mathbb{R}, such that (c1,n,c2,n,,cn,n)(c_{1,n},c_{2,n},\cdots,c_{n,n}) satisfies (5.4),(5.5)(\ref{eq35}),(\ref{eq36}) and

(5.9) a2=Q2(t,c1,n,,cn,n)Q02(t,c1,n,,cn,n)=R2(t,c1,n,,cn,n)ea0a12a_{2}=\cfrac{Q_{2}(t_{*},c_{1,n},\cdots,c_{n,n})}{Q_{0}^{2}(t_{*},c_{1,n},\cdots,c_{n,n})}=\cfrac{R_{2}(t_{*},c_{1,n},\cdots,c_{n,n})}{e^{a_{0}}}-a_{1}^{2}

simultaneously. Now let

ϕ(x)=cx0,n\phi(x)=c_{x_{0},n}

for x=x1x0x1[x0]x=\cdots x_{-1}x_{0}x_{1}\cdots\in[x_{0}] on the full shift space Λn\Lambda_{n}^{\mathbb{Z}}. It is a locally constant potential. According to (5.3) and (5.9), we have

(5.10) P′′(tϕ)=Q2(t,c1,n,,cn,n)Q02(t,c1,n,,cn,n)=a2.P^{\prime\prime}(t_{*}\phi)=\cfrac{Q_{2}(t_{*},c_{1,n},\cdots,c_{n,n})}{Q_{0}^{2}(t_{*},c_{1,n},\cdots,c_{n,n})}=a_{2}.

Since (c1,n,c2,n,,cn,n)(c_{1,n},c_{2,n},\cdots,c_{n,n}) satisfies (5.4)(\ref{eq35}) and (5.5)(\ref{eq36}), we have

(5.11) P(tϕ)=Q2(t,c1,n,,cn,n)Q02(t,c1,n,,cn,n)=a0P(t_{*}\phi)=\cfrac{Q_{2}(t_{*},c_{1,n},\cdots,c_{n,n})}{Q_{0}^{2}(t_{*},c_{1,n},\cdots,c_{n,n})}=a_{0}

while

(5.12) P(tϕ)=Q2(t,c1,n,,cn,n)Q02(t,c1,n,,cn,n)=a1.P^{\prime}(t_{*}\phi)=\cfrac{Q_{2}(t_{*},c_{1,n},\cdots,c_{n,n})}{Q_{0}^{2}(t_{*},c_{1,n},\cdots,c_{n,n})}=a_{1}.

Note that P(tϕ)P(t\phi) is analytic with respect to tt on (α,)(\alpha,\infty) for any α>0\alpha>0, so there exists some δn>0\delta_{n}>0, such that (1.6) holds on [tδn,t+δn][t_{*}-\delta_{n},t_{*}+\delta_{n}], considering (5.10), (5.11) and (5.12). ∎

In the following we illustrate some dependent relationship between

{mt,a0,a1,n,Mt,a0,a1,n}n\{m_{t_{*},a_{0},a_{1},n},M_{t_{*},a_{0},a_{1},n}\}_{n\in\mathbb{N}}

and some particular t,a0,a1,nt_{*},a_{0},a_{1},n satisfying (1.5). There should be some universal relationship between them, while we hope the following observations will provide some hints. The first one is that it is possible for mt,a0,a1,n=0m_{t_{*},a_{0},a_{1},n}=0 for some t,a0,a1,nt_{*},a_{0},a_{1},n.

5.10 Proposition.

Let t>0t_{*}>0 and (a0,a1)2(a_{0},a_{1})\in\mathbb{R}^{2} satisfy (1.5). Then mt,a0,a1,n=0m_{t_{*},a_{0},a_{1},n}=0 for n2n\geq 2 if and only if

(5.13) a1=a0lognt.a_{1}=\cfrac{a_{0}-\log n}{t_{*}}.
Proof.

Note that mt,a0,a1,n=0m_{t_{*},a_{0},a_{1},n}=0 is equivalent to say that there exists some locally constant potential ϕ\phi on Λn\Lambda_{n}^{\mathbb{Z}} such that P′′(tϕ)=0P^{\prime\prime}(t_{*}\phi)=0 according to Theorem 1.8. By [PP, Proposition 4.12], this happens if and only if ϕ\phi is a constant potential on Λn\Lambda_{n}^{\mathbb{Z}}. In this case we have

ϕ(x)=a0lognt\phi(x)=\cfrac{a_{0}-\log n}{t_{*}}

for any xΛnx\in\Lambda_{n}^{\mathbb{Z}}, which implies (5.13). ∎

This result does not tell things about the sequence

{mt,a0,a1,n}n large enough \{m_{t_{*},a_{0},a_{1},n}\}_{\ n\in\mathbb{N}\mbox{ large enough }}

for given t,a0,a1t_{*},a_{0},a_{1}, since (5.13) will never be true for any nn large enough for fixed t,a0,a1t_{*},a_{0},a_{1}. The following result describes some limit behaviour of the sequence

{Mt,a0,a1,n}n large enough \{M_{t_{*},a_{0},a_{1},n}\}_{\ n\in\mathbb{N}\mbox{ large enough }}

for t=1,a0=2,a1=1t_{*}=1,a_{0}=2,a_{1}=1.

5.11 Proposition.

Let t=1,a0=2,a1=1t_{*}=1,a_{0}=2,a_{1}=1, in symbols of Theorem 1.8, we have

(5.14) limnM1,2,1,n=.\lim_{n\rightarrow\infty}M_{1,2,1,n}=\infty.

To justify Proposition 5.11, we first illustrate some basic properties about the function zetzze^{t_{*}z} for t>0t_{*}>0.

5.12 Lemma.

For t>0t_{*}>0, zetzze^{t_{*}z} is strictly decreasing on (,1t)(-\infty,-\cfrac{1}{t_{*}}), strictly increasing on (1t,)(-\cfrac{1}{t_{*}},\infty), while it attains its minimum 1te1-\cfrac{1}{t_{*}}e^{-1} at z=1tz=-\cfrac{1}{t_{*}}. It admits one and only one inflection in (,1t)(-\infty,-\cfrac{1}{t_{*}}).

Proof.

One can check these conclusions by some direct computations on the first and second derivatives of the function zetzze^{t_{*}z}. ∎

In Figure 3 we depict the graph of ς(z)=zez\varsigma(z)=ze^{z}.

Refer to caption
Figure 3. Graph of ς(z)=zez\varsigma(z)=ze^{z}
Refer to caption
Figure 4. Γ5.15\Gamma_{\ref{eq46}} and Γ5.16\Gamma_{\ref{eq47}}
Proof of Proposition 5.11.

Since we are considering the limit behaviour of M1,2,1,nM_{1,2,1,n}, we always assume nn is large enough throughout the proof. Now consider the following two equations

(5.15) (n1)eza+ezb=e2(n-1)e^{z_{a}}+e^{z_{b}}=e^{2}

and

(5.16) (n1)zaeza+zbezb=e2(n-1)z_{a}e^{z_{a}}+z_{b}e^{z_{b}}=e^{2}

with unknowns za,zbz_{a},z_{b}. Let

Γ5.15={(za,zb)2:za,zb satisfy (5.15)}\Gamma_{\ref{eq46}}=\{(z_{a},z_{b})\in\mathbb{R}^{2}:z_{a},z_{b}\mbox{ satisfy }(\ref{eq46})\}

and

Γ5.16={(za,zb)2:za,zb satisfy (5.16)}\Gamma_{\ref{eq47}}=\{(z_{a},z_{b})\in\mathbb{R}^{2}:z_{a},z_{b}\mbox{ satisfy }(\ref{eq47})\}.

We describe the graph of Γ5.15\Gamma_{\ref{eq46}} and Γ5.16\Gamma_{\ref{eq47}} separately in the following. Γ5.15\Gamma_{\ref{eq46}} is a 11 dimensional smooth curve with two asymptotes za=2log(n1)z_{a}=2-\log(n-1) and zb=2z_{b}=2. It is strictly decreasing when we consider the curve as the graph of the function

zb=log(e2(n1)eza)z_{b}=\log\big{(}e^{2}-(n-1)e^{z_{a}}\big{)}

for za(,2log(n1))z_{a}\in(-\infty,2-\log(n-1)). Γ5.16\Gamma_{\ref{eq47}} is also a 11 dimensional smooth curve with two asymptotes za=ς1(e2n1)z_{a}=\varsigma^{-1}(\cfrac{e^{2}}{n-1}) and zb=ς1(e2)z_{b}=\varsigma^{-1}(e^{2}). When we consider the Γ5.16\Gamma_{\ref{eq47}} as the graph of the function

zb=η(za)z_{b}=\eta(z_{a})

as the implicit function induced by (5.16)(\ref{eq47}), it is strictly increasing for za(,1)z_{a}\in(-\infty,-1), strictly decreasing for za(1,ς1(e2n1))z_{a}\in(-1,\varsigma^{-1}(\cfrac{e^{2}}{n-1})), with its maximum ς1(e2+(n1)e1)\varsigma^{-1}(e^{2}+(n-1)e^{-1}) attained at za=1z_{a}=-1. Let ςl1(e2n1)\varsigma_{l}^{-1}(-\cfrac{e^{2}}{n-1}) be the smaller one of the two intersections of zb=2z_{b}=2 and Γ5.16\Gamma_{\ref{eq47}}, then Γ5.15\Gamma_{\ref{eq46}} and Γ5.16\Gamma_{\ref{eq47}} must intersection at some unique point ca,n(,ςl1(e2n1))c_{a,n}\in(-\infty,\varsigma_{l}^{-1}(-\cfrac{e^{2}}{n-1})). Obviously

limnca,n=\lim_{n\rightarrow\infty}c_{a,n}=-\infty

since limnςl1(e2n1)=\lim_{n\rightarrow\infty}\varsigma_{l}^{-1}(-\cfrac{e^{2}}{n-1})=-\infty. Now we analyse the order of ca,nc_{a,n} with respect to nn as nn\rightarrow\infty. Let

za,n=lognloglogn+log11z_{a,n}=-\log n-\log\log n+\log 1-1.

One can check that

limn,zb2((n1)eza,n+ezb)=e2\lim_{n\rightarrow\infty,z_{b}\rightarrow 2}((n-1)e^{z_{a,n}}+e^{z_{b}})=e^{2}

while

limn,zb2((n1)za,neza,n+zbezb)=e2\lim_{n\rightarrow\infty,z_{b}\rightarrow 2}((n-1)z_{a,n}e^{z_{a,n}}+z_{b}e^{z_{b}})=e^{2}.

These imply that

ca,n=lognloglogn+o(loglogn)c_{a,n}=-\log n-\log\log n+o(\log\log n).

Note that (ca,n,ca,n,,ca,n,η(ca,n))Γ5.4,5.5n(c_{a,n},c_{a,n},\cdots,c_{a,n},\eta(c_{a,n}))\in\Gamma_{\ref{eq35},\ref{eq36}}^{n} for t=1,a0=2,a1=1t_{*}=1,a_{0}=2,a_{1}=1. Now

R2(ca,n,ca,n,,ca,n,η(ca,n))=(n1)ca,n2eca,n+(η(ca,n))2eη(ca,n)=(n1)(lognloglogn+o(loglogn))2elognloglogn+o(loglogn)+4e2+o(1)=logn+o(logn),\begin{array}[]{ll}&R_{2}(c_{a,n},c_{a,n},\cdots,c_{a,n},\eta(c_{a,n}))\vspace{3mm}\\ =&(n-1)c_{a,n}^{2}e^{c_{a,n}}+(\eta(c_{a,n}))^{2}e^{\eta(c_{a,n})}\vspace{3mm}\\ =&(n-1)(-\log n-\log\log n+o(\log\log n))^{2}e^{-\log n-\log\log n+o(\log\log n)}+4e^{2}+o(1)\vspace{3mm}\\ =&\log n+o(\log n),\end{array}

from which it is easy to see that

limnR2(ca,n,ca,n,,ca,n,η(ca,n))=\lim_{n\rightarrow\infty}R_{2}(c_{a,n},c_{a,n},\cdots,c_{a,n},\eta(c_{a,n}))=\infty.

This forces

limnM1,2,1,n=\lim_{n\rightarrow\infty}M_{1,2,1,n}=\infty,

considering (5.8).

We provide the readers with the curves Γ5.15\Gamma_{\ref{eq46}} and Γ5.16\Gamma_{\ref{eq47}} in Figure 4. Obviously some more general conclusions are available if one considers variations of the parameters t,a0,a1t_{*},a_{0},a_{1} in Proposition 5.11. At last we provide the readers with some solutions {ca,n}n\{c_{a,n}\}_{n\in\mathbb{N}} and {η(ca,n)}n\{\eta(c_{a,n})\}_{n\in\mathbb{N}} in Table 3, from which one can see the order of decay and increase of the sequences with respect to nn clearly.

Table 3. {ca,n}n\{c_{a,n}\}_{n\in\mathbb{N}} and {η(ca,n)}n\{\eta(c_{a,n})\}_{n\in\mathbb{N}}
nn ca,nc_{a,n} η(ca,n)\eta(c_{a,n})
1010 -1.8599539391797653780996686364493 1.7634042477581860636342812520981
10210^{2} -4.6278529940301947157458180305676 1.8580906928560505140960875180438
10310^{3} -7.2278923365046354303919671475052 1.8965708210067454817129699066334
10410^{4} -9.7529279223041958189401940128674 1.9180710389285259082138396366755
10510^{5} -12.23426184122178540565187685582 1.9319494203818796717151866525306
10610^{6} -14.686689485112383196253350885528 1.941701042038176132682488585943
10710^{7} -17.118475509130338419321449219176 1.9489507180131363431129601417792
10810^{8} -19.534737736752111249670741176574 1.9545628133690736391913141129777
10910^{9} -21.938877884281897893422087428599 1.9590417833080193886068703580662
101010^{10} -24.333277592346602338263750350022 1.9627027620469153955488959845337
101110^{11} -26.719672172461371813735932628894 1.9657531814729595378854181456218
101210^{12} -29.099366670257435261982274861811 1.9683353707111573738492465130807
101310^{13} -31.473368167571030624456199153849 1.970550350496947761285545176838
101410^{14} -33.842470627269595326611535858951 1.9724718685216929582206115029034
101510^{15} -36.20731141238751139407393422892 1.9741550583546827046855344007126
101610^{16} -38.568410155198951836337896822881 1.97564198636943790477268372057
101710^{17} -40.926196222869058989174011616314 1.9769653208730088904749619599928
101810^{18} -43.28102858421294787781225809291 1.9781508271703613365389080750692
101910^{19} -45.633210475623427729647938869856 1.9792191056459012534062976747755
102010^{20} -47.983000423353389741328990557576 1.9801868284846851379610473178804
102110^{21} -50.330620660008332271820694306839 1.9810676363715292020369862557429
102210^{22} -52.676263643082855194671803053742 1.9818727996772032079642260800619
102310^{23} -55.020097168291592849066888176454 1.9826117134133018944596936081392
102410^{24} -57.362268427077060922578379063246 1.9832922728467949817312209115653
102510^{25} -59.702907260160132201351723856461 1.9839211621102961084222105523408
102610^{26} -62.042128791447074538616865826092 1.9845040784885601043186175801529
102710^{27} -64.380035579030470553978577616248 1.9850459085371711281404342988732
102810^{28} -66.716719386002755126963619613768 1.9855508677057357884921072471682
102910^{29} -69.052262649137714881922574449762 1.9860226120088820356292880321385
103010^{30} -71.386739705385277326962820249044 1.9864643280735340181774784668139
103110^{31} -73.7202178226698188293210417966 1.9868788063025456510398996161088
103210^{32} -76.052758071376257724956806229201 1.9872685007417625212636588061233
103310^{33} -78.384416065240707497606345034329 1.9876355783911370649894789906831
103410^{34} -80.715242594490126828808238297291 1.9879819600725889180558042388717
103510^{35} -83.045284169538297201269228695051 1.9883093544968926933418986392956
103610^{36} -85.374583490011093204926910773042 1.9886192868162322855194994642595
103710^{37} -87.703179851099500408441885242821 1.9889131226778662869710690898493
103810^{38} -90.031109497045012553979249690171 1.9891920885858755212588911444123
103910^{39} -92.358405929815521230622254914238 1.9894572892164831961499157326311
104010^{40} -94.685100179630439886817678169988 1.9897097222064583485549741614556

References

  • [BDL] V. Baladi, M. Demers and C. Liverani, Exponential Decay of Correlations for Finite Horizon Sinai Billiard Flows, Inventiones mathematicae, volume 211, pp. 39-177 (2018).
  • [BKR] J. Bochi, A. Katok and F. Rodriguez Hertz, Flexibility of Lyapunov exponents, Ergod. Theory Dyn. Syst., 42(2): Anatole Katok Memorial Issue Part 1: Special Issue of Ergodic Theory and Dynamical Systems, February 2022, pp. 554-591.
  • [Bow] R. Bowen, Equilibrium States and the Ergodic Theory of Anosov Diffeomorphisms, Lecture Notes in Mathematics 470, Springer, 1975.
  • [CP] Z. Coelho and W. Parry, Central Limit asymptotics for shifts of finite type, Israel J. Math., Vol 69, N.2, (1990).
  • [Fel] W. Feller, An Introduction to Probability Theory and its Applications, Vol. 2, 2nd ed., John Wiley & Sons, New York, 1971.
  • [GLP] P. Giulietti, C. Liverani and M. Pollicott, Anosov Flows and Dynamical Zeta Functions, Annals of Mathematics, 178, 2 (2013) 687-773.
  • [GT] D. Gilbarg and N. Trudinger, Elliptic Partial Differential Equations of Second Order, New York: Springer, 1983.
  • [IRV] G. Iommi, F. Riquelme and A. Velozo, Entropy in the cusp and phase transitions for geodesic flows, Isr. J. Math., 225 (2018) 609-659.
  • [IT1] G. Iommi and M. Todd, Transience in dynamical systems, Ergod. Theory Dyn. Syst., 33(5) (2013) 1450-1476.
  • [IT2] G. Iommi and M. Todd, Differentiability of the pressure in non-compact spaces, arXiv:2010.10250 [math.DS], 2020.
  • [Kar] W. Karush, Minima of Functions of Several Variables with Inequalities as Side Constraints (M.Sc. thesis), Dept. of Mathematics, Univ. of Chicago, 1939.
  • [KS1] M. Kotani and T. Sunada, The pressure and higher correlations for an Anosov diffeomorphism, Ergod. Theory Dyn. Syst., 21 (2001), no. 3, 807-821.
  • [KS2] M. Kotani and T. Sunada, A Central Limit Theorem for the Simple Random Walk on a Crystal Lattice, Proceedings of the Second ISAAC Congress, pp 1-6, 2000.
  • [KQ] T. Kucherenko and A. Quas, Flexibility of the Pressure Function, Comm. Math. Phys., to appear.
  • [KQW] T. Kucherenko, A. Quas and C. Wolf, Multiple phase transitions on compact symbolic systems, Adv. in Math., 385 (2021).
  • [KT] H. Kuhn and A. Tucker, Nonlinear programming, Proceedings of 2nd Berkeley Symposium, Berkeley: University of California Press, pp. 481-492, 1951.
  • [Lal] S. Lalley, Ruelle’s Perron-Frobenius theorem and central limit theorem for additive functionals of one-dimensional Gibbs states, Proc. Conf. in honour of H. Robbins, 1985.
  • [Lan] S. Lang, Fundamentals of Differential Geometry, Graduate Texts in Mathematics 191, New York: Springer, 1999.
  • [Lop1] A. Lopes, The Dimension spectrum and a mathematical model for phase transition, Adv. in Appl. Math., 11 No. 4 (1990), 475-502.
  • [Lop2] A. Lopes, The first order level 2 phase transition in thermodynamic formalism, J. Sta. Phys., 60 Nos 3/4 (1990), 395-411.
  • [Lop3] A. Lopes, The Zeta Function, non-differentiability of the pressure, and the critical exponent of transition, Adv. in Math., 101 (1993), 133-165.
  • [PP] W. Parry and M. Pollicott, Zeta functions and the periodic orbit structure of hyperbolic dynamics, Astérisque, Vol. 187-188, 1990.
  • [Rou] J. Rousseau-Egèle, Un théorèdme de la limite locale pour une classe de transformations dilatantes et monotones par morceaux, Ann. of Prob., 11 (1983), 772-788.
  • [Rue1] D. Ruelle, Thermodynamic formalism: The mathematical structures of equilibrium statistical mechanics, Second edition, Cambridge University Press, Cambridge, 2004.
  • [Rue2] D. Ruelle, Statistical mechanics of a one-dimensional lattice gas, Comm. Math. Phys., 9 (1968), 267-278.
  • [Sar] O. Sarig, On an example with a non-analytic topological pressure, C. R. Acad. Sci. Paris, Sér. I Math., 330 (2000) 311-315.
  • [Tim] A. Timan, Theory of Approximation of Functions of a Real Variable, translated by J. Berry from ’Teoriya priblizheniya funktsii deistvitel’nogo peremennogo’, Pergamon Press LTD., 1963.
  • [Wal1] P. Walters, An introduction to ergodic theory, Graduate Texts in Mathematics 79, Springer, 1981.
  • [Wal2] P. Walters, A necessary and sufficient condition for a two-sided continuous function to be cohomologous to a one-sided continuous function, Dyn. Syst., 18(2) (2003) 131-138.