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

Space spanned by characteristic exponents

Zhuchao Ji Institute for Theoretical Sciences, Westlake University, Hangzhou 310030, China [email protected] Junyi Xie Beijing International Center for Mathematical Research, Peking University, Beijing 100871, China [email protected]  and  Geng-Rui Zhang School of Mathematical Sciences, Peking University, Beijing 100871, China [email protected]
Abstract.

We prove several rigidity results on multiplier and length spectrum. For example, we show that for every non-exceptional rational map f:1()1()f:\mathbb{P}^{1}(\mathbb{C})\to\mathbb{P}^{1}(\mathbb{C}) of degree d2d\geq 2, the \mathbb{Q}-vector space generated by the characteristic exponents (that are not -\infty) of periodic points of ff has infinite dimension. This answers a stronger version of a question of Levy and Tucker. Our result can also be seen as a generalization of recent results of Ji-Xie and of Huguin which proved Milnor’s conjecture about rational maps having integer multipliers. We also get a characterization of postcritically finite maps by using its length spectrum. Finally as an application of our result, we get a new proof of the Zariski dense orbit conjecture for endomorphisms on (1)N,N1(\mathbb{P}^{1})^{N},N\geq 1.

1. Introduction

Let f:11f:\mathbb{P}^{1}\to\mathbb{P}^{1} be a rational map over \mathbb{C} of degree d2d\geq 2. Our aim is to study the \mathbb{Q}-vector space spanned by the characteristic exponents of periodic points of a rational map on 1()\mathbb{P}^{1}(\mathbb{C}) and prove some rigidity results.

1.1. Multiplier, length and characteristic exponent

Let z01()z_{0}\in\mathbb{P}^{1}(\mathbb{C}) be a periodic point of ff with exact period nn. Define nf(z0):=nn_{f}(z_{0}):=n be this period. We write n(z0)n(z_{0}) for simplicity when the map ff is clear. The multiplier ρf(z0)\rho_{f}(z_{0}) of ff at z0z_{0} is defined to be the differential dfn(z0)df^{n}(z_{0})\in\mathbb{C}. We write ρ(z0)\rho(z_{0}) for simplicity when the map ff is clear. The length of ff at z0z_{0} is the norm |ρf(z0)|.|\rho_{f}(z_{0})|. The multiplier and the length are invariant under conjugacy. The characteristic exponent of ff at z0z_{0} is defined to be χf(z0):=n1log|ρf(z0)|\chi_{f}(z_{0}):=n^{-1}{\rm log}\lvert\rho_{f}(z_{0})\rvert.

Denote by Per(f)(){\rm Per}(f)(\mathbb{C}) the set of all periodic points in 1()\mathbb{P}^{1}(\mathbb{C}) of ff and define Per(f)():={z0Per(f)():ρf(z0)0}{\rm Per}^{*}(f)(\mathbb{C}):=\{z_{0}\in{\rm Per}(f)(\mathbb{C}):\rho_{f}(z_{0})\neq 0\}. When the base field \mathbb{C} is clear, we also write Per(f){\rm Per}(f) and Per(f){\rm Per}^{*}(f) for simplicity.

The Lyapunov exponent (of the maximal entropy measure) of ff is defined by

f:=1()log|df|𝑑μf,\mathcal{L}_{f}:=\int_{\mathbb{P}^{1}(\mathbb{C})}{\rm log}|df|d\mu_{f},

where μf\mu_{f} is the unique maximal entropy measure, and the norm of the differential is computed with respect to the spherical metric.

1.2. Exceptional maps

In complex dynamics, the exceptional maps defined below are often considered as exceptional examples among all rational maps. We may view them as rational maps on 1()\mathbb{P}^{1}(\mathbb{C}) related to algebraic groups.

Definition 1.1.

Let f:11f:\mathbb{P}^{1}\to\mathbb{P}^{1} be an endomorphism over \mathbb{C} of degree d2d\geq 2.

  • It is called Lattès if it is semi-conjugate to an endomorphism on an elliptic curve. Further it is called flexible Lattès if it is semi-conjugate to the multiplication by an integer nn on an elliptic curve for some |n|2.|n|\geq 2. Otherwise, it is called rigid Lattès.

  • We say that ff is of monomial type if it semi-conjugate to the map zznz\mapsto z^{n} on 1\mathbb{P}^{1} for some integer nn with |n|2.|n|\geq 2.

  • We call ff exceptional if it is Lattès or of monomial type. An endomorphism ff is exceptional if and only if some iterate fkf^{k} is exceptional (k>0k\in\mathbb{Z}_{>0}).

1.3. Statement of the main results

We fix an embedding of the algebraic closure ¯\overline{\mathbb{Q}} of \mathbb{Q} in \mathbb{C} and identify ¯\overline{\mathbb{Q}} as a subfield of \mathbb{C}, hence any number field is a subfield of \mathbb{C}. Denote the usual absolute value on \mathbb{C} by |||\cdot|.

Our first result shows that the definition field of a non-flexible Lattès rational map is determined by its length spectrum.

Theorem 1.2.

Let f:11f:\mathbb{P}_{\mathbb{C}}^{1}\to\mathbb{P}_{\mathbb{C}}^{1} be a rational map of degree at least 22. Assume that ff is not a flexible Lattès map and for every xPer(f)()x\in{\rm Per}(f)(\mathbb{C}), |ρf(x)|¯|\rho_{f}(x)|\in\overline{\mathbb{Q}}. Then ff is defined over ¯.\overline{\mathbb{Q}}.

In Theorem 2.1, we indeed proved a more general version of Theorem 1.2, in which ¯\overline{\mathbb{Q}} can be replaced to any algebraically closed subfield of \mathbb{C} which is invariant under the complex conjugation.

McMullen’s rigidity of multiplier spectrum [McM87] with a standard spread out argument implies that, for a rational map ff of degree at least 22 which is not flexible Lattès, if its multipliers at periodic points are all algebraic, then ff is defined over ¯\overline{\mathbb{Q}}. Theorem 1.2 is a generalization of this result from multiplier spectrum to length spectrum (which contains less information). The rigidity of length spectrum was proved in [JX23b, Theorem 1.5]. However, the spread out argument does not apply directly in this case as the length spectrum map (and its square) is not algebraic on the moduli space of rational maps. Indeed as shown in [JX23b, Section 8.1], its square is not even real algebraic. In Section 2.3, we introduce a way to do the spread out argument respecting the real structure using Weil restriction. Another difficulty in the length spectrum case is the lack of noetherianity for semi-algebraic subsets. We overcome this difficulty using the notion of admissible subsets introduced in [JX23b].

The following two results concern the \mathbb{Q}-vector space spanned by the characteristic exponents of periodic points.

Theorem 1.3.

Let f:1()1()f:\mathbb{P}^{1}(\mathbb{C})\to\mathbb{P}^{1}(\mathbb{C}) be a rational map of degree d2d\geq 2. Suppose that ff is not exceptional. Then the \mathbb{Q}-vector space generated by {χf(z):zPer(f)}\{\chi_{f}(z):z\in{\rm Per}^{*}(f)\} in \mathbb{R} has infinite dimension.

Theorem 1.4.

Let f:1()1()f:\mathbb{P}^{1}(\mathbb{C})\to\mathbb{P}^{1}(\mathbb{C}) be a rational map of degree d2d\geq 2. Assume that there exists a number field KK such that

(1.1) z0Per(f),n=n(z0)>0,|ρf(z0)|nK.\forall z_{0}\in{\rm Per}(f),\ \exists n=n(z_{0})\in\mathbb{Z}_{>0},\ \lvert\rho_{f}(z_{0})\rvert^{n}\in K.

Then ff is exceptional.

Finitely many nonzero elements z1,,zNz_{1},\dots,z_{N} in a commutative ring RR are called multiplicatively independent if for all triples (m1,,mN)(m_{1},\dots,m_{N}) of integers, z1m1zNmN=1z_{1}^{m_{1}}\cdots z_{N}^{m_{N}}=1 if and only if m1==mN=0m_{1}=\cdots=m_{N}=0. A sequence (zn)n=1(z_{n})_{n=1}^{\infty} in R{0}R\setminus\{0\} is called multiplicatively independent if any its finite subsequence is multiplicatively independent. Theorem 1.3 immediately implies the existence of infinitely many multipliers for a non-exceptional ff whose absolute values are multiplicatively independent.

Corollary 1.5.

Let f:1()1()f:\mathbb{P}^{1}(\mathbb{C})\to\mathbb{P}^{1}(\mathbb{C}) be a rational map of degree d2d\geq 2. Suppose that ff is not exceptional. Then there exists a sequence (xj)j=1(x_{j})_{j=1}^{\infty} in Per(f){\rm Per}^{*}(f) such that the sequence (|ρf(xj)|)j=1(\lvert\rho_{f}(x_{j})\rvert)_{j=1}^{\infty} is multiplicatively independent in \mathbb{R}.

1.4. Motivations and previous results

Milnor’s conjecture

Milnor [Mil06] has showed that an exceptional rational map f:1()1()f:\mathbb{P}^{1}(\mathbb{C})\to\mathbb{P}^{1}(\mathbb{C}) of degree d2d\geq 2 must have all its multipliers of periodic points in the ring of integers 𝒪K\mathcal{O}_{K} for some imaginary quadratic number field KK, and in fact in \mathbb{Z} when ff is not a rigid Lattès map. Milnor conjectured that the converse is also true. Milnor’s conjecture was recently proved by Ji and Xie:

Theorem 1.6 ([JX23b, Theorem 1.13]).

Let f:1()1()f:\mathbb{P}^{1}(\mathbb{C})\to\mathbb{P}^{1}(\mathbb{C}) be a rational map of degree d2d\geq 2. Assume that there exists an imaginary quadratic field KK such that all multipliers of ff belong to 𝒪K\mathcal{O}_{K}. Then ff is exceptional.

See also [BGHR22] for a different proof. Recently Huguin generalized the above result using different approach:

Theorem 1.7 ([Hug23, Theorem 7]).

Let f:1()1()f:\mathbb{P}^{1}(\mathbb{C})\to\mathbb{P}^{1}(\mathbb{C}) be a rational map of degree d2d\geq 2. Assume that there exists a number field KK such that all multipliers of ff belong to KK. Then ff is exceptional.

Since our assumption (1.1) in Theorem 1.4 is weaker than that of Theorem 1.7, Theorem 1.4 is a generalization of Theorem 1.7. Indeed our assumption (1.1) is even weaker than the condition that there is a number field KK such that

(1.2) z0Per(f),n>0,(ρf(z0))nK.\forall z_{0}\in{\rm Per}(f),\ \exists n\in\mathbb{Z}_{>0},\ (\rho_{f}(z_{0}))^{n}\in K.

A question of Levy and Tucker

On the other hand, in the 2014 AIM workshop Postcritically Finite Maps In Complex And Arithmetic Dynamics, Levy [Lev14] and Tucker [Tuc14] asked the following question independently:

Question 1.8.

Let f:1()1()f:\mathbb{P}^{1}(\mathbb{C})\to\mathbb{P}^{1}(\mathbb{C}) be a non-exceptional rational map of degree d2d\geq 2 and let SS be the set of all multipliers of periodic points of ff. Take the subgroup of \mathbb{C}^{\ast} generated by S{0}S\setminus\{0\}. Is that true that the rank of this group is infinite?

It is not hard to see that our Corollary 1.5 gives a positive answer to (a generalized version of) Levy and Tucker’s question.

1.5. Sketch of the proofs

We have explained the proof of Theorem 1.2 before. Here we explain the proofs of Theorem 1.3 and Theorem 1.4.

We first give the idea of the proof of Theorem 1.4. We argue by contradiction and suppose that ff is not exceptional. The first step is to reduce to the case where ff is defined over ¯\overline{\mathbb{Q}}. This can be done using our Theorem 1.2. After enlarging KK, we may assume that ff is defined over KK. In the second step, we combine the arithmetic equidistribution theorem with a result of Zdunik [Zdu14] on the Lyapunov exponent to get a contradiction. This argument is inspired by Huguin’s proof of Theorem 1.7. Not like the case of Theorem 1.7, we can not apply the equidistribution theorem to the one dimensional dynamical system f:11f:\mathbb{P}^{1}\to\mathbb{P}^{1} directly. Our idea is to consider the two dimensional endomorphism F:=f×f¯F:=f\times\overline{f} on 1×1\mathbb{P}^{1}\times\mathbb{P}^{1} instead. More precisely, applying a result of Zdunik [Zdu14], we get a sequence (xn)n=1(x_{n})_{n=1}^{\infty} of distinct periodic points such that

limn+χf(xn)=a>f.\lim\limits_{n\to+\infty}\chi_{f}(x_{n})=a>\mathcal{L}_{f}.

Consider the endomorphism F:=f×f¯F:=f\times\overline{f} on 1×1\mathbb{P}^{1}\times\mathbb{P}^{1} and Γ:={pn=(zn,zn¯)}¯Zar1×1\Gamma:=\overline{\{p_{n}=(z_{n},\overline{z_{n}})\}}^{\rm Zar}\subseteq\mathbb{P}^{1}\times\mathbb{P}^{1}. By [GTZ11], the Dynamical Manin-Mumford conjecture holds for FF. Hence we may assume that Γ\Gamma is FF-invariant.

Let νn\nu_{n} be the discrete probability measure equally supported at the union of Galois orbits of iterates of pnp_{n} under FF. Then νn\nu_{n} converges weakly to the canonical measure μ\mu on Γ\Gamma with respect to FF by an equidistribution-type theorem (Theorem 3.1), which is a reformulation of [Yua08, Theorem 3.1], see Section 3 for details. Applying νnμ\nu_{n}\to\mu to the continuous test functions max{log|det(dF)|,A}{\rm max}\{{\rm log}\lvert{\rm det}(dF)\rvert,A\} (AA\in\mathbb{R}) and letting AA\to-\infty, we get

2alog|det(dF)|𝑑μ,2a\leq\int{\rm log}\lvert{\rm det}(dF)\rvert d\mu,

which is impossible since the right hand side equals to 2f<2a2\mathcal{L}_{f}<2a by a direct computation.

Next we sketch the proof of Theorem 1.3. According to [DH93], postcritically finite (PCF) maps are defined over ¯\overline{\mathbb{Q}} in the moduli space d\mathcal{M}_{d} of rational maps of degree dd, except for the family of flexible Lattès maps. So it suffices to consider the following two cases: 1). ff is defined over ¯\overline{\mathbb{Q}}, and 2). ff is not PCF. For the first case the conclusion follows from Theorem 1.4. For the second case, we need to develop some new techniques, which are presented in Section 5. In Section 5, we consider some pseudo linear algebra (which means that the domain may not be the whole vector space), and the vector space 𝔻(k)=k\mathbb{D}(k)_{\mathbb{Q}}=k^{*}\otimes_{\mathbb{Z}}\mathbb{Q} for a field kk of characteristic zero. We will actually prove a theorem (Theorem 5.6) stronger than the non-PCF case of Theorem 1.3, see Section 5 and 6 for details. To prove Theorem 5.6, in Section 6.1 we first deal with the case that ff is defined over ¯\overline{\mathbb{Q}}. A key ingredient in this step is [BGKT12, Lemma 4.1] which is a consequence of Siegel’s theorem on SS-integral points. The existence of a no preperiodic critical point is essentially used in here. In Section 6.2, we consider the general case and finish the proof. This is achieved by reducing to the case that ff is defined over ¯\overline{\mathbb{Q}} via an algebraic-geometric argument and techniques in Section 5.

1.6. Applications

The Zariski-dense orbit Conjecture

By applying Corollary 1.5 we can give a new proof of a special case of the Zariski-dense orbit conjecture.

Zariski-dense orbit Conjecture (=ZDO).

Let kk be an algebraically closed field of characteristic 0. Given an irreducible quasiprojective variety XX over kk and a dominant rational self-map ff on XX. If we have {gk(X):gf=g}=k\{g\in k(X):g\circ f=g\}=k where k(X)k(X) is the function field of XX, then there exists xX(k)x\in X(k) whose forward orbit under ff is well-defined and Zariski-dense in XX.

Remark 1.9.

The converse of ZDO is easy. For some progressions of ZDO, see e.g.[ABR11], [AC08], [MS14], [Xie17] and [Xie22].

As an application of Corollary 1.5, we give a new proof of (the most difficult part of) a special case of ZDO, which was firstly proved in [Xie22, Theorem 1.16].

Theorem 1.10.

Let X=1××1X=\mathbb{P}^{1}\times\cdots\times\mathbb{P}^{1} be the variety of product of NN copies of projective line over an algebraically closed field kk of characteristic 0. Suppose that f:XXf:X\to X is an endomorphism of form f1××fNf_{1}\times\cdots\times f_{N} where fj:11f_{j}:\mathbb{P}^{1}\to\mathbb{P}^{1} is a non-constant rational map for 1jN1\leq j\leq N. The ZDO holds for XX and ff.

Remark 1.11.

We note that every dominant endomorphism f:(1)N(1)Nf:(\mathbb{P}^{1})^{N}\to(\mathbb{P}^{1})^{N} over an algebraically closed field kk of characteristic zero must be of form f1××fNf_{1}\times\cdots\times f_{N}, after replacing ff by a suitable positive-integer iterate.

The original proof of Theorem 1.10 in [Xie22] relies on the solution of the (adelic) Zariski dense orbit conjecture on smooth projective surfaces [Xie22, Theorem 1.15], the notion of adelic topology introduced in [Xie22, Section 3] and a classification result on invariant subvarieties of f:(1)N(1)Nf:(\mathbb{P}^{1})^{N}\to(\mathbb{P}^{1})^{N} [Xie22, Proposition 9.2] (see also [MS14] and [GNY18]). When n=2n=2, Pakovich gave another proof [Pak23] using his classification of invariant curves in 1×1\mathbb{P}^{1}\times\mathbb{P}^{1} and some height argument. In our new proof, we don’t need the ingredients mentioned above.

A characterization of PCF maps

We also show that one can decide whether a rational map f:1()1()f:\mathbb{P}^{1}(\mathbb{C})\to\mathbb{P}^{1}(\mathbb{C}) of degree d2d\geq 2 is PCF with the information of its multiplier spectrum or length spectrum on periodic points.

Theorem 1.12.

Let f:1()1()f:\mathbb{P}^{1}(\mathbb{C})\to\mathbb{P}^{1}(\mathbb{C}) be a rational map of degree d2d\geq 2. Then the followings are equivalent:

(1) ff is PCF;

(2) ρf(x)¯\rho_{f}(x)\in\overline{\mathbb{Q}} for all xPer(f)()x\in{\rm Per}(f)(\mathbb{C}) and the \mathbb{Q}-subspace V=V(f)V=V(f) of \mathbb{R} is of finite dimension, where VV is generated over \mathbb{Q} by {log|NKx/(ρf(x))|:xPer(f)()}\{\log\lvert N_{K_{x}/\mathbb{Q}}(\rho_{f}(x))\rvert:x\in{\rm Per}^{*}(f)(\mathbb{C})\};

(3) |ρf(x)|¯\lvert\rho_{f}(x)\rvert\in\overline{\mathbb{Q}} for all xPer(f)()x\in{\rm Per}(f)(\mathbb{C}) and the \mathbb{Q}-subspace W=W(f)W=W(f) of \mathbb{R} is of finite dimension, where WW is generated over \mathbb{Q} by {log|NLx/(|ρf(x)|)|:xPer(f)()}\{\log\lvert N_{L_{x}/\mathbb{Q}}(\lvert\rho_{f}(x)\rvert)\rvert:x\in{\rm Per}^{*}(f)(\mathbb{C})\}.

Here KxK_{x} (resp. LxL_{x}) is any number field containing ρf(x)\rho_{f}(x) (resp. |ρf(x)|\lvert\rho_{f}(x)\rvert) and NKx/N_{K_{x}/\mathbb{Q}} (resp. NLx/N_{L_{x}/\mathbb{Q}}) is the norm map for the extension Kx/K_{x}/\mathbb{Q} (resp. Lx/L_{x}/\mathbb{Q}), i.e. the determinant of the \mathbb{Q}-linear transformation induced by multiplication by ρf(x)\rho_{f}(x) (resp. |ρf(x)||\rho_{f}(x)|).

Clearly, the subspaces V,WV,W above is independent of the choices of the fields Kx,LxK_{x},L_{x}, respectively.

The proofs of Theorem 1.10 and Theorem 1.12 will be given in Section 6.

Acknowledgement

The second-named author Junyi Xie would like to thank Thomas Gauthier, Vigny Gabriel, Charles Favre and Serge Cantat for helpful discussions.

The first-named author would like to thank Beijing International Center for Mathematical Research in Peking University for the invitation. The second and third-named authors Junyi Xie and Geng-Rui Zhang are supported by NSFC Grant (No.12271007).

2. Rational maps with algebraic lengths

Let KK be an algebraically closed subfield of \mathbb{C} which is invariant under the complex conjugate τ\tau i.e. τ(K)=K.\tau(K)=K. The aim of this section is the following result.

Theorem 2.1.

Let f:11f:\mathbb{P}_{\mathbb{C}}^{1}\to\mathbb{P}_{\mathbb{C}}^{1} be a rational map of degree d2d\geq 2. Assume that ff is not a flexible Lattès map and for every xPer(f)()x\in{\rm Per}(f)(\mathbb{C}), |ρf(x)|K|\rho_{f}(x)|\in K. Then ff is defined over K.K.

Applying Theorem 2.1 to the case K=¯K=\overline{\mathbb{Q}}, we get Theorem 1.2.

2.1. Weil restriction

Recall that KK is an algebraically closed field of \mathbb{C} such that τ(K)=K.\tau(K)=K. Set L:=Kτ=K.L:=K^{\tau}=K\cap\mathbb{R}. For example, if K=K=\mathbb{C}, then L=.L=\mathbb{R}. We need the following easy lemma.

Lemma 2.2.

We have K=L+iL,K=L+iL, in particular [K:L]=2.[K:L]=2.

Proof of Lemma 2.2.

Since KK is algebraically closed, iKi\in K. In particular, KL.K\neq L. For every uKu\in K, we may write

u=u+τ(u)2+uτ(u)2iiu=\frac{u+\tau(u)}{2}+\frac{u-\tau(u)}{2i}i

and both u+τ(u)2\frac{u+\tau(u)}{2} and uτ(u)2i\frac{u-\tau(u)}{2i} are contained in LL. This concludes the proof. ∎

We briefly recall the notion of Weil restriction. See [Poo17, Section 4.6] and [BLR90, Section 7.6] for more information.

Denote by Var/KVar_{/K} (resp. Var/LVar_{/L}) the category of varieties over KK (resp. LL). For every variety XX over KK, there is a unique variety R(X)R(X) over LL represents the functor Var/LSetsVar_{/L}\to Sets sending VVar/LV\in Var_{/L} to Hom(VLK,X).{\rm Hom}(V\otimes_{L}K,X). It is called the Weil restriction of XX. The functor XR(X)X\mapsto R(X) is called the Weil restriction. One has the canonical morphism ψK:X(K)R(X)(L).\psi_{K}:X(K)\to R(X)(L). When K=K=\mathbb{C}, this map is a real analytic diffeomorphism. One may view X(K)X(K) as an LL-algebraic variety via ψK.\psi_{K}.

Definition 2.3.

The LL-Zariski topology on X(K)X(K) is the restriction of the Zariski topology on R(X)R(X) via ψK\psi_{K}. A subset YY of X(K)X(K) is LL-algebraic if it is closed in the LL-Zariski topology. When K=K=\mathbb{C}, the LL-Zariski topology is exactly the real Zariski topology as in [JX23b, Section 8.1.1].

By (iii) of Proposition 2.5 below, the LL-Zariski topology is stronger than the Zariski topology on X(K).X(K).

When K=K=\mathbb{C}, roughly speaking, the Weil restriction is just constructed by splitting a complex variable zz into two real variables x,yx,y via z=x+iyz=x+iy. For the convenience of the reader, in the following example, we show the concrete construction of R(X)R(X) when XX is affine.

Example 2.4.

First assume that X=𝔸KNX=\mathbb{A}^{N}_{K}. Then R(X)=𝔸L2N.R(X)=\mathbb{A}^{2N}_{L}. The map

ψK:𝔸LN(L)=KN𝔸L2N(L)=2N\psi_{K}:\mathbb{A}^{N}_{L}(L)=K^{N}\to\mathbb{A}^{2N}_{L}(L)=\mathbb{R}^{2N}

sends (z1,,zN)(z_{1},\dots,z_{N}) to (x1,y1,x2,y2,,xN,yN)(x_{1},y_{1},x_{2},y_{2},\dots,x_{N},y_{N}) where zj=xj+iyjz_{j}=x_{j}+iy_{j}.

Consider the algebra 𝔹:=K[I]/(I2+1)KIK\mathbb{B}:=K[I]/(I^{2}+1)\simeq K\oplus IK. Every fK[z1,,zN]f\in K[z_{1},\dots,z_{N}] defines an element

F:=f(x1+Iy1,,xN+IyN)𝔹[x1,y1,,xN,yN].F:=f(x_{1}+Iy_{1},\dots,x_{N}+Iy_{N})\in\mathbb{B}[x_{1},y_{1},\dots,x_{N},y_{N}].

Since

𝔹[x1,y1,,xN,yN]=K[x1,y1,,xN,yN]IK[x1,y1,,xN,yN],\mathbb{B}[x_{1},y_{1},\dots,x_{N},y_{N}]=K[x_{1},y_{1},\dots,x_{N},y_{N}]\oplus IK[x_{1},y_{1},\dots,x_{N},y_{N}],

FF can be uniquely decomposed to

F=r(f)+Ii(f)F=r(f)+Ii(f)

where r(f),i(f)K[x1,y1,,xN,yN].r(f),i(f)\in K[x_{1},y_{1},\dots,x_{N},y_{N}].

More generally, if XX is the closed subvariety of 𝔸KN=SpecK[z1,,zM]\mathbb{A}^{N}_{K}={\rm Spec}K[z_{1},\dots,z_{M}] defined by the ideal (f1,,fs)(f_{1},\dots,f_{s}), then R(X)R(X) is the closed subvariety of

R(𝔸KN)=𝔸L2N=SpecL[x1,y1,,xN,yN]R(\mathbb{A}^{N}_{K})=\mathbb{A}^{2N}_{L}={\rm Spec}\,L[x_{1},y_{1},\dots,x_{N},y_{N}]

defined by the ideal generated by r(f1),i(f1),,r(fs),i(fs)r(f_{1}),i(f_{1}),\dots,r(f_{s}),i(f_{s}).

We list some basic properties of Weil restriction without proof.

Propsition 2.5.

Let X,YVar/KX,Y\in Var_{/K}, then we have the following properties:

  • if XX is irreducible, then R(X)R(X) is irreducible;

  • dimR(X)=2dimX;\dim R(X)=2\dim X;

  • if f:YXf:Y\to X is a closed (resp. open) immersion, then the induced morphism R(f):R(Y)R(X)R(f):R(Y)\to R(X) is a closed (resp. open) immersion.

We still denote by τ\tau the restriction of τ\tau to K.K. Denote by XτX^{\tau} the base change of XX by the field extension τ:KK\tau:K\to K. This induces a morphism of schemes (over \mathbb{Z}) τ:XτX\tau:X^{\tau}\to X. It is not a morphism of schemes over KK. It is clear that (Xτ)τ=X.(X^{\tau})^{\tau}=X.

Example 2.6.

If XX is the subvariety of 𝔸KN=SpecK[z1,,zN]\mathbb{A}^{N}_{K}={\rm Spec}K[z_{1},\dots,z_{N}] defined by the equations Iai,IzI=0,i=1,,s\sum_{I}a_{i,I}z^{I}=0,i=1,\dots,s Then XτX^{\tau} is the subvariety of 𝔸KN\mathbb{A}^{N}_{K} defined by Iτ(ai,I)zI=0,i=1,,s\sum_{I}\tau(a_{i,I})z^{I}=0,i=1,\dots,s. The map τ:X=(Xτ)τXτ\tau:X=(X^{\tau})^{\tau}\to X^{\tau} sends a point (z1,,zN)X(K)(z_{1},\dots,z_{N})\in X(K) to (τ(z1),,τ(zN))Xτ(K)(\tau(z_{1}),\dots,\tau(z_{N}))\in X^{\tau}(K).

The following result due to Weil is useful for computing the Weil restriction.

Propsition 2.7.

[Poo17, Exercise 4.7] We have a canonical isomorphism

R(X)LKX×Xτ.R(X)\otimes_{L}K\simeq X\times X^{\tau}.

Under this isomorphism,

R(X)(L)={(z1,z2)X(K)×Xτ(K)|z2=τ(z1)}R(X)(L)=\{(z_{1},z_{2})\in X(K)\times X^{\tau}(K)|\,\,z_{2}=\tau(z_{1})\}

and ψK\psi_{K} sends zX(K)z\in X(K) to (z,τ(z))R(X)(L).(z,\tau(z))\in R(X)(L).

2.2. Admissible subsets

In this section,we recall the notion of admissible subsets on real algebraic varieties introduced in [JX23b].

Let XX be a variety over \mathbb{R}.

Definition 2.8.

[JX23b, Section 8.2] A closed subset VV of X()X(\mathbb{R}) is called admissible if there is a morphism f:YXf:Y\to X of real algebraic varieties and a Zariski closed subset VYV^{\prime}\subseteq Y such that V=f(V())V=f(V^{\prime}(\mathbb{R})) and ff is étale at every point in V().V^{\prime}(\mathbb{R}).

In particular, every algebraic subset of X()X(\mathbb{R}) is admissible.

Remark 2.9.

Denote by JJ the non-étale locus for ff in VV. We have JV()=.J\cap V(\mathbb{R})=\emptyset. Since we may replace VV by VJV\setminus J, in the above definition we may further assume that ff is étale.

Propsition 2.10.

[JX23b, Remarks 8.14, 8.15 and Proposition 8.16] We have the following basic properties:

  • (1)

    Let YY be a Zariski closed subset of XX. If VV is admissible as a subset of X()X(\mathbb{R}), then VYV\cap Y is admissible as a subset of Y()Y(\mathbb{R}).

  • (2)

    An admissible subset is semialgebraic.

  • (3)

    Let V1,V2V_{1},V_{2} be two admissible closed subsets of X()X(\mathbb{R}). Then V1V2V_{1}\cap V_{2} is admissible.

The following theorem shows that admissible subsets satisfy the descending chain condition.

Theorem 2.11.

[JX23b, Theorem 8.17] Let Vn,n0V_{n},n\geq 0 be a sequence of decreasing admissible subsets of X()X(\mathbb{R}). Then there is N0N\geq 0 such that Vn=VNV_{n}=V_{N} for all nN.n\geq N.

2.3. Transcendental points

Let XKX_{K} be a variety over KK and X:=XKKX:=X_{K}\otimes_{K}\mathbb{C}. We think that XKX_{K} as a model of XX over K.K.

Denote by πK:XXK\pi_{K}:X\to X_{K} the natural projection. For any point xX()x\in X(\mathbb{C}), define Z(x)KZ(x)_{K} to be the Zariski closure of πK(x)\pi_{K}(x) and Z(x):=πK1(Z(x)K)Z(x):=\pi_{K}^{-1}(Z(x)_{K}). It is clear that Z(x)Z(x) is irreducible. We call Z(x)Z(x) the /K\mathbb{C}/K-closure of xx w.r.t the model XK.X_{K}. We say that xx is transcendental if dimZ(x)1\dim Z(x)\geq 1 and call dimZ(x)\dim Z(x) the transcendental degree of x.x.

The notion of transcendental points (on curves) was introduced in [XY23, Section 4.1] and it plays important role in [XY23] on the geometric Bombieri-Lang conjecture and [JX23a] on the dynamical André-Oort conjecture . Roughly speaking, a very general point in Z(x)Z(x) satisfies the same algebraic properties as x.x. In this paper, we study lengths of periodic points in whose definition we need the norm map ||:0|\cdot|:\mathbb{C}\to\mathbb{R}_{\geq 0} which is not algebraic. However ||2:|\cdot|^{2}:\mathbb{C}\to\mathbb{R} is real algebraic. For this reason we need to generalize the above notions to respect the real structure.

The Weil restriction R(X)R(X) of XX w.r.t. /\mathbb{C}/\mathbb{R} is a real algebraic variety. We have R(X)=R(XK)L.R(X)=R(X_{K})\otimes_{L}\mathbb{R}. Denote by πL:R(X)R(XK)\pi_{L}:R(X)\to R(X_{K}) the natural projection. For every xX()x\in X(\mathbb{C}), let Y(x)LY(x)_{L} be the Zariski closure of πL(ψ(x))\pi_{L}(\psi_{\mathbb{C}}(x)) and Y(x):=πL1(Y(x)L)Y(x):=\pi_{L}^{-1}(Y(x)_{L}). Set Z(x):=ψ1(Y(x)())Z^{\mathbb{R}}(x):=\psi_{\mathbb{C}}^{-1}(Y(x)(\mathbb{R})) which is a real Zariski closed subset of X().X(\mathbb{C}).

We now give a more concrete description of Z(x)Z(x) and Z(x)Z^{\mathbb{R}}(x). Let UKU_{K} be an affine open neighborhood of πK(x)\pi_{K}(x). Set U:=πK1(UK)=UKKU:=\pi^{-1}_{K}(U_{K})=U_{K}\otimes_{K}\mathbb{C}. We have a natural embedding πK:𝒪(UK)𝒪(U)\pi_{K}^{*}:\mathcal{O}(U_{K})\hookrightarrow\mathcal{O}(U). We can view elements in 𝒪K(U):=πK(𝒪(UK))\mathcal{O}^{K}(U):=\pi_{K}^{*}(\mathcal{O}(U_{K})) as the algebraic functions on U()U(\mathbb{C}) defined over K.K. Then we have

Z(x)U={yU|h(y)=0 for every h𝒪K(U) with h(x)=0}Z(x)\cap U=\{y\in U|\,\,h(y)=0\text{ for every }h\in\mathcal{O}^{K}(U)\text{ with }h(x)=0\}

and Z(x)Z(x) is the Zariski closure of Z(x)U.Z(x)\cap U.

As 𝒪(R(U))=𝒪(R(U))\mathcal{O}(R(U)_{\mathbb{C}})=\mathcal{O}(R(U))\otimes_{\mathbb{R}}\mathbb{C}, every h𝒪(R(U))h\in\mathcal{O}(R(U)_{\mathbb{C}}) can be viewed as a \mathbb{C}-valued algebraic function on R(U)().R(U)(\mathbb{R}). Every h𝒪(R(U))h\in\mathcal{O}(R(U)_{\mathbb{C}}) induces a function hψh\circ\psi_{\mathbb{C}} on U().U(\mathbb{C}). The functions of this form are exactly the \mathbb{C}-valued real algebraic functions on U()U(\mathbb{C}). Denote by 𝒞alg(U)\mathcal{C}^{\mathbb{R}-{\rm alg}}(U) the \mathbb{R}-algebra of \mathbb{C}-valued real algebraic functions on U().U(\mathbb{C}). Since algebraic functions are real algebraic, we have a natural embedding 𝒪(U)𝒞alg(U).\mathcal{O}(U)\subseteq\mathcal{C}^{\mathbb{R}-{\rm alg}}(U). By Proposition 2.7, we have

𝒞alg(U)𝒪(U)τ(𝒪(U)).\mathcal{C}^{\mathbb{R}-{\rm alg}}(U)\simeq\mathcal{O}(U)\otimes_{\mathbb{C}}\tau(\mathcal{O}(U)).

Let 𝒪L(R(U)):=πL(𝒪(R(UK)))\mathcal{O}^{L}(R(U)):=\pi_{L}^{*}(\mathcal{O}(R(U_{K}))) be the set of algebraic functions defined over LL on R(U).R(U). Let 𝒞alg,L(U)\mathcal{C}^{\mathbb{R}-{\rm alg},L}(U) the image of 𝒪L(R(U))LK\mathcal{O}^{L}(R(U))\otimes_{L}K in 𝒞alg(U)\mathcal{C}^{\mathbb{R}-{\rm alg}}(U), which is the set of \mathbb{C}-valued real algebraic functions on U()U(\mathbb{C}) defined over L.L. It is clear that 𝒪K(U)𝒞alg,L(U).\mathcal{O}^{K}(U)\subseteq\mathcal{C}^{\mathbb{R}-{\rm alg},L}(U). By Proposition 2.7, we have

𝒞alg,L(U)𝒪K(U)Kτ(𝒪K(U)).\mathcal{C}^{\mathbb{R}-{\rm alg},L}(U)\simeq\mathcal{O}^{K}(U)\otimes_{K}\tau(\mathcal{O}^{K}(U)).

We have

Z(x)U()={yU()|h(y)=0 for every h𝒞alg,L(U) with h(x)=0}Z^{\mathbb{R}}(x)\cap U(\mathbb{C})=\{y\in U(\mathbb{C})|\,\,h(y)=0\text{ for every }h\in\mathcal{C}^{\mathbb{R}-{\rm alg},L}(U)\text{ with }h(x)=0\}

and Z(x)Z^{\mathbb{R}}(x) is the real Zariski closure of Z(x)U().Z^{\mathbb{R}}(x)\cap U(\mathbb{C}). This implies the following lemma.

Lemma 2.12.

Let fK:XKXKf_{K}:X^{\prime}_{K}\to X_{K} be a morphisms between KK-varieties. Set X:=XKKX^{\prime}:=X^{\prime}_{K}\otimes_{K}\mathbb{C} and let f:XXf:X^{\prime}\to X be the morphism induced by f.f. Let xX()x^{\prime}\in X^{\prime}(\mathbb{C}) and xX()x\in X(\mathbb{C}) with f(x)=xf(x^{\prime})=x. Then we have f(Z(x))Z(x).f(Z^{\mathbb{R}}(x^{\prime}))\subseteq Z^{\mathbb{R}}(x).

Lemma 2.13.

We have Z(x)Z(x)Z^{\mathbb{R}}(x)\subseteq Z(x) and Z(x)Z^{\mathbb{R}}(x) is Zariski dense in Z(x).Z(x). In particular, if xx is transcendental, then dimZ(x)>1.\dim_{\mathbb{R}}Z^{\mathbb{R}}(x)>1.

Proof.

It is clear that Z(x)Z(x)Z^{\mathbb{R}}(x)\subseteq Z(x). After replacing XKX_{K} by an affine open neighborhood of πK(x)\pi_{K}(x). We may assume that XK,XX_{K},X are affine. Let h𝒪(X)h\in\mathcal{O}(X) such that h(Z(x))=0.h(Z^{\mathbb{R}}(x))=0. Let ej,jJe_{j},j\in J be a KK-basis of \mathbb{C}. We may assume that 0J0\in J and 1=e0.1=e_{0}. Write h1=jJgjejh\otimes_{\mathbb{C}}1=\sum_{j\in J}g_{j}e_{j}. Then gj𝒞alg,L(X)g_{j}\in\mathcal{C}^{\mathbb{R}-{\rm alg},L}(X) and gj(x)=0g_{j}(x)=0.

Let fn,nNf_{n},n\in N be a KK-basis of 𝒪K(X).\mathcal{O}^{K}(X). We may assume that 0N0\in N and 1=f0.1=f_{0}. Write

gj=m,nNbj,m,nfmτ(fn).g_{j}=\sum_{m,n\in N}b_{j,m,n}f_{m}\otimes\tau(f_{n}).

The we get

h1=jJ,m,nNbj,m,nejfmτ(fn).h\otimes_{\mathbb{C}}1=\sum_{j\in J,m,n\in N}b_{j,m,n}e_{j}f_{m}\otimes\tau(f_{n}).

As ejfmτ(fn),jJ,m,nNe_{j}f_{m}\otimes\tau(f_{n}),j\in J,m,n\in N forms a KK-basis of 𝒞alg(X)\mathcal{C}^{\mathbb{R}-{\rm alg}}(X), we have

bj,m,n=0b_{j,m,n}=0

for every n0.n\neq 0. So gj=mNbj,m,0fm1𝒪K(X).g_{j}=\sum_{m\in N}b_{j,m,0}f_{m}\otimes 1\in\mathcal{O}^{K}(X). Since gj(x)=0g_{j}(x)=0, gj|Z(x)=0.g_{j}|_{Z(x)}=0. Then we have h|Z(x)=0h|_{Z(x)}=0 which concludes the proof. ∎

Lemma 2.14.

Assume that XKX_{K} is affine. Let h𝒞alg,L(X)h\in\mathcal{C}^{\mathbb{R}-{\rm alg},L}(X). For xX()x\in X(\mathbb{C}), if h(x)Kh(x)\in K, then hh is constant on Z(x).Z^{\mathbb{R}}(x).

Proof.

Write h=gψh=g\circ\psi_{\mathbb{C}} where g𝒪L(R(X))LKg\in\mathcal{O}^{L}(R(X))\otimes_{L}K. Write g=πL(g1)+πL(g2)ig=\pi_{L}^{*}(g_{1})+\pi_{L}^{*}(g_{2})i where g1,g2𝒪(R(XK)).g_{1},g_{2}\in\mathcal{O}(R(X_{K})). Since h(x)Kh(x)\in K, πL(g1)(ψ(x)),πL(g2)(ψ(x))L.\pi_{L}^{*}(g_{1})(\psi(x)),\pi_{L}^{*}(g_{2})(\psi(x))\in L. The map πL|ψ(x):ψ(x)Y(x)L\pi_{L}|_{\psi(x)}:\psi(x)\to Y(x)_{L} induces an embedding 𝒪(Y(x)L)L.\mathcal{O}(Y(x)_{L})\hookrightarrow L. The image of gi|Y(x)L,i=1,2g_{i}|_{Y(x)_{L}},i=1,2 are contained in LL. Hence gi|Y(x)L,i=1,2g_{i}|_{Y(x)_{L}},i=1,2 are contained in L.L. This implies that πL(g1),πL(g2)\pi_{L}^{*}(g_{1}),\pi_{L}^{*}(g_{2}) are constant on Y(x)(L)Y(x)(L), hence hh is constant on Z(x).Z^{\mathbb{R}}(x). This concludes the proof. ∎

2.4. Moduli space of rational maps

For d2,d\geq 2, let Ratd\text{Rat}_{d} be the space of degree dd endomorphisms on 1\mathbb{P}^{1}. It is a smooth quasi-projective variety of dimension 2d+12d+1 [Sil12]. Let FLdRatdFL_{d}\subseteq\text{Rat}_{d} be the locus of flexible Lattès maps, which is Zariski closed in Ratd\text{Rat}_{d}. The group PGL2=Aut(1){\rm PGL}_{2}={\rm Aut}(\mathbb{P}^{1}) acts on Ratd\text{Rat}_{d} by conjugacy. The geometric quotient

d:=Ratd/PGL2\mathcal{M}_{d}:=\text{Rat}_{d}/{\rm PGL}_{2}

is the (coarse) moduli space of endomorphisms of degree dd [Sil12]. The moduli space d=Spec(𝒪(Ratd)PGL2)\mathcal{M}_{d}={\rm Spec}(\mathcal{O}(\text{Rat}_{d})^{{\rm PGL}_{2}}) is an affine variety of dimension 2d22d-2 [Sil07, Theorem 4.36(c)]. Let Ψ:Ratdd\Psi:\text{Rat}_{d}\to\mathcal{M}_{d} be the quotient morphism. Set [FLd]:=Ψ(FLd).[FL_{d}]:=\Psi(FL_{d}). The above construction works over any algebraically closed field of characteristic 0 and commutes with base changes.

For every n>0n\in\mathbb{Z}_{>0}, let Pern(fRatd){\rm Per}_{n}(f_{\text{Rat}_{d}}) be the closed subvariety of Ratd×1\text{Rat}_{d}\times\mathbb{P}^{1} of the nn-periodic points of fRatd.f_{\text{Rat}_{d}}. Let ϕn:Pern(fRatd)Ratd\phi_{n}:{\rm Per}_{n}(f_{\text{Rat}_{d}})\to\text{Rat}_{d} be the first projection. It is a finite map of degree dn+1.d^{n}+1. Let λn:Pern(fRatd)𝔸1\lambda_{n}:{\rm Per}_{n}(f_{\text{Rat}_{d}})\to\mathbb{A}^{1} be the morphism (ft,x)dftn(x)𝔸1(f_{t},x)\mapsto df_{t}^{n}(x)\in\mathbb{A}^{1}. View Pern(fRatd){\rm Per}_{n}(f_{\text{Rat}_{d}}) as the moduli space of endomorphisms of degree dd with a marked nn-periodic point. We also denote it by Ratd[n]\text{Rat}_{d}[n] or Ratd1[n]\text{Rat}^{1}_{d}[n].

Let s1,,sns_{1},\dots,s_{n} be a sequence of elements in 0\mathbb{Z}_{\geq 0} with s1sns_{1}\leq\dots\leq s_{n} and sidi!+1.s_{i}\leq d^{i!}+1. We construct the space Rd(s1,,sn)R_{d}(s_{1},\dots,s_{n}) of rational functions of degree dd with sns_{n} marked n!n!-periodic points (counting with multiplicities) and in which there are sn1s_{n-1} (n1)!(n-1)!-periodic points (counting with multiplicities) …and in which there are s1s_{1} 11-periodic points (counting with multiplicities) as follows: Consider the fiber product (Ratd[n!])/Ratdsn(\text{Rat}_{d}[n!])^{s_{n}}_{/\text{Rat}_{d}} of sns_{n} copies of Ratd[n!]\text{Rat}_{d}[n!] over Ratd.\text{Rat}_{d}. For ij{1,,dn!+1}i\neq j\in\{1,\dots,d^{n!}+1\}, let πi,j:(Ratd[n!])/Ratdsn(Ratd[n!])/Ratd2\pi_{i,j}:(\text{Rat}_{d}[n!])^{s_{n}}_{/\text{Rat}_{d}}\to(\text{Rat}_{d}[n!])^{2}_{/\text{Rat}_{d}} be the projection to the i,ji,j coordinates. The diagonal Δ(Ratd[n!])/Ratd2\Delta\subseteq(\text{Rat}_{d}[n!])^{2}_{/\text{Rat}_{d}} is an irreducible component of (Ratd[n!])/Ratd2(\text{Rat}_{d}[n!])^{2}_{/\text{Rat}_{d}}. Consider the open subset

U:=(Ratd[n!])/Ratdsn(ij{1,,dn!+1}πi,j1(Δ)).U:=(\text{Rat}_{d}[n!])^{s_{n}}_{/\text{Rat}_{d}}\setminus(\cup_{i\neq j\in\{1,\dots,d^{n!}+1\}}\pi_{i,j}^{-1}(\Delta)).

Let UU^{\prime} be the subset of UU of points (f,x1,,xsn)(f,x_{1},\dots,x_{s_{n}}) satisfying fm!(xi)=xif^{m!}(x_{i})=x_{i} for every m=1,,nm=1,\dots,n and i=1,,smi=1,\dots,s_{m}. This set is open and closed in U.U. We then define Rd(s1,,sn)R_{d}(s_{1},\dots,s_{n}) to be the Zariski closure of UU^{\prime} in (Ratd[n!])/Ratdsn.(\text{Rat}_{d}[n!])^{s_{n}}_{/\text{Rat}_{d}}. For mnm\leq n, define ϕn,m:Rd(s1,,sn)Rd(s1,,sm)\phi_{n,m}:R_{d}(s_{1},\dots,s_{n})\to R_{d}(s_{1},\dots,s_{m}) the morphism (f,x1,,xsn)(f,x1,,xsm).(f,x_{1},\dots,x_{s_{n}})\mapsto(f,x_{1},\dots,x_{s_{m}}). Moreover, denote by ϕn,0:Rd(s1,,sn)Ratd\phi_{n,0}:R_{d}(s_{1},\dots,s_{n})\to{\rm Rat}_{d} the morphism (f,x1,,xsn)f.(f,x_{1},\dots,x_{s_{n}})\mapsto f. For m1m2nm_{1}\leq m_{2}\leq n, we have ϕm2,m1ϕn,m2=ϕn,m1.\phi_{m_{2},m_{1}}\circ\phi_{n,m_{2}}=\phi_{n,m_{1}}. Let λs1,,sn:Rd(s1,,sn)𝔸sn\lambda_{s_{1},\dots,s_{n}}:R_{d}(s_{1},\dots,s_{n})\to\mathbb{A}^{s_{n}} the morphism defined by

(f,x1,,xsn)(dfn!(x1),,dfn!(xsn)).(f,x_{1},\dots,x_{s_{n}})\mapsto(df^{n!}(x_{1}),\dots,df^{n!}(x_{s_{n}})).

Since ϕn\phi_{n} is étale at every point xPern(fRatd)λs1,,sn1(1),x\in{\rm Per}_{n}(f_{\text{Rat}_{d}})\setminus\lambda_{s_{1},\dots,s_{n}}^{-1}(1), ϕn,0\phi_{n,0} is étale at every point x(λs1,,sn)1((𝔸1{1})sn).x\in(\lambda_{s_{1},\dots,s_{n}})^{-1}((\mathbb{A}^{1}\setminus\{1\})^{s_{n}}).

Define d(s1,,sn):=Rd(s1,,sn)/PGL2\mathcal{M}_{d}(s_{1},\dots,s_{n}):=R_{d}(s_{1},\dots,s_{n})/{\rm PGL}_{2} to be the moduli space of endomorphisms of degree dd on 1\mathbb{P}^{1} with sns_{n} marked n!n!-periodic points (counting with multiplicities) and in which there are sn1s_{n-1} (n1)!(n-1)!-periodic points (counting with multiplicities) …and in which there are s1s_{1} 11-periodic points (counting with multiplicities). The morphisms ϕn,m\phi_{n,m}, λs1,,sn\lambda_{s_{1},\dots,s_{n}} descent to [ϕn,m]:d(s1,,sn)d(s1,,sm)[\phi_{n,m}]:\mathcal{M}_{d}(s_{1},\dots,s_{n})\to\mathcal{M}_{d}(s_{1},\dots,s_{m}) when m=1,,nm=1,\dots,n, [ϕn,0]:d(s1,,sn)d[\phi_{n,0}]:\mathcal{M}_{d}(s_{1},\dots,s_{n})\to\mathcal{M}_{d} and [λs1,,sn]:d(s1,,sn)𝔸sn.[\lambda_{s_{1},\dots,s_{n}}]:\mathcal{M}_{d}(s_{1},\dots,s_{n})\to\mathbb{A}^{s_{n}}. Then [ϕn,0][\phi_{n,0}] is étale at every point x[λs1,,sm]1((𝔸1{1})sn).x\in[\lambda_{s_{1},\dots,s_{m}}]^{-1}((\mathbb{A}^{1}\setminus\{1\})^{s_{n}}).

2.5. Length maps

For d2,d\geq 2, let s1,,sns_{1},\dots,s_{n} be a sequence of elements in 0\mathbb{Z}_{\geq 0} with s1sns_{1}\leq\dots\leq s_{n} and sidi!+1.s_{i}\leq d^{i!}+1. Let

|λs1,,sn|:d(s1,,sn)()0sn|\lambda_{s_{1},\dots,s_{n}}|:\mathcal{M}_{d}(s_{1},\dots,s_{n})(\mathbb{C})\to\mathbb{R}_{\geq 0}^{s_{n}}

be the composition of

[λs1,,sn]:d(s1,,sn)()sn[\lambda_{s_{1},\dots,s_{n}}]:\mathcal{M}_{d}(s_{1},\dots,s_{n})(\mathbb{C})\to\mathbb{C}^{s_{n}}

and the norm map

(a1,,asn)sn(|a1|,,|asn|)0sn.(a_{1},\dots,a_{s_{n}})\in\mathbb{C}^{s_{n}}\mapsto(|a_{1}|,\dots,|a_{s_{n}}|)\in\mathbb{R}_{\geq 0}^{s_{n}}.

Define

qs1,,sn:d(s1,,sn)()0snq_{s_{1},\dots,s_{n}}:\mathcal{M}_{d}(s_{1},\dots,s_{n})(\mathbb{C})\to\mathbb{R}_{\geq 0}^{s_{n}}

be the composition of

|λs1,,sn|:d(s1,,sn)()0sn|\lambda_{s_{1},\dots,s_{n}}|:\mathcal{M}_{d}(s_{1},\dots,s_{n})(\mathbb{C})\to\mathbb{R}_{\geq 0}^{s_{n}}

and the map

(a1,,asn)0sn(a12,,asn2)0sn.(a_{1},\dots,a_{s_{n}})\in\mathbb{R}_{\geq 0}^{s_{n}}\mapsto(a_{1}^{2},\dots,a_{s_{n}}^{2})\in\mathbb{R}_{\geq 0}^{s_{n}}.

It is clear that

qs1,,sn𝒞alg,L(d(s1,,sn)()).q_{s_{1},\dots,s_{n}}\in\mathcal{C}^{\mathbb{R}-{\rm alg},L}(\mathcal{M}_{d}(s_{1},\dots,s_{n})(\mathbb{C})).

Here the model of d(s1,,sn)\mathcal{M}_{d}(s_{1},\dots,s_{n})_{\mathbb{C}} over KK is taken to be d(s1,,sn)K.\mathcal{M}_{d}(s_{1},\dots,s_{n})_{K}.

By Lemma 2.14, for every xd(s1,,sn)()x\in\mathcal{M}_{d}(s_{1},\dots,s_{n})(\mathbb{C}), if qs1,,sn(x)Lsnq_{s_{1},\dots,s_{n}}(x)\in L^{s_{n}}, then qs1,,sn|V(x)q_{s_{1},\dots,s_{n}}|_{V^{\mathbb{R}}(x)} is constant. Hence for every xd(s1,,sn)()x\in\mathcal{M}_{d}(s_{1},\dots,s_{n})(\mathbb{C}), if |λs1,,sn|(x)Lsn|\lambda_{s_{1},\dots,s_{n}}|(x)\in L^{s_{n}}, then |λs1,,sn||V(x)|\lambda_{s_{1},\dots,s_{n}}||_{V^{\mathbb{R}}(x)} is constant.

2.6. Rigidity of length spectrum

In this section, we recall the rigidity of length spectrum proved by Ji and Xie [JX23b].

Let ff be an endomorphism of 1()\mathbb{P}^{1}(\mathbb{C}) of degree d2d\geq 2. As in [JX23b, Section 8.3] the length spectrum L(f)={L(f)n,n1}L(f)=\{L(f)_{n},n\geq 1\} of ff is a sequence of finite multisets111A multiset is a set except allowing multiple instances for each of its elements. The number of the instances of an element is called the multiplicity. For example: {a,a,b,c,c,c}\{a,a,b,c,c,c\} is a multiset of cardinality 66, the multiplicities for a,b,ca,b,c are 2,1,3, respectively., where L(f)n:=Ln(f)L(f)_{n}:=L_{n}(f) is the multiset of norms of multipliers of all fixed points of fn.f^{n}. In particular, L(f)L(f) is a multiset of non-negative real numbers of cardinality dn+1d^{n}+1. For every n0n\geq 0, let RL(f)nRL(f)_{n} be the sub-multiset of L(f)nL(f)_{n} consisting of all elements >1.>1. We call RL(f):={RL(f)n,n1}RL(f):=\{RL(f)_{n},n\geq 1\} the repelling length spectrum of ff and RL(f):={RL(f)n:=RL(f)n!,n1}RL^{*}(f):=\{RL^{*}(f)_{n}:=RL(f)_{n!},n\geq 1\} the main repelling length spectrum of ff. We have dn+1#RL(f)ndn+1Md^{n}+1\geq\#RL(f)_{n}\geq d^{n}+1-M for some M0M\geq 0. It is clear that the difference dn!+1#RL(f)nd^{n!}+1-\#RL^{*}(f)_{n} is increasing and bounded. As L(f),RL(f)L(f),RL(f) and RL(f)RL^{*}(f) are invariant under conjugacy, they descent on d()\mathcal{M}_{d}(\mathbb{C}). For every [f]d()[f]\in\mathcal{M}_{d}(\mathbb{C}), define L([f]):=L(f),RL([f]):=RL(f)L([f]):=L(f),RL([f]):=RL(f) and RL([f]):=RL(f)RL^{*}([f]):=RL^{*}(f) for any ff in the class [f][f].

Let Ω\Omega be the set of sequences An,n1A_{n},n\geq 1 of multisets consisting of real numbers of norm strictly larger than 11 satisfying #Andn!+1\#A_{n}\leq d^{n!}+1 and for every aAna\in A_{n} with multiplicity mm, an+1An+1a^{n+1}\in A_{n+1} with multiplicity at least mm. For A,BΩA,B\in\Omega, we write ABA\subseteq B if AnBnA_{n}\subseteq B_{n} for every n1n\geq 1. An element A=(An)ΩA=(A_{n})\in\Omega is called big if dn!+1#And^{n!}+1-\#A_{n} is bounded. For every endomorphism ff of 1()\mathbb{P}^{1}(\mathbb{C}) of degree dd, we have RL(f)ΩRL^{*}(f)\in\Omega and it is big.

Theorem 2.15.

[JX23b, Theorem 8.25] If AΩA\in\Omega is big, then the set

{fd()[FLd]|ARL(f)}\{f\in\mathcal{M}_{d}(\mathbb{C})\setminus[FL_{d}]|\,\,A\subseteq RL^{*}(f)\}

is finite.

2.7. Proof of Theorem 2.1

Let f:11f:\mathbb{P}_{\mathbb{C}}^{1}\to\mathbb{P}_{\mathbb{C}}^{1} be a rational map of degree d2d\geq 2. Assume that ff is not a flexible Lattès map and for every xPer(f)()x\in{\rm Per}(f)(\mathbb{C}), |ρf(x)|K|\rho_{f}(x)|\in K. We want to show that [f]d()[f]\in\mathcal{M}_{d}(\mathbb{C}) is not transcendental over KK for the model (d)K.(\mathcal{M}_{d})_{K}. Now assume that [f][f] is transcendental.

Set A:=RL(f)ΩA:=RL^{*}(f)\in\Omega, which is big. Set sn:=#An.s_{n}:=\#A_{n}. We may pick a sequence of periodic points xi,i1x_{i},i\geq 1 such that for every n1n\geq 1, x1,,xsnx_{1},\dots,x_{s_{n}} are fixed by fn!f^{n!} and An={|ρ(xi)|n!,i=1,,sn}.A_{n}=\{|\rho(x_{i})|^{n!},i=1,\dots,s_{n}\}. Let [fn](s1,,sn)()[f_{n}]\in\mathcal{M}(s_{1},\dots,s_{n})(\mathbb{C}) be the point presented by (f,x1,,xsn).(f,x_{1},\dots,x_{s_{n}}). It is clear that [ϕn,0]([fn])=[f][\phi_{n,0}]([f_{n}])=[f] for every n1n\geq 1. Since [f][f] is transcendental, for every n1n\geq 1, [fn][f_{n}] is transcendental. By Lemma 2.13, dimZ(fn)1\dim_{\mathbb{R}}Z^{\mathbb{R}}(f_{n})\geq 1 for every n1.n\geq 1. Our assumption implies that |λs1,,sn|([fn])Lsn|\lambda_{s_{1},\dots,s_{n}}|([f_{n}])\in L^{s_{n}}. The last paragraph of Section 2.5 shows that |λs1,,sn||\lambda_{s_{1},\dots,s_{n}}| is constant on Z(fn).Z^{\mathbb{R}}(f_{n}). As |λs1,,sn|([fn])(1,+)sn|\lambda_{s_{1},\dots,s_{n}}|([f_{n}])\in(1,+\infty)^{s_{n}}, [ϕn,0][\phi_{n,0}] is étale in a neighborhood of Z(fn).Z^{\mathbb{R}}(f_{n}). Since [ϕn,0][\phi_{n,0}] is a finite map, Vn:=[ϕn,0](Z(fn))V_{n}:=[\phi_{n,0}](Z^{\mathbb{R}}(f_{n})) is closed in d()\mathcal{M}_{d}(\mathbb{C}). Then VnV_{n} is an admissible subset of d()\mathcal{M}_{d}(\mathbb{C}). Moreover, by Lemma 2.12, Vn,n1V_{n},n\geq 1 is decreasing. By Theorem 2.11, there is N1N\geq 1 such that Vn=VNV_{n}=V_{N} for nN.n\geq N. Then for every gVNg\in V_{N}, we have ARL([g]).A\subseteq RL^{*}([g]). Since [f][FLd][f]\not\in[FL_{d}], Z([fN])Z^{\mathbb{R}}([f_{N}]) is real irreducible and dimZ([fN])1\dim_{\mathbb{R}}Z^{\mathbb{R}}([f_{N}])\geq 1, VN(d()[FLd])V_{N}\cap(\mathcal{M}_{d}(\mathbb{C})\setminus[FL_{d}]) is infinite. This contradicts to Theorem 2.15. This concludes the proof. ∎

3. An equidistribution theorem

The following equidistribution-type theorem is a reformulation of [Yua08, Theorem 3.1]. We only state it in the case where the canonical height of XX is 0, since this case often appear in the dynamical settings. Our statement is slightly stronger than [Yua08, Theorem 3.1] as our SnS_{n} may contain several Galois orbits. We follow the terminology in [Yua08].

Theorem 3.1.

Let KK be a number field and XX be a projective variety over KK. Fix an embedding of KK into \mathbb{C}. Let ¯\overline{\mathcal{L}} be a metrized line bundle on XX such that \mathcal{L} is ample and the metric is semipositive. Let μ:=deg(X)1c1(¯)dimX\mu:={\rm deg}_{\mathcal{L}}(X)^{-1}c_{1}(\overline{\mathcal{L}})^{{\rm dim}\ X}_{\mathbb{C}} be the canonical probability measure on X()X(\mathbb{C}) associated to ¯\overline{\mathcal{L}}. For n>0n\in\mathbb{Z}_{>0}, let SnS_{n} be a countable subset of X(K¯)X(\overline{K}) which is Gal(¯/K){\rm Gal}(\overline{\mathbb{Q}}/K)-invariant. For ySny\in S_{n}, given real numbers an,y0a_{n,y}\geq 0 such that ySnan,y=1\sum_{y\in S_{n}}a_{n,y}=1 and an,y=an,σya_{n,y}=a_{n,\sigma y} for all ySny\in S_{n} and σGal(¯/K)\sigma\in{\rm Gal}(\overline{\mathbb{Q}}/K). Assume that (Sn)n=1(S_{n})_{n=1}^{\infty} satisfies the following two conditions:

(1) (small) ySnan,yh¯(y)0\sum_{y\in S_{n}}a_{n,y}h_{\overline{\mathcal{L}}}(y)\to 0 as n+n\to+\infty, here h¯h_{\overline{\mathcal{L}}} is the height function associated with ¯\overline{\mathcal{L}} ;

(2) (generic) for any proper subvariety VXV\subsetneqq X of XX, ySnVan,y0\sum_{y\in S_{n}\cap V}a_{n,y}\to 0 as n+n\to+\infty.
Then the measure μn:=ySnan,yδy\mu_{n}:=\sum_{y\in S_{n}}a_{n,y}\delta_{y} converges weakly to μ\mu on X()X(\mathbb{C}) as n+n\to+\infty where δy\delta_{y} denotes the Dirac measure at the point yy, i.e., for all continuous function gg on X()X(\mathbb{C}), we have

(3.3) limnySnan,yg(y)=X()g𝑑μ.\lim\limits_{n\to\infty}\sum_{y\in S_{n}}a_{n,y}g(y)=\int_{X(\mathbb{C})}gd\mu.
Proof.

Our proof is a small modification of the one for [Yua08, Theorem 3.1].

Let dd be the dimension of XX. We say that a continuous function ff on X()X(\mathbb{C}) is smooth if there exists an embedding of X()X(\mathbb{C}) into a projection manifold YY such that ff can be extended to a smooth function on YY. As in [Zha98], by the Stone-Weierstrass theorem, continuous functions on X()X(\mathbb{C}) can be approximated uniformly by smooth functions. Then we suffice to prove (3.3) for all smooth real-valued function ff on X()X(\mathbb{C}). Fix such a function ff. Let v0v_{0} be the archimedean place of KK corresponding to the fixed embedding KK\hookrightarrow\mathbb{C}. For a real function gg on X()X(\mathbb{C}) and a metrized line bundle 𝒢¯=(𝒢,)\overline{\mathcal{G}}=(\mathcal{G},\lVert\cdot\rVert) on XX, we define the twist 𝒢¯(g):=(,)\overline{\mathcal{G}}(g):=(\mathcal{M},\lVert\cdot\rVert^{\prime}) to be the line bundle 𝒢\mathcal{G} on XX with the metric sv0=sv0eg\lVert s\rVert^{\prime}_{v_{0}}=\lVert s\rVert_{v_{0}}e^{-g} and sv=sv\lVert s\rVert^{\prime}_{v}=\lVert s\rVert_{v} for any vv0v\neq v_{0}. Let ϵ>0\epsilon>0. By the adelic Minkowski’s theorem (cf. [BG06, Appendix C]) and [Yua08, Lemma 3.3], for a fixed place ω0K\omega_{0}\in\mathcal{M}_{K} and N>0N\in\mathbb{Z}_{>0}, there exists a nonzero small section sNΓ(X,N)s_{N}\in\Gamma(X,N\mathcal{L}) such that

logsNω0c^1(¯(ϵf))d+1+O(ϵ2)(d+1)deg(X)N+o(N)=(h¯(ϵf)(X)+O(ϵ2))N+o(N)\log\lVert s_{N}\rVert^{\prime}_{\omega_{0}}\leq-\frac{\hat{c}_{1}(\overline{\mathcal{L}}(\epsilon f))^{d+1}+O(\epsilon^{2})}{(d+1){\rm deg}_{\mathcal{L}}(X)}N+o(N)=(-h_{\overline{\mathcal{L}}(\epsilon f)}(X)+O(\epsilon^{2}))N+o(N)

and logsNω0\log\lVert s_{N}\rVert^{\prime}_{\omega}\leq 0 for all ωω0\omega\neq\omega_{0}, where ω\lVert\cdot\rVert^{\prime}_{\omega} denotes the metric of N¯(ϵf)N\overline{\mathcal{L}}(\epsilon f). For a point yy, denote by y¯\overline{y} its Zariski closure. For N,n>0N,n\in\mathbb{Z}_{>0}, denote the vanishing locus of sNs_{N} by VNXV_{N}\subsetneqq X, using the condition of ¯\overline{\mathcal{L}}, we have

ySnan,yh¯(ϵf)(y)ySn,y¯VNan,ydeg(y)1(vzO(y)(N1logsN(z)v))+0\displaystyle\sum_{y\in S_{n}}a_{n,y}h_{\overline{\mathcal{L}}(\epsilon f)}(y)\geq\sum_{y\in S_{n},\overline{y}\in V_{N}}a_{n,y}{\rm deg}(y)^{-1}\left(\sum_{v}\sum_{z\in O(y)}(-N^{-1}\log\lVert s_{N}(z)\rVert_{v}^{\prime})\right)+0
\displaystyle\geq (ySn,y¯VNan,y)(h¯(ϵf)(X)+O(ϵ2)+oN(1)).\displaystyle\left(\sum_{y\in S_{n},\overline{y}\notin V_{N}}a_{n,y}\right)(h_{\overline{\mathcal{L}}(\epsilon f)}(X)+O(\epsilon^{2})+o_{N}(1)).

Let n+n\to+\infty, the generic condition (2) implies that

lim infn+ySnan,yh¯(ϵf)(y)h¯(ϵf)(X)+O(ϵ2)+oN(1).\liminf_{n\to+\infty}\sum_{y\in S_{n}}a_{n,y}h_{\overline{\mathcal{L}}(\epsilon f)}(y)\geq h_{\overline{\mathcal{L}}(\epsilon f)}(X)+O(\epsilon^{2})+o_{N}(1).

Let N+N\to+\infty, then

lim infn+ySnan,yh¯(ϵf)(y)h¯(ϵf)(X)+O(ϵ2).\liminf_{n\to+\infty}\sum_{y\in S_{n}}a_{n,y}h_{\overline{\mathcal{L}}(\epsilon f)}(y)\geq h_{\overline{\mathcal{L}}(\epsilon f)}(X)+O(\epsilon^{2}).

By the definition, it is easy to see that

ySnan,yh¯(ϵf)(y)=ySnan,yh¯(y)+ϵX()f𝑑μn\sum_{y\in S_{n}}a_{n,y}h_{\overline{\mathcal{L}}(\epsilon f)}(y)=\sum_{y\in S_{n}}a_{n,y}h_{\overline{\mathcal{L}}}(y)+\epsilon\int_{X(\mathbb{C})}fd\mu_{n}

and

h¯(ϵf)(X)=h¯(X)+ϵ1deg(X)X()fc1(¯)d+O(ϵ2).h_{\overline{\mathcal{L}}(\epsilon f)}(X)=h_{\overline{\mathcal{L}}}(X)+\epsilon\frac{1}{{\rm deg}_{\mathcal{L}}(X)}\int_{X(\mathbb{C})}fc_{1}(\overline{\mathcal{L}})_{\mathbb{C}}^{d}+O(\epsilon^{2}).

With the small condition (1), dividing ϵ\epsilon and setting ϵ0+\epsilon\to 0^{+}, we get

lim infn+X()f𝑑μn1deg(X)X()fc1(¯)d=X()f𝑑μ.\liminf_{n\to+\infty}\int_{X(\mathbb{C})}fd\mu_{n}\geq\frac{1}{{\rm deg}_{\mathcal{L}}(X)}\int_{X(\mathbb{C})}fc_{1}(\overline{\mathcal{L}})_{\mathbb{C}}^{d}=\int_{X(\mathbb{C})}fd\mu.

Replacing ff by f-f in the above inequality, we get the other direction and thus

limn+X()f𝑑μn=X()f𝑑μ.\lim_{n\to+\infty}\int_{X(\mathbb{C})}fd\mu_{n}=\int_{X(\mathbb{C})}fd\mu.

Remark 3.2.

The same idea also applies for a non-archimedean place or the algebraic case, which gives the full analogy of [Yua08, Theorem 3.1 and 3.2].

In order to check the “generic” condition in Theorem 3.1, we need the following lemma. The proof uses the ergodic theory with respect to the constructible topology (on algebraic varieties) introduced by Xie in [Xie23].

Lemma 3.3.

Let KK be a number field and XX be a projective variety over KK. Given a dominant endomorphism f:XXf:X\to X and a sequence (xn)n=1(x_{n})_{n=1}^{\infty} of periodic points in X(K¯)X(\overline{K}) under ff. Assume that (xn)n=1(x_{n})_{n=1}^{\infty} is generic in XX, i.e., there does not exist a proper Zariski closed subset ZXZ\subsetneqq X containing all xnx_{n} except for finitely many. Then for every proper subvariety VXV\subsetneqq X, we have

(3.4) #(VOf(xn))#Of(xn)0,asn+,\frac{\#(V\cap O_{f}(x_{n}))}{\#O_{f}(x_{n})}\to 0,\ \text{as}\ n\to+\infty,

where Of(xn)O_{f}(x_{n}) is the (forward) orbit of xnx_{n} under ff.

Proof.

Clearly, we suffice to show that for any subsequence (nk)k(n_{k})_{k} of (n)n=1(n)_{n=1}^{\infty}, there exists a subsubsequence (nkl)l(n_{k_{l}})_{l} such that

#(VOf(xnkl))#Of(xnkl)0,asl+.\frac{\#(V\cap O_{f}(x_{n_{k_{l}}}))}{\#O_{f}(x_{n_{k_{l}}})}\to 0,\ \text{as}\ l\to+\infty.

Given a proper subvariety VXV\subsetneqq X and fix VV. Let |X|\lvert X\rvert be XX equipped with the constructible topology (i.e. the topology of XX generated by all its Zariski closed and open subsets) and 1(|X|)\mathcal{M}^{1}(\lvert X\rvert) be the space of all probability Radon measures on |X|\lvert X\rvert with the topology of weak convergence relative to all continuous functions on |X|\lvert X\rvert. Then 1(|X|)\mathcal{M}^{1}(\lvert X\rvert) is sequentially compact (cf. [Xie23, Corollary 1.14]). For n>0n\in\mathbb{Z}_{>0}, set

mn=(#Of(xn))1zOf(xn)δz.m_{n}=(\#O_{f}(x_{n}))^{-1}\sum_{z\in O_{f}(x_{n})}\delta_{z}.

By the sequentially compactness of 1(|X|)\mathcal{M}^{1}(\lvert X\rvert), we suffice to show that for any subsequence (nk)k(n_{k})_{k} of (n)n=1(n)_{n=1}^{\infty} with mnkmask+m_{n_{k}}\to m\ \text{as}\ k\to+\infty in 1(|X|)\mathcal{M}^{1}(\lvert X\rvert) for some m1(|X|)m\in\mathcal{M}^{1}(\lvert X\rvert), we have

#(VOf(xnk))#Of(xnk)0,ask+.\frac{\#(V\cap O_{f}(x_{n_{k}}))}{\#O_{f}(x_{n_{k}})}\to 0,\ \text{as}\ k\to+\infty.

Without loss of generality, we may assume that (mn)(m_{n}) itself converges to a measure m1(|X|)m\in\mathcal{M}^{1}(\lvert X\rvert); and we suffice to show (3.4) in this case. As fmn=mnf_{*}m_{n}=m_{n}, we see that fm=mf_{*}m=m. Then according to [Xie23, Lemma 5.3], mm must be of form m=ySayδOf(y)m=\sum_{y\in S}a_{y}\delta_{O_{f}(y)}, where SS is a countable set of periodic elements in |X|\lvert X\rvert under ff, ay0a_{y}\in\mathbb{R}_{\geq 0} with ySay=1\sum_{y\in S}a_{y}=1, and δOf(y)=(#Of(y))1zOf(y)δz\delta_{O_{f}(y)}=(\#O_{f}(y))^{-1}\sum_{z\in O_{f}(y)}\delta_{z} for ySy\in S. Denote the characteristic function of V|X|V\subsetneqq\lvert X\rvert by 1V1_{V}, then 1V1_{V} is continuous with respect to the constructible topology. As mnmm_{n}\to m, we get

#(VOf(xn))#Of(xn)=1V𝑑mn1V𝑑m,asn+.\frac{\#(V\cap O_{f}(x_{n}))}{\#O_{f}(x_{n})}=\int 1_{V}dm_{n}\to\int 1_{V}dm,\ \text{as}\ n\to+\infty.

Suppose that (3.4) fails. Then there must be a ySy\in S with ay>0a_{y}>0 and VOf(y)V\cap O_{f}(y)\neq\emptyset. Denote the exact period of yy under ff by kk. Let YY be the Zariski closure of {y}\{y\}. Then Yj=0k1fj(V)Y\subseteq\cup_{j=0}^{k-1}f^{\circ j}(V), hence YY is also a proper Zariski closed subset of XX. Note that

#(YOf(xn))#Of(xn)=1Y𝑑mn1Y𝑑mayk>0,asn+.\frac{\#(Y\cap O_{f}(x_{n}))}{\#O_{f}(x_{n})}=\int 1_{Y}dm_{n}\to\int 1_{Y}dm\geq\frac{a_{y}}{k}>0,\ \text{as}\ n\to+\infty.

Hence for every sufficiently large integer n1n\gg 1, we have xnj=0fj(Y)=j=0k1fj(Y)x_{n}\in\cup_{j=0}^{\infty}f^{\circ j}(Y)=\cup_{j=0}^{k-1}f^{\circ j}(Y); but j=0k1fj(Y)\cup_{j=0}^{k-1}f^{\circ j}(Y) has dimension strictly smaller than dimX{\rm dim}\ X by the noetherian condition, contradicting the assumption that (xn)n=1(x_{n})_{n=1}^{\infty} is generic in XX. ∎

4. Proofs of Theorem 1.4 and the defined over ¯\overline{\mathbb{Q}} case of Theorem 1.3

Proof of Theorem 1.4.

Assume that ff is not exceptional. By Theorem 1.2, our assumption implies that ff is defined over ¯\overline{\mathbb{Q}} (after a conjugate over \mathbb{C}), hence over a number field KK. After replacing KK by a finite extension of KK, we may assume that both ff and f¯\overline{f} are defined over KK. Here we denote by f¯\overline{f} the rational map obtained from ff via replacing the coefficients by their complex conjugates. According to [Hug23, Theorem 9 and Lemma 11] (cf. [Zdu14]), there exists a sequence (xn)n=1(x_{n})_{n=1}^{\infty} of distinct points in Per(f){\rm Per}^{*}(f) such that

a:=limnχf(xn)>f,a:=\lim\limits_{n\to\infty}\chi_{f}(x_{n})>\mathcal{L}_{f},

where the limit exists and is finite.

Clearly, f=f¯\mathcal{L}_{f}=\mathcal{L}_{\overline{f}}. For an arbitrary xPer(f)x\in{\rm Per}(f), we have x¯Per(f¯)\overline{x}\in{\rm Per}(\overline{f}), nf(x)=nf¯(x¯)n_{f}(x)=n_{\overline{f}}(\overline{x}) and ρf(x)=ρf¯(x¯)¯\rho_{f}(x)=\overline{\rho_{\overline{f}}(\overline{x})}, hence χf(x)=χf¯(x¯)\chi_{f}(x)=\chi_{\overline{f}}(\overline{x}).

Consider the morphism F:=f×f¯:1×11×1F:=f\times\overline{f}:\mathbb{P}^{1}\times\mathbb{P}^{1}\to\mathbb{P}^{1}\times\mathbb{P}^{1} over KK. For n>0n\in\mathbb{Z}_{>0}, set pn=(xn,xn¯)Per(F)p_{n}=(x_{n},\overline{x_{n}})\in{\rm Per}(F). Let Γ\Gamma be the Zariski closure of {pn:n>0}\{p_{n}:n\in\mathbb{Z}_{>0}\} in 1×1\mathbb{P}^{1}\times\mathbb{P}^{1}. As (pn)n=1(p_{n})_{n=1}^{\infty} is pairwise distinct, by the noetherian condition, we have dimΓ1{\rm dim}\ \Gamma\geq 1. After taking a subsequence, we may assume that Γ\Gamma is irreducible and that (pn)n=1(p_{n})_{n=1}^{\infty} is generic in Γ\Gamma.

There are 2 cases: dimΓ=2\dim\Gamma=2 or 11. When dimΓ=2{\rm dim}\ \Gamma=2, then Γ=1×1\Gamma=\mathbb{P}^{1}\times\mathbb{P}^{1} and the canonical probability measure on Γ\Gamma relative to FF is μ:=μf×μf¯\mu:=\mu_{f}\times\mu_{\overline{f}}, where μf\mu_{f} and μf¯\mu_{\overline{f}} are the canonical measures on 1\mathbb{P}^{1} relative to ff and f¯\overline{f}, respectively. When dimΓ=1{\rm dim}\ \Gamma=1, by the dynamical Manin-Mumford problem for FF on 1×1\mathbb{P}^{1}\times\mathbb{P}^{1}, proved in [GTZ11], Γ\Gamma is periodic under FF. After replacing ff by fmf^{m} for some suitable m>0m\in\mathbb{Z}_{>0}, we may assume that Γ\Gamma is FF-invariant. Still denote by μ\mu the canonical probability measure on Γ\Gamma relative to FF. In all cases, let πj:Γ1\pi_{j}:\Gamma\to\mathbb{P}^{1} be the jj-th projection on Γ\Gamma for j=1,2j=1,2. Then we have

deg(π1)μ=π1μf,deg(π2)μ=π2μf¯.\deg(\pi_{1})\mu=\pi_{1}^{*}\mu_{f},\ \deg(\pi_{2})\mu=\pi_{2}^{*}\mu_{\overline{f}}.

For n>0n\in\mathbb{Z}_{>0}, set

νn=1nf(xn)[Kn:K]j=0nf(xn)1τGal(Kn/K)δFj(τ(pn)),\nu_{n}=\frac{1}{n_{f}(x_{n})[K_{n}:K]}\sum_{j=0}^{n_{f}(x_{n})-1}\sum_{\tau\in{\rm Gal}(K_{n}/K)}\delta_{F^{\circ j}(\tau(p_{n}))},

where KnK_{n} is the Galois closure of K(xn)K(x_{n}) over KK in ¯\overline{\mathbb{Q}} and K():=KK(\infty):=K.

Claim: νn\nu_{n} converges weakly to μ\mu as n+n\to+\infty.

We prove the claim using Theorem 3.1. Let \mathcal{L} be the line bundle π1𝒪1(1)π2𝒪1(1)\pi_{1}^{*}\mathcal{O}_{\mathbb{P}^{1}}(1)\otimes\pi_{2}^{*}\mathcal{O}_{\mathbb{P}^{1}}(1) on Γ\Gamma. Then F|ΓdF|_{\Gamma}^{*}\mathcal{L}\cong\mathcal{L}^{\otimes d}. By [Zha95], there exists a unique semipositive metric over \mathcal{L} making F|ΓdF|_{\Gamma}^{*}\mathcal{L}\cong\mathcal{L}^{\otimes d} an isometry; denote \mathcal{L} with this metric by ¯\overline{\mathcal{L}}. We need to check the conditions (1) and (2) in Theorem 3.1. The condition (1) is trivial, since h¯(Γ)=0h_{\overline{\mathcal{L}}}(\Gamma)=0 and the height of any periodic algebraic point relative to ¯\overline{\mathcal{L}} is zero. For the condition (2), let VV be an arbitrary proper subvariety of Γ\Gamma and fix VV. By consider the finitely many images of VV under Galois transformations, the generic condition (2) follows from Lemma 3.3. Thus the claim is true.

Let n>0n\in\mathbb{Z}_{>0}, take m>0m\in\mathbb{Z}_{>0} such that |ρf(xn)|mK\lvert\rho_{f}(x_{n})\rvert^{m}\in K by the assumption, and write l=nf(xn)l=n_{f}(x_{n}). For every τGal(¯/K)\tau\in{\rm Gal}(\overline{\mathbb{Q}}/K) and 0jl10\leq j\leq l-1, we have

det(dFl(Fj(τ(pn))))m=det(dFl(τ(pn)))m=τ(det(dFl(pn))m)\displaystyle{\rm det}(dF^{\circ l}(F^{\circ j}(\tau(p_{n}))))^{m}={\rm det}(dF^{\circ l}(\tau(p_{n})))^{m}=\tau({\rm det}(dF^{\circ l}(p_{n}))^{m})
=\displaystyle= τ(ρf(xn)mρf¯(xn¯)m)=τ(|ρf(xn)|2m)=|ρf(xn)|2m,\displaystyle\tau(\rho_{f}(x_{n})^{m}\rho_{\overline{f}}(\overline{x_{n}})^{m})=\tau(\lvert\rho_{f}(x_{n})\rvert^{2m})=\lvert\rho_{f}(x_{n})\rvert^{2m},

hence |det(dFl(Fj(τ(pn))))|=|ρf(xn)|2\lvert{\rm det}(dF^{l}(F^{\circ j}(\tau(p_{n}))))\rvert=\lvert\rho_{f}(x_{n})\rvert^{2}. Then by the definition of νn\nu_{n}, we have

log|det(dF)|𝑑νn=1llog|det(dFl)|𝑑νn\displaystyle\int{\rm log}\lvert{\rm det}(dF)\rvert d\nu_{n}=\frac{1}{l}\int{\rm log}\lvert{\rm det}(dF^{\circ l})\rvert d\nu_{n}
=\displaystyle= 1l2[Kn:K]j=0l1τGal(Kn/K)log|det(dFl(Fj(τ(pn))))|\displaystyle\frac{1}{l^{2}[K_{n}:K]}\sum_{j=0}^{l-1}\sum_{\tau\in{\rm Gal}(K_{n}/K)}{\rm log}\lvert{\rm det}(dF^{\circ l}(F^{\circ j}(\tau(p_{n}))))\rvert
=\displaystyle= 1l2[Kn:K]j=0l1τGal(Kn/K)log|ρf(xn)|2\displaystyle\frac{1}{l^{2}[K_{n}:K]}\sum_{j=0}^{l-1}\sum_{\tau\in{\rm Gal}(K_{n}/K)}{\rm log}\lvert\rho_{f}(x_{n})\rvert^{2}
=\displaystyle= 2llog|ρf(xn)|=2χf(xn).\displaystyle\frac{2}{l}{\rm log}\lvert\rho_{f}(x_{n})\rvert=2\chi_{f}(x_{n}).

For any AA\in\mathbb{R}, since the function max{log|det(dF)|,A}{\rm max}\{{\rm log}\lvert{\rm det}(dF)\rvert,A\} is continuous, we have

2a=2limnχf(xn)=limnlog|det(dF)|𝑑νn\displaystyle 2a=2\lim\limits_{n\to\infty}\chi_{f}(x_{n})=\lim\limits_{n\to\infty}\int{\rm log}\lvert{\rm det}(dF)\rvert d\nu_{n}
\displaystyle\leq limnmax{log|det(dF)|,A}𝑑νn\displaystyle\lim\limits_{n\to\infty}\int{\rm max}\{{\rm log}\lvert{\rm det}(dF)\rvert,A\}d\nu_{n}
=\displaystyle= max{log|det(dF)|,A}𝑑μ.\displaystyle\int{\rm max}\{{\rm log}\lvert{\rm det}(dF)\rvert,A\}d\mu.

Let AA\to-\infty, by the monotone convergence theorem, we have

(4.5) 2f<2alog|det(dF)|𝑑μ.2\mathcal{L}_{f}<2a\leq\int{\rm log}\lvert{\rm det}(dF)\rvert d\mu.

When dimΓ=2{\rm dim}\ \Gamma=2, it is clear that

log|det(dF)|𝑑μ=1×1log|det(d(f×f¯))|d(μf×μf¯)=f+f¯=2f,\int{\rm log}\lvert{\rm det}(dF)\rvert d\mu=\int_{\mathbb{P}^{1}\times\mathbb{P}^{1}}{\rm log}\lvert{\rm det}(d(f\times\overline{f}))\rvert d(\mu_{f}\times\mu_{\overline{f}})=\mathcal{L}_{f}+\mathcal{L}_{\overline{f}}=2\mathcal{L}_{f},

contradicting (4.5). When dimΓ=1{\rm dim}\ \Gamma=1, then

log|det(dF)|𝑑μ=Γlog|det(d(f×f¯))|𝑑μ\displaystyle\int{\rm log}\lvert{\rm det}(dF)\rvert d\mu=\int_{\Gamma}{\rm log}\lvert{\rm det}(d(f\times\overline{f}))\rvert d\mu
=\displaystyle= Γlog|det(df)π1|𝑑μ+Γlog|det(df¯)π2|𝑑μ\displaystyle\int_{\Gamma}{\rm log}\lvert{\rm det}(df)\circ\pi_{1}\rvert d\mu+\int_{\Gamma}{\rm log}\lvert{\rm det}(d\overline{f})\circ\pi_{2}\rvert d\mu
=\displaystyle= Γlog|det(df)π1|𝑑π1μfdeg(π1)+Γlog|det(df¯)π2|𝑑π2μf¯deg(π2)\displaystyle\int_{\Gamma}{\rm log}\lvert{\rm det}(df)\circ\pi_{1}\rvert d\frac{\pi_{1}^{*}\mu_{f}}{{\rm deg}(\pi_{1})}+\int_{\Gamma}{\rm log}\lvert{\rm det}(d\overline{f})\circ\pi_{2}\rvert d\frac{\pi_{2}^{*}\mu_{\overline{f}}}{{\rm deg}(\pi_{2})}
=\displaystyle= 1log|det(df)|𝑑μf+1log|det(df¯)|𝑑μf¯\displaystyle\int_{\mathbb{P}^{1}}{\rm log}\lvert{\rm det}(df)\rvert d\mu_{f}+\int_{\mathbb{P}^{1}}{\rm log}\lvert{\rm det}(d\overline{f})\rvert d\mu_{\overline{f}}
=\displaystyle= f+f¯=2f,\displaystyle\mathcal{L}_{f}+\mathcal{L}_{\overline{f}}=2\mathcal{L}_{f},

contradicting (4.5). Therefore, ff must be exceptional. We have finished the proof. ∎

Proof of Theorem 1.3 when ff is defined over ¯\overline{\mathbb{Q}}.

Assume that ff is defined over ¯\overline{\mathbb{Q}}, hence ff is defined over a number field KK. Suppose that the Theorem 1.3 does not hold for ff. Let VV be the \mathbb{Q}-span of χf(Per(f))\chi_{f}({\rm Per}^{*}(f)) in \mathbb{R}, then dimV<{\rm dim}_{\mathbb{Q}}V<\infty. We can take M>0M\in\mathbb{Z}_{>0} and x1,,xMPer(f)x_{1},\dots,x_{M}\in{\rm Per}^{*}(f) such that χf(x1),,χf(xM)\chi_{f}(x_{1}),\dots,\chi_{f}(x_{M}) generate VV over \mathbb{Q}. By enlarging KK, we may assume that |ρf(x1)|,,|ρf(xM)|K\lvert\rho_{f}(x_{1})\rvert,\dots,\lvert\rho_{f}(x_{M})\rvert\in K. Then for every z0Per(f)z_{0}\in{\rm Per}^{*}(f), χf(z0)\chi_{f}(z_{0}) is a linear combination of χf(x1),,χf(xM)\chi_{f}(x_{1}),\dots,\chi_{f}(x_{M}) over \mathbb{Q}; then it is easy to see that there exists n>0,n1,,nMn\in\mathbb{Z}_{>0},n_{1},\dots,n_{M}\in\mathbb{Z} such that

|ρf(z0)|n=|ρf(x1)|n1|ρf(xM)|nMK,\lvert\rho_{f}(z_{0})\rvert^{n}=\lvert\rho_{f}(x_{1})\rvert^{n_{1}}\cdots\lvert\rho_{f}(x_{M})\rvert^{n_{M}}\in K,

which contradicts Theorem 1.4 since ff is not exceptional by the assumption. ∎

5. Some linear algebras

5.1. Pseudo linear algebra

Let V,WV,W be two \mathbb{R}-linear spaces. A pseudo morphism f:VWf:V\to W is a pair (Vf,f)(V_{f},f) where VfV_{f} is a linear subspace of VV and f:VfWf:V_{f}\to W is an \mathbb{R}-linear map. If xVVfx\in V\setminus V_{f}, we write f(x)=f(x)=\infty. When W=W=\mathbb{R}, we say that ff is a pseudo linear function.

Denote by PHom(V,W){\rm PHom}(V,W) the set of pseudo morphisms from VV to WW. For f,gPHom(V,W)f,g\in{\rm PHom}(V,W), we define f+gf+g to be the pair (VfVg,f|VfVg+g|VfVg)(V_{f}\cap V_{g},f|_{V_{f}\cap V_{g}}+g|_{V_{f}\cap V_{g}}). Then PHom(V,W){\rm PHom}(V,W) is a commutative semigroup with ++ as the operation. We denote by 0 the pair (V,0)(V,0). We have 0+f=f0+f=f for all fPHom(V,W)f\in{\rm PHom}(V,W). For every aa\in\mathbb{R}, we define afaf to be the pair (Vf,af)(V_{f},af). We note that f+(f)=(Vf,0)f+(-f)=(V_{f},0), which is not 0 if VfVV_{f}\neq V. We have an natural embedding Hom(V,W)PHom(V,W){\rm Hom}(V,W)\hookrightarrow{\rm PHom}(V,W).

For fPHom(U,V)f\in{\rm PHom}(U,V) and gPHom(V,W)g\in{\rm PHom}(V,W), we define their composition gfg\circ f to be (Uff1(Vg),gf|Uff1(Vg))PHom(U,W)(U_{f}\cap f^{-1}(V_{g}),g\circ f|_{U_{f}\cap f^{-1}(V_{g})})\in{\rm PHom}(U,W). Observe that if f(v)=f(v)=\infty, then gf(v)=g\circ f(v)=\infty.

Fix a subset OO of VV. Denote the set of positive real numbers by +\mathbb{R}^{+}, and set 0:=+{0}\mathbb{R}_{\geq 0}:=\mathbb{R}^{+}\cup\{0\}.

Definition 5.1.

A sequence (fi)i=1(f_{i})_{i=1}^{\infty} in PHom(V,W){\rm PHom}(V,W) is said to be an OO-sequence if the following conditions are satisfied:

(i) fi(O)0{}f_{i}(O)\subseteq\mathbb{R}_{\geq 0}\cup\{\infty\} for i1i\geq 1; (ii) for every λO\lambda\in O, the set {i1:fi(λ)0}\{i\geq 1:f_{i}(\lambda)\neq 0\} is finite.
Clearly, an infinite subsequence of an OO-sequence is still an OO-sequence.

Definition 5.2.

Let (λi)i=1(\lambda_{i})_{i=1}^{\infty} be a sequence in OO and (fi)i=1(f_{i})_{i=1}^{\infty} be a sequence in PHom(V,){\rm PHom}(V,\mathbb{R}). We say that ((λi)i=1,(fi)i=1)((\lambda_{i})_{i=1}^{\infty},(f_{i})_{i=1}^{\infty}) is an upper triangle OO-system (resp. weak upper triangle OO-system) if the following conditions hold:

(i) (fi)i=1(f_{i})_{i=1}^{\infty} is an OO-sequence;

(ii) fi(λi)+f_{i}(\lambda_{i})\in\mathbb{R}^{+} (resp. fi(λi)+{})f_{i}(\lambda_{i})\in\mathbb{R}^{+}\cup\{\infty\}) for i1i\geq 1;

(iii) fj(λi)=0f_{j}(\lambda_{i})=0 for j>i1j>i\geq 1.
Clearly, an upper triangle OO-system is a weak upper triangle OO-system.

Lemma 5.3.

Let ((λi)i=1,(fi)i=1)((\lambda_{i})_{i=1}^{\infty},(f_{i})_{i=1}^{\infty}) be a weak upper triangle OO-system. Then (λi)i=1(\lambda_{i})_{i=1}^{\infty} are linearly independent over \mathbb{R}.

Proof.

Since f1(λ1)0f_{1}(\lambda_{1})\neq 0, we see that λ10\lambda_{1}\neq 0. Then we only need to show that for all l2l\geq 2, λl\lambda_{l} is not contained in span{λi:il1}{\rm span}_{\mathbb{R}}\{\lambda_{i}:i\leq l-1\}. Otherwise, λl=i=1l1aiλi\lambda_{l}=\sum_{i=1}^{l-1}a_{i}\lambda_{i} for some l2,ai,1il1l\geq 2,a_{i}\in\mathbb{R},1\leq i\leq l-1. Then fl(λl)=i=1l1aifl(λi)=0f_{l}(\lambda_{l})=\sum_{i=1}^{l-1}a_{i}f_{l}(\lambda_{i})=0, which contradicts to our assumption. ∎

Let τ:VV\tau:V\to V be an involution (i.e. τ2=id\tau^{2}=\mathrm{id}).

Lemma 5.4.

Assume that τ(O)O\tau(O)\subseteq O. Let ((λi)i=1,(fi)i=1)((\lambda_{i})_{i=1}^{\infty},(f_{i})_{i=1}^{\infty}) be an upper triangle OO-system. Then there exists a strictly increasing sequence (mi)i=1(m_{i})_{i=1}^{\infty} in >0\mathbb{Z}_{>0} such that the pair ((λmi+τ(λmi))i=1,(fmi)i=1)((\lambda_{m_{i}}+\tau(\lambda_{m_{i}}))_{i=1}^{\infty},(f_{m_{i}})_{i=1}^{\infty}) is a weak upper triangle OO^{\prime}-system, where O={λmi+τ(λmi):i>0}O^{\prime}=\{\lambda_{m_{i}}+\tau(\lambda_{m_{i}}):i\in\mathbb{Z}_{>0}\}.

Proof.

It is clear that (fi)i=1(f_{i})_{i=1}^{\infty} is also an OO^{\prime}-sequence. We construct (mi)i=1(m_{i})_{i=1}^{\infty} recursively. Set m1:=1m_{1}:=1. As τ(λ1)O\tau(\lambda_{1})\in O, we have f1(τ(λ1))0{}f_{1}(\tau(\lambda_{1}))\in\mathbb{R}_{\geq 0}\cup\{\infty\}. Since f1(λ1)+f_{1}(\lambda_{1})\in\mathbb{R}^{+}, we have

fm1(λm1+τ(λm1))=f1(λ1)+f1(τ(λ1))+{}.f_{m_{1}}(\lambda_{m_{1}}+\tau(\lambda_{m_{1}}))=f_{1}(\lambda_{1})+f_{1}(\tau(\lambda_{1}))\in\mathbb{R}^{+}\cup\{\infty\}.

Assume that we have constructed m1,,mlm_{1},\dots,m_{l} satisfying the conditions for weak upper triangle systems. Since (fi)i=1(f_{i})_{i=1}^{\infty} is an OO-system and τ(λ1),,τ(λl)O\tau(\lambda_{1}),\cdots,\tau(\lambda_{l})\in O, there exists ml+1>mlm_{l+1}>m_{l} such that fml+1(τ(λi))=0f_{m_{l+1}}(\tau(\lambda_{i}))=0 for all i=1,,li=1,\dots,l. Then for all i=1,,li=1,\dots,l, we have

fml+1(λmi+τ(λmi))=fml+1(λmi)+fml+1(τ(λmi))=0;f_{m_{l+1}}(\lambda_{m_{i}}+\tau(\lambda_{m_{i}}))=f_{m_{l+1}}(\lambda_{m_{i}})+f_{m_{l+1}}(\tau(\lambda_{m_{i}}))=0;

also,

fml+1(λml+1+τ(λml+1))=fml+1(λml+1)+fml+1(τ(λml+1))+{}f_{m_{l+1}}(\lambda_{m_{l+1}}+\tau(\lambda_{m_{l+1}}))=f_{m_{l+1}}(\lambda_{m_{l+1}})+f_{m_{l+1}}(\tau(\lambda_{m_{l+1}}))\in\mathbb{R}^{+}\cup\{\infty\}

and

fmi(λml+1+τ(λml+1))=fmi(λml+1)+fmi(τ(λml+1))0{}.f_{m_{i}}(\lambda_{m_{l+1}}+\tau(\lambda_{m_{l+1}}))=f_{m_{i}}(\lambda_{m_{l+1}})+f_{m_{i}}(\tau(\lambda_{m_{l+1}}))\in\mathbb{R}_{\geq 0}\cup\{\infty\}.

We conclude the proof. ∎

By Lemma 5.4 and Lemma 5.3 we get the following result.

Corollary 5.5.

Assume that τ(O)O\tau(O)\subseteq O. Let ((λi)i=1,(fi)i=1)((\lambda_{i})_{i=1}^{\infty},(f_{i})_{i=1}^{\infty}) be an upper triangle OO-system. Then dimspan{21(λi+τ(λi)):i1}={\rm dim}_{\mathbb{R}}{\rm span}_{\mathbb{R}}\{2^{-1}(\lambda_{i}+\tau(\lambda_{i})):i\geq 1\}=\infty.

Note that the discussion in this subsection also applies with \mathbb{R} replaced by any ordered field FF.

5.2. Linear algebra for multiplication

For every field kk of characteristic 0, denote by μk\mu_{k} the subgroup of roots of unity in kk. Denote by rog:k𝔻(k):=k/μk{\rm rog}:k^{*}\to\mathbb{D}(k):=k^{*}/\mu_{k} the quotient map. Extend rog{\rm rog} to a map rog:k𝔻(k){}{\rm rog}:k\to\mathbb{D}(k)\cup\{\infty\} by sending 0 to \infty. Here we use the notation rog{\rm rog} since it is an analogy of the classical log\log function to some extent. The embedding kk¯k\hookrightarrow\overline{k} gives a natural embedding 𝔻(k)𝔻(k¯)\mathbb{D}(k)\hookrightarrow\mathbb{D}(\overline{k}) as multiplicative abelian groups. Write 𝔻(k):=𝔻(k)\mathbb{D}(k)_{\mathbb{Q}}:=\mathbb{D}(k)\otimes_{\mathbb{Z}}\mathbb{Q}, where 𝔻(k)\mathbb{D}(k) is as a multiplicative commutative group, hence a \mathbb{Z}-module; then 𝔻(k)\mathbb{D}(k)_{\mathbb{Q}} is the subspace of 𝔻(k¯)\mathbb{D}(\overline{k}) spanned by 𝔻(k)\mathbb{D}(k) over \mathbb{Q}. Write 𝔻(k):=𝔻(k)\mathbb{D}(k)_{\mathbb{R}}:=\mathbb{D}(k)\otimes_{\mathbb{Z}}\mathbb{R}.

Let AkA\subseteq k be an integral domain with Frac(A)=k{\rm Frac}(A)=k. Define 𝔻(A):=rog(A{0})\mathbb{D}(A):={\rm rog}(A\setminus\{0\})\subseteq 𝔻(k)\mathbb{D}(k), which is a subsemigroup of 𝔻(k)\mathbb{D}(k). For every prime ideal pp of AA, the surjective projection AA/pA\to A/p induces a surjective morphism sp:𝔻(A){}s_{p}:\mathbb{D}(A)\cup\{\infty\}\to 𝔻(A/p){}\mathbb{D}(A/p)\cup\{\infty\}. In fact, we may view sps_{p} as a pseudo morphism

sp:𝔻(k)𝔻(Frac(A/p))s_{p}:\mathbb{D}(k)_{\mathbb{R}}\to\mathbb{D}({\rm Frac}(A/p))_{\mathbb{R}}

with domain Vsp:=(Ap)V_{s_{p}}:=(A\setminus p)\otimes_{\mathbb{Z}}\mathbb{R}.

5.3. Norms

Let kk be a field of characteristic 0. For every finite field extension k~\widetilde{k} over kk , denote by Nk~/k:k~kN_{\widetilde{k}/k}:\widetilde{k}\to k the norm map. We define a morphism 𝐧k:𝔻(k¯)𝔻(k)\mathbf{n}_{k}:\mathbb{D}(\overline{k})_{\mathbb{Q}}\to\mathbb{D}(k)_{\mathbb{Q}} by

𝐧k:rog(x)[l:k]1rog(Nl/k(x)),\mathbf{n}_{k}:{\rm rog}(x)\mapsto[l:k]^{-1}{\rm rog}(N_{l/k}(x)),

where ll is any finite extension over kk containing xx. We may check that 𝐧k\mathbf{n}_{k} is well defined and is \mathbb{Q}-linear. We also denote by 𝐧k\mathbf{n}_{k} its \mathbb{R}-linear extension 𝐧k:𝔻(k¯)𝔻(k)\mathbf{n}_{k}:\mathbb{D}(\overline{k})_{\mathbb{R}}\to\mathbb{D}(k)_{\mathbb{R}}. When the field kk is clear, we also write 𝐧\mathbf{n} for 𝐧k\mathbf{n}_{k}.

5.4. Valuations

Assume that KK is a number field. Denote by K\mathcal{M}_{K} the set of all places of KK. For every vKv\in\mathcal{M}_{K}, denote by v:𝔻(K)v:\mathbb{D}(K)_{\mathbb{R}}\to\mathbb{R} the \mathbb{R}-linear map given by

rog(x)log(|x|v),xK.{\rm rog}(x)\mapsto-\log(\lvert x\rvert_{v}),\ x\in K^{*}.

It is easy to check that this map is well-defined and \mathbb{R}-linear. We also denote by v:𝔻(K)v:\mathbb{D}(K)_{\mathbb{Q}}\to\mathbb{R} its restriction. For every a𝔻(K)a\in\mathbb{D}(K)_{\mathbb{R}}, the set {vK:v(a)0}\{v\in\mathcal{M}_{K}:v(a)\neq 0\} is finite.

Let SS be a finite subset of K\mathcal{M}_{K} containing all the archimedean places. Let 𝒪K,S\mathcal{O}_{K,S} be the ring of SS-integers in KK. Let 𝒪\mathcal{O} be the integral closure of 𝒪K,S\mathcal{O}_{K,S} in K¯\overline{K}. For every vKSv\in\mathcal{M}_{K}\setminus S and λ𝒪\lambda\in\mathcal{O}, we have v𝐧(λ)0v\circ\mathbf{n}(\lambda)\geq 0. Write KS={v1,v2,}\mathcal{M}_{K}\setminus S=\{v_{1},v_{2},\dots\}. Then (vi)i=1Hom(𝔻(K),)(v_{i})_{i=1}^{\infty}\subseteq{\rm Hom}(\mathbb{D}(K)_{\mathbb{R}},\mathbb{R}) is an 𝒪K,S\mathcal{O}_{K,S}-sequence and (vi𝐧)i=1Hom(𝔻(K¯),)(v_{i}\circ\mathbf{n})_{i=1}^{\infty}\subseteq{\rm Hom}(\mathbb{D}(\overline{K})_{\mathbb{R}},\mathbb{R}) is an 𝒪\mathcal{O}-sequence.

5.5. Complex conjugation and absolute value

Denote by τ:\tau:\mathbb{C}\to\mathbb{C} the complex conjugation. Then \mathbb{R} is the fixed field τ\mathbb{C}^{\tau} of τ\tau. As \mathbb{Q}-vector spaces, we have an identification 𝔻()=/{±1},rog(a)log|a|\mathbb{D}(\mathbb{R})=\mathbb{R}^{*}/\{\pm 1\}\to\mathbb{R},{\rm rog}(a)\mapsto\log\lvert a\rvert, where the latter log\log is the classical one on +\mathbb{R}^{+}. Using this identification, the absolute value on \mathbb{C} can be viewed as the norm 𝐧:𝔻()𝔻()\mathbf{n}_{\mathbb{R}}:\mathbb{D}(\mathbb{C})\to\mathbb{D}(\mathbb{R}) sending rog(x){\rm rog}(x) to 21(rog(x)+rog(τ(x)))2^{-1}({\rm rog}(x)+{\rm rog}(\tau(x))).

Let 𝐤\mathbf{k} be an algebraically closed subfield of \mathbb{C} stable under the complex conjugation. Still denote by τGal(𝐤/)\tau\in{\rm Gal}(\mathbf{k}/\mathbb{Q}) the restriction of the complex conjugation on 𝐤\mathbf{k}. Note that τ\tau is an involution. Denote by 𝐤τ\mathbf{k}^{\tau} the τ\tau-fixed subfield of 𝐤\mathbf{k}. Then the restriction of the absolute value 𝐧\mathbf{n}_{\mathbb{R}} on 𝐤\mathbf{k} is 𝐧𝐤τ:𝔻(𝐤)𝔻(𝐤τ),rog(x)21(rog(x)+rog(τ(x)))\mathbf{n}_{\mathbf{k}^{\tau}}:\mathbb{D}(\mathbf{k})\to\mathbb{D}(\mathbf{k}^{\tau}),{\rm rog}(x)\mapsto 2^{-1}({\rm rog}(x)+{\rm rog}(\tau(x))).

We shall prove the following result.

Theorem 5.6.

Assume that 𝐤\mathbf{k} is an algebraically closed field of characteristic 0. Let τGal(𝐤/)\tau\in{\rm Gal}(\mathbf{k}/\mathbb{Q}) be an element with τ2=id\tau^{2}=\mathrm{id}. If f:11f:\mathbb{P}^{1}\to\mathbb{P}^{1} is an endomorphism over 𝐤\mathbf{k} of degree at least 22 which is not PCF, then the \mathbb{Q}-subspace in 𝔻(𝐤τ)\mathbb{D}(\mathbf{k}^{\tau})_{\mathbb{Q}} spanned by {𝐧𝐤τ(rog(ρf(x))):xPer(f)(𝐤)}\{\mathbf{n}_{\mathbf{k}^{\tau}}({\rm rog}(\rho_{f}(x))):x\in{\rm Per}^{*}(f)(\mathbf{k})\} is of infinite dimension.

Take 𝐤=\mathbf{k}=\mathbb{C} and let τ\tau be the complex conjugation, then Theorem 5.6 implies Theorem 1.3 in the case that ff is a non-PCF map.

Remark 5.7.

Setting τ=id\tau={\rm id}, then from Theorem 5.6 we get the following result:

Assume that 𝐤\mathbf{k} is an algebraically closed field of characteristic 0. If f:11f:\mathbb{P}^{1}\to\mathbb{P}^{1} is an endomorphism over 𝐤\mathbf{k} of degree at least 22 which is not PCF, then the \mathbb{Q}-subspace in 𝔻(𝐤)\mathbb{D}(\mathbf{k})_{\mathbb{Q}} spanned by {rog(ρf(x)):xPer(f)(𝐤)}\{{\rm rog}(\rho_{f}(x)):x\in{\rm Per}^{*}(f)(\mathbf{k})\} is of infinite dimension.

6. Proofs of Theorem 5.6 and Theorem 1.3

6.1. Proof of Theorem 5.6: the case 𝐤=¯\mathbf{k}=\overline{\mathbb{Q}}

Let τ\tau be an element in Gal(¯/){\rm Gal}(\overline{\mathbb{Q}}/\mathbb{Q}) with τ2=\tau^{2}= id.

Denote by 𝒞f\mathcal{C}_{f} the set of critical points of ff. Since ff is not postcritically finite, there exists o𝒞fo\in\mathcal{C}_{f} such that the (forward) orbit Of(o)O_{f}(o) of oo is infinite. We fix this critical point oo. Let XX be the union of all (forward) orbits of periodic critical points of ff. Then XX is finite.

Pick a number field KK satisfying τ(K)=K\tau(K)=K and such that f,of,o and all points in XX are defined over KK.

Denote by K\mathcal{M}_{K} the set of places of KK. Let BKB\subseteq\mathcal{M}_{K} be a finite set containing all the archimedean places, satisfying τ(B)=B\tau(B)=B and such that for every vK\B,fv\in\mathcal{M}_{K}\backslash B,f has good reduction at vv. Then we have τ(𝒪K,B)=𝒪K,B\tau(\mathcal{O}_{K,B})=\mathcal{O}_{K,B}. For xPer(f)(𝐤)x\in{\rm Per}(f)(\mathbf{k}), set λ(x)=(nf(x))1rog(ρf(x))𝔻(𝐤){}\lambda(x)=(n_{f}(x))^{-1}{\rm rog}(\rho_{f}(x))\in\mathbb{D}(\mathbf{k})_{\mathbb{R}}\cup\{\infty\}. Recall that nf(x)n_{f}(x) is the exact periods of xx and ρf(x)\rho_{f}(x) is the multiplier of xx. Then for all xPer(f)(𝐤)x\in{\rm Per}^{*}(f)(\mathbf{k}), we have 𝐧K(λ(x))𝔻(𝒪K,B)\mathbf{n}_{K}(\lambda(x))\in\mathbb{D}(\mathcal{O}_{K,B}).

Denote by v\mathbb{C}_{v} the completion of the algebraically closure of KvK_{v}. Every embedding σ:𝐤v\sigma:\mathbf{k}\hookrightarrow\mathbb{C}_{v} gives a bijection σ:Per(f)(𝐤)Per(f)(v)\sigma:{\rm Per}(f)(\mathbf{k})\to{\rm Per}(f)(\mathbb{C}_{v}). Observe that for every xPer(f)(𝐤)x\in{\rm Per}(f)(\mathbf{k}), we have σ(λ(x))=λ(σ(x))\sigma(\lambda(x))=\lambda(\sigma(x)).

For every vKBv\in\mathcal{M}_{K}\setminus B and x1(v)x\in\mathbb{P}^{1}(\mathbb{C}_{v}), denote by x~1(Kv~¯)\tilde{x}\in\mathbb{P}^{1}(\overline{\widetilde{K_{v}}}) the reduction of xx in the special fiber at vv and fv:Kv~¯1Kv~¯1f_{v}:\mathbb{P}^{1}_{\overline{\widetilde{K_{v}}}}\to\mathbb{P}^{1}_{\overline{\widetilde{K_{v}}}} the reduction of ff. After enlarging BB, we may assume that o~Xv\tilde{o}\notin X_{v} where XvX_{v} is the reduction of XX in 1(Kv~)1(Kv~¯)\mathbb{P}^{1}(\widetilde{K_{v}})\subseteq\mathbb{P}^{1}(\overline{\widetilde{K_{v}}})

Observe that for every vKB,xPer(f)(𝐤)v\in\mathcal{M}_{K}\setminus B,x\in{\rm Per}(f)(\mathbf{k}) of exact period n1n\geq 1 and any embedding σ:𝐤v\sigma:\mathbf{k}\hookrightarrow\mathbb{C}_{v}, we have v(λ(x))0v(\lambda(x))\geq 0. Moreover the followings are equivalent:

(i) v(λ(x))>0v(\lambda(x))>0;

(ii) there exists an embedding σ:𝐤v\sigma:\mathbf{k}\hookrightarrow\mathbb{C}_{v} such that (fvn)(σ(x)~)=0(f^{n}_{v})^{\prime}(\widetilde{\sigma(x)})=0;

(iii) there exists an embedding σ:𝐤v,q𝒞f\sigma:\mathbf{k}\hookrightarrow\mathbb{C}_{v},q\in\mathcal{C}_{f} and m0m\geq 0, such that σ(q)~\widetilde{\sigma(q)} is periodic for fvf_{v} and σ(x)~=fvm(σ(q)~)\widetilde{\sigma(x)}=f_{v}^{m}(\widetilde{\sigma(q)}).

For vKBv\in\mathcal{M}_{K}\setminus B, denote by PvP_{v} the union of all orbits of periodic critical points of fvf_{v}. Then PvP_{v} is finite. For every vKB,qPvv\in\mathcal{M}_{K}\setminus B,q\in P_{v}, there exists a unique periodic point yPer(f)(v𝐤)y\in{\rm Per}(f)(\mathbb{C}_{v}\cap\mathbf{k}) such that y~=q\tilde{y}=q. Then there exists a unique Gal(𝐤/K){\rm Gal}(\mathbf{k}/K)-orbit O(q)O(q) in 𝐤\mathbf{k} such that for some (then every) xO(q)x\in O(q), there exists an embedding σ:𝐤v\sigma:\mathbf{k}\hookrightarrow\mathbb{C}_{v} such that σ(x)~=q\widetilde{\sigma(x)}=q (here O(q)O(q) is the orbit of qq). In particular, we have XvPvX_{v}\subseteq P_{v} and qXvO(q)=X\cup_{q\in X_{v}}O(q)=X. It follows that the set

Qv:={xPer(f)(𝐤):v(λ(x))>0}=qPvO(q)Q_{v}:=\{x\in{\rm Per}(f)(\mathbf{k}):v(\lambda(x))>0\}=\bigcup_{q\in P_{v}}O(q)

is finite. Moreover, Qv=XQ_{v}=X if and only if Pv=XvP_{v}=X_{v}.

Lemma 6.1.

The set S:={vKB:PvXv}S:=\{v\in\mathcal{M}_{K}\setminus B:P_{v}\setminus X_{v}\neq\emptyset\} is infinite.

Proof.

By [BGKT12, Lemma 4.1], there are infinitely may vKBv\in\mathcal{M}_{K}\setminus B, for which there exists n>0n\in\mathbb{Z}_{>0} such that fvn(o~)=o~f_{v}^{n}(\tilde{o})=\tilde{o}. For such vv, we have o~PvXv\tilde{o}\in P_{v}\setminus X_{v}, which concludes the proof. ∎

Lemma 6.2.

There exists a sequence (xi)i=1(x_{i})_{i=1}^{\infty} in Per(f)(𝐤){\rm Per}^{*}(f)(\mathbf{k}) and a sequence (vj)j=1(v_{j})_{j=1}^{\infty} in KB\mathcal{M}_{K}\setminus B such that vi(λ(xi))>0v_{i}(\lambda(x_{i}))>0 for i1i\geq 1 and vj(λ(xi))=0v_{j}(\lambda(x_{i}))=0 for jij\neq i. In particular, ((λ(xi))i=1,(vi)i=1)((\lambda(x_{i}))_{i=1}^{\infty},(v_{i})_{i=1}^{\infty}) is an upper triangle 𝔻(𝒪K,B)\mathbb{D}(\mathcal{O}_{K,B})-system for 𝔻(K)\mathbb{D}(K)_{\mathbb{R}}.

Proof.

We construct these two sequences recursively.

By Lemma 6.1, SS is infinite. Pick v1Sv_{1}\in S, then there exists x1Qv1Xx_{1}\in Q_{v_{1}}\setminus X\subseteq Per(f)(𝐤){\rm Per}^{*}(f)(\mathbf{k}). We have v1(λ(x1))>0v_{1}(\lambda(x_{1}))>0.

Assume that we have constructs x1,,xmPer(f)(𝐤)x_{1},\dots,x_{m}\in{\rm Per}^{*}(f)(\mathbf{k}) and v1,,vmv_{1},\dots,v_{m}\in KB\mathcal{M}_{K}\setminus B such that vj(λ(xi))0v_{j}(\lambda(x_{i}))\geq 0 and the quality holds if and only if jij\neq i. The set i=1mQviX\cup_{i=1}^{m}Q_{v_{i}}\setminus X is finite. Then there exists a finite set TmKT_{m}\subseteq\mathcal{M}_{K} such that for all xi=1mQviXx\in\cup_{i=1}^{m}Q_{v_{i}}\setminus X, and vKTmv\in\mathcal{M}_{K}\setminus T_{m}, we have v(x)=0v(x)=0. By Lemma 6.1, there exists vm+1v_{m+1}\in S({v1,,vm}Tm)S\setminus(\{v_{1},\dots,v_{m}\}\cup T_{m}). Then we have vm+1(xi)=0v_{m+1}(x_{i})=0 for i=1,,mi=1,\dots,m. Pick xm+1Qvm+1Xx_{m+1}\in Q_{v_{m+1}}\setminus X. We have vm+1(xm+1)>0v_{m+1}(x_{m+1})>0. It follows that xm+1i=1mQvix_{m+1}\notin\cup_{i=1}^{m}Q_{v_{i}}. Then vi(xm+1)=0v_{i}(x_{m+1})=0 for i=1,,mi=1,\dots,m. We conclude the proof of Lemma 6.2. ∎

Then we conclude the proof by Corollary 5.5.

6.2. Proof of Theorem 5.6: the general case

Denote by 𝒞f\mathcal{C}_{f} the set of critical points of ff. Since ff is not PCF, there is an o𝒞fo\in\mathcal{C}_{f} which is not preperiodic. We fix this critical point oo. Fix a subfield KK of 𝐤\mathbf{k} such that K/K/\mathbb{Q} is finite generated, τ(K)=K\tau(K)=K, and o,fo,f are defined over KK. Without loss of generality, we may assume that 𝐤=K¯\mathbf{k}=\overline{K}.

Take a finite generated \mathbb{Z}-subalgebra AA of KK with Frac(A)=K{\rm Frac}(A)=K and τ(A)=A\tau(A)=A. After shrinking Spec(A){\rm Spec}(A), we may assume that there exists an endomorphism fA:A1A1f_{A}:\mathbb{P}^{1}_{A}\to\mathbb{P}^{1}_{A} over AA whose restriction fK:K1K1f_{K}:\mathbb{P}^{1}_{K}\to\mathbb{P}^{1}_{K} over the generic fiber K1\mathbb{P}^{1}_{K} satisfies f=fKK𝐤f=f_{K}\otimes_{K}\mathbf{k}.

For every cSpec(A¯)(¯)c\in{\rm Spec}(A\otimes_{\mathbb{Z}}\overline{\mathbb{Q}})(\overline{\mathbb{Q}}), denote by fcf_{c} the specialization of fAf_{A} at cc, and oco_{c} the specialization of oo at cc. Then oco_{c} is a critical point of fcf_{c}. By [GX18, Lemma 3.3], there exists cSpec(A¯)(¯)c\in{\rm Spec}(A\otimes_{\mathbb{Z}}\overline{\mathbb{Q}})(\overline{\mathbb{Q}}) such that the orbit of oco_{c} is infinite. In particular, fcf_{c} is not PCF. There exists a number field LK¯L\subseteq\overline{K} such that cc is defined over LL. Denote by A1A_{1} the algebra generated by A,𝒪LA,\mathcal{O}_{L} and τ(𝒪L)\tau(\mathcal{O}_{L}); we may replace Spec(A){\rm Spec}(A) by some Zariski open set of Spec(A1){\rm Spec}(A_{1}) for which fA:A1A1f_{A}:\mathbb{P}^{1}_{A}\to\mathbb{P}^{1}_{A} is still everywhere well-defined. We may view Spec(A){\rm Spec}(A) as an 𝒪L\mathcal{O}_{L}-scheme, and pick a point cSpec(A𝒪LL)c\in{\rm Spec}(A\otimes_{\mathcal{O}_{L}}L) such that the orbit of oco_{c} is infinite. After shrinking Spec(A){\rm Spec}(A), the Zariski closure of cc in Spec(A){\rm Spec}(A) is isomorphic to Spec(𝒪L,S){\rm Spec}(\mathcal{O}_{L,S}) for a finite set of places SLS\subseteq\mathcal{M}_{L} containing all archimedean places. It corresponds to a prime ideal pp of AA.

Denote by sp:𝔻(K)𝔻(L)s_{p}:\mathbb{D}(K)_{\mathbb{R}}\dashrightarrow\mathbb{D}(L)_{\mathbb{R}} the pseudo morphism as in Section 5. We have sp(𝔻(A))𝔻(𝒪L,S){}s_{p}(\mathbb{D}(A))\subseteq\mathbb{D}(\mathcal{O}_{L,S})\cup\{\infty\}. Then for every vLSv\in\mathcal{M}_{L}\setminus S and λ𝔻(A)\lambda\in\mathbb{D}(A), we have v(λ)0{}v(\lambda)\in\mathbb{R}_{\geq 0}\cup\{\infty\}. Moreover, for every λ𝔻(A)\lambda\in\mathbb{D}(A), if sp(λ)s_{p}(\lambda)\neq\infty, then there are only finitely many vLSv\in\mathcal{M}_{L}\setminus S for which v(sp(λ))0v(s_{p}(\lambda))\neq 0.

For every yPer(f)(K¯)y\in{\rm Per}(f)(\overline{K}), denote by ycy_{c} the set of xPer(fc)(L¯)x\in{\rm Per}(f_{c})(\overline{L}) with whose image is contained in the image of yy in A1\mathbb{P}_{A}^{1}. For every yPer(f)(K¯)y\in{\rm Per}(f)(\overline{K}), ycy_{c} is finite and nonempty. On the other hand, for every xPer(fc)(L¯)x\in{\rm Per}(f_{c})(\overline{L}), the set of yPer(f)(K¯)y\in{\rm Per}(f)(\overline{K}) with xycx\in y_{c} is finite and nonempty. Moreover, if xycx\in y_{c}, then

sp(𝐧K(λ(y)))=𝐧L(λ(x)).s_{p}(\mathbf{n}_{K}(\lambda(y)))=\mathbf{n}_{L}(\lambda(x)).

Since the set of xPer(fc)(L¯)x\in{\rm Per}(f_{c})(\overline{L}) with 𝐧L(λ(x))=\mathbf{n}_{L}(\lambda(x))=\infty is finite, the set

Wc:={yPer(f)(K¯):sp(𝐧K(λ(y)))=}W_{c}:=\{y\in{\rm Per}(f)(\overline{K}):s_{p}(\mathbf{n}_{K}(\lambda(y)))=\infty\}

is also finite. Similarly Wτ(c):={yPer(f)(K¯):sτ(p)(𝐧K(λ(y)))=}W_{\tau(c)}:=\{y\in{\rm Per}(f)(\overline{K}):s_{\tau(p)}(\mathbf{n}_{K}(\lambda(y)))=\infty\} is finite.

By Lemma 6.2, there exists (yi)i=1(y_{i})_{i=1}^{\infty} in Per(fc)(L¯){\rm Per}(f_{c})(\overline{L}) and (vi)i=1(v_{i})_{i=1}^{\infty} in LS\mathcal{M}_{L}\setminus S such that ((𝐧L(λ(yi)))i=1,(vi)i=1)((\mathbf{n}_{L}(\lambda(y_{i})))_{i=1}^{\infty},(v_{i})_{i=1}^{\infty}) is an upper triangle 𝔻(𝒪L,S)\mathbb{D}(\mathcal{O}_{L,S})-system for 𝔻(L)\mathbb{D}(L)_{\mathbb{R}}. For every i>0i\in\mathbb{Z}_{>0}, there exists xiPer(f)(K¯)x_{i}\in{\rm Per}(f)(\overline{K}) such that the image of yiy_{i} is contained in the Zariski closure of the image of xix_{i} in A1\mathbb{P}_{A}^{1}. We have

sp(𝐧K(λ(xi)))=𝐧L(λ(yi)).s_{p}(\mathbf{n}_{K}(\lambda(x_{i})))=\mathbf{n}_{L}(\lambda(y_{i})).

After removing finite terms, we may assume that yiWcWτ(c)y_{i}\notin W_{c}\cup W_{\tau(c)} for all i1i\geq 1. It follows that 𝐧L(λ(yi))rog(A(pτ(p)))\mathbf{n}_{L}(\lambda(y_{i}))\in{\rm rog}(A\setminus(p\cup\tau(p))) for i1i\geq 1. Observe that (visp)i=1(v_{i}\circ s_{p})_{i=1}^{\infty} is a rog(A(pτ(p))){\rm rog}(A\setminus(p\cup\tau(p)))-sequence. It follows that ((𝐧L(λ(yi)))i=1,(visp)i=1)((\mathbf{n}_{L}(\lambda(y_{i})))_{i=1}^{\infty},(v_{i}\circ s_{p})_{i=1}^{\infty}) is an upper triangle rog(A(pτ(p))){\rm rog}(A\setminus(p\cup\tau(p)))-system for 𝔻(K)\mathbb{D}(K)_{\mathbb{R}}. Since rog(A(pτ(p))){\rm rog}(A\setminus(p\cup\tau(p))) is invariant under τ\tau, we conclude the proof by Corollary 5.5.

6.3. Proof of Theorem 1.3

There are two cases:

1. The case ff is PCF. In this case according to [DH93], PCF maps are defined over ¯\overline{\mathbb{Q}} in the moduli space d\mathcal{M}_{d} of rational maps of degree dd, except for the family of flexible Lattès maps. So after a conjugacy by an elements in PGL2(){\rm PGL}_{2}(\mathbb{C}), ff is defined over ¯\overline{\mathbb{Q}}, and Theorem 1.3 was already proved in the end of Section 4.

2. The case ff is not PCF. Then Theorem 1.3 is a consequence of Theorem 5.6. This finishes the proof of Theorem 1.3.

7. Proofs of the Applications

7.1. Proof of Theorem 1.10

Without loss of generality, we may assume that kk is of finite transcendence degree over .\mathbb{Q}. Fix an embedding of kk into \mathbb{C}. We view ff as an endomorphism on XX defined over \mathbb{C}. According to [Xie22, Theorem 3.34], we may assume that all fj:11f_{j}:\mathbb{P}^{1}\to\mathbb{P}^{1} has degree at least 22 for 1jN1\leq j\leq N.

Assume first that all fjf_{j} are not exceptional, 1jN1\leq j\leq N. Corollary 1.5 implies that we can take xjPer(fj)()x_{j}\in{\rm Per}^{*}(f_{j})(\mathbb{C}) for 1jN1\leq j\leq N such that ρf1(x1),,ρfN(xN)\rho_{f_{1}}(x_{1}),\cdots,\rho_{f_{N}}(x_{N}) are multiplicatively independent in \mathbb{C}. After replacing ff by an iterate, we may assume that fj(xj)=xjf_{j}(x_{j})=x_{j} for 1jN1\leq j\leq N, and the multipliers (ρfj(xj)=fj(xj))j=1N(\rho_{f_{j}}(x_{j})=f_{j}^{\prime}(x_{j}))_{j=1}^{N} are still multiplicatively independent. Denote x=(x1,,xN)X(k)x=(x_{1},\dots,x_{N})\in X(k). Then xx is a fixed point of ff (smooth in the fixed locus of ff) such that the eigenvalues of df|xdf|_{x} are nonzero and multiplicatively independent. Then the conclusion follows from [ABR11].

Assume that all fjf_{j} are exceptional, 1jN1\leq j\leq N. This case is easy, and we just refer to the proof in the first several paragraphs of [Xie22, Section 9.3].

We may assume that 0sN0\leq s\leq N such that f1,,fsf_{1},\cdots,f_{s} are not exceptional and fs+1,,fNf_{s+1},\cdots,f_{N} are exceptional. Let l(f)=min{s,Ns}0l(f)={\rm min}\{s,N-s\}\geq 0. Then we have done in the case l(f)=0l(f)=0. Then an induction on l(f)l(f) will prove this corollary, as shown in the last several paragraphs of [Xie22, Section 9.3].

7.2. Proof of Theorem 1.12

Using the terminology and notations in Section 5, it is clear that (2)(2) and (3)(3) is equivalent to the following (2)(2)^{\prime} and (3)(3)^{\prime}, respectively.

(2)(2)^{\prime} ρf(x)¯\rho_{f}(x)\in\overline{\mathbb{Q}} for all xPer(f)()x\in{\rm Per}(f)(\mathbb{C}) and the \mathbb{Q}-subspace of 𝔻()\mathbb{D}(\mathbb{Q})_{\mathbb{Q}} generated by 𝐧(rog(ρf(x)))\mathbf{n}_{\mathbb{Q}}({\rm rog}(\rho_{f}(x))) for xPer(f)()x\in{\rm Per}^{*}(f)(\mathbb{C}) is of finite dimension over \mathbb{Q}.

(3)(3)^{\prime} |ρf(x)|¯\lvert\rho_{f}(x)\rvert\in\overline{\mathbb{Q}} for all xPer(f)()x\in{\rm Per}(f)(\mathbb{C}) and the \mathbb{Q}-subspace of 𝔻()\mathbb{D}(\mathbb{Q})_{\mathbb{Q}} generated by 𝐧(rog(|ρf(x)|))\mathbf{n}_{\mathbb{Q}}({\rm rog}(\lvert\rho_{f}(x)\lvert)) for xPer(f)()x\in{\rm Per}^{*}(f)(\mathbb{C}) is of finite dimension over \mathbb{Q}.

Now we prove that (1),(2),(3)(1),(2)^{\prime},(3)^{\prime} are equivalent.
(1) \Rightarrow (2)(2)^{\prime} and (3)(3)^{\prime}:

Suppose that ff is PCF. By [DH93], PCF maps are defined over ¯\overline{\mathbb{Q}} in d\mathcal{M}_{d}, except for the family of flexible Lattès maps. If ff is flexible Lattès, then according to [Mil06, Lemma 5.6], ρf(x)\rho_{f}(x)\in\mathbb{Z} for all xPer(f)()x\in{\rm Per}(f)(\mathbb{C}). If ff is defined over ¯\overline{\mathbb{Q}}, then clearly ρf(x)¯\rho_{f}(x)\in\overline{\mathbb{Q}} for all xPer(f)()x\in{\rm Per}(f)(\mathbb{C}). Thus, we always have ρf(x),|ρf(x)|¯\rho_{f}(x),\lvert\rho_{f}(x)\rvert\in\overline{\mathbb{Q}} for all xPer(f)()x\in{\rm Per}(f)(\mathbb{C}).

Suppose that (2)(2)^{\prime} is false, then

dimspan{𝐧(rog(ρf(x))):xPer(f)()}=.{\rm dim}_{\mathbb{Q}}{\rm span}_{\mathbb{Q}}\{\mathbf{n}_{\mathbb{Q}}({\rm rog}(\rho_{f}(x))):x\in{\rm Per}^{*}(f)(\mathbb{C})\}=\infty.

By [Mil06, Corollary 3.9], ff cannot be a flexible Lattès map, hence ff is defined over ¯\overline{\mathbb{Q}}, and over a number field KK. We use the notation and ideas in the case of Section 6.1 where τ=Id\tau={\rm Id}. Let BKB\subseteq\mathcal{M}_{K} be a finite set containing all the archimedean places such that for every vKB,fv\in\mathcal{M}_{K}\setminus B,f has good reduction at vv. For every vKBv\in\mathcal{M}_{K}\setminus B, the reduction fvf_{v} are still PCF and its critical orbits from those of ff. Then as in Section 6.1, it is easy to see that the set

𝒲:={xPer(f)():v(𝐧K(λ(x)))=0,vKB}\mathcal{W}:=\{x\in{\rm Per}^{*}(f)(\mathbb{C}):v(\mathbf{n}_{K}(\lambda(x)))=0,\forall v\in\mathcal{M}_{K}\setminus B\}

is co-finite in Per(f)(){\rm Per}^{*}(f)(\mathbb{C}). It is well-known that rank(𝒪K,B×)=#B1<{\rm rank}(\mathcal{O}_{K,B}^{\times})=\#B-1<\infty (cf. [Nar04, Theorem 3.12]). Note that 𝐧K(rog(ρf(x)))𝔻(𝒪K,B)\mathbf{n}_{K}({\rm rog}(\rho_{f}(x)))\in\mathbb{D}(\mathcal{O}_{K,B}) for all xPer(f)()x\in{\rm Per}^{*}(f)(\mathbb{C}). Then we deduce that

dimspan{𝐧K(rog(ρf(x))):xPer(f)()}<,{\rm dim}_{\mathbb{Q}}{\rm span}_{\mathbb{Q}}\{\mathbf{n}_{K}({\rm rog}(\rho_{f}(x))):x\in{\rm Per}^{*}(f)(\mathbb{C})\}<\infty,

which implies dimspan{𝐧(rog(ρf(x))):xPer(f)()}<,{\rm dim}_{\mathbb{Q}}{\rm span}_{\mathbb{Q}}\{\mathbf{n}_{\mathbb{Q}}({\rm rog}(\rho_{f}(x))):x\in{\rm Per}^{*}(f)(\mathbb{C})\}<\infty, contradicting the assumption. Thus (2)(2)^{\prime} must hold.

(3)(3)^{\prime} follows similar to the above paragraph, corresponding to the case where τ\tau is the complex conjugate of Section 6.1.
(2)(2)^{\prime} \Rightarrow (1):

Suppose that ff is not PCF. In particularly, ff is not flexible Lattès. Denote by ZZ the set of conjugacy classes [g]d()[g]\in\mathcal{M}_{d}(\mathbb{C}) such that ff and gg have the same multiplier spectrum. By [Sil98, Theorem 4.5] and (2)(2)^{\prime}, ZZ is Zariski closed in d()\mathcal{M}_{d}(\mathbb{C}) and it is defined over ¯\overline{\mathbb{Q}}. By [McM87, Corollary 2.3], ZZ consists of finitely many points and possibly a curve of flexible Lattès maps. Since ff is not flexible Latteś, then we may assume that ff is defined over ¯\overline{\mathbb{Q}}, hence over a number field KK. By the argument in Section 6.1 of the case τ=Id\tau={\rm Id}, we have

dimspan{𝐧K(rog(ρf(x))):xPer(f)(),𝐧K(rog(ρf(x)))𝔻(𝒪K,B)}=,{\rm dim}_{\mathbb{Q}}{\rm span}_{\mathbb{Q}}\{\mathbf{n}_{K}({\rm rog}(\rho_{f}(x))):x\in{\rm Per}^{*}(f)(\mathbb{C}),\mathbf{n}_{K}({\rm rog}(\rho_{f}(x)))\in\mathbb{D}(\mathcal{O}_{K,B})\}=\infty,

where BKB\subseteq\mathcal{M}_{K} is a finite set containing all the archimedean places such that for every vKB,fv\in\mathcal{M}_{K}\setminus B,f has good reduction at vv. After enlarging BB, we may assume that BB is invariant under every σGal(K/)\sigma\in{\rm Gal}(K/\mathbb{Q}). Indeed, a small modification of the proof of Lemma 6.2 shows that there exists a sequence (xi)i=1(x_{i})_{i=1}^{\infty} in Per(f)(𝐤){\rm Per}^{*}(f)(\mathbf{k}) and a sequence (vj)j=1(v_{j})_{j=1}^{\infty} in KB\mathcal{M}_{K}\setminus B satisfy the following conditions:

vi(𝐧K(λ(xi)))>0 for all i1;\displaystyle v_{i}(\mathbf{n}_{K}(\lambda(x_{i})))>0\text{ for all }i\geq 1;
σ(vj)(𝐧K(λ(xi)))=0 for all ij and σGal(K/).\displaystyle\sigma(v_{j})(\mathbf{n}_{K}(\lambda(x_{i})))=0\text{ for all }i\neq j\text{ and }\sigma\in{\rm Gal}(K/\mathbb{Q}).

For i1i\geq 1, let pip_{i} be the prime number below viv_{i}, let B~\widetilde{B} be the restriction of BB to \mathcal{M}_{\mathbb{Q}}. Then it is easy to see that the pair ((𝐧(λ(xi)))i=1,(vpi)i=1)((\mathbf{n}_{\mathbb{Q}}(\lambda(x_{i})))_{i=1}^{\infty},(v_{p_{i}})_{i=1}^{\infty}) is an upper triangle 𝔻(,B~)\mathbb{D}(\mathcal{M}_{\mathbb{Q},\widetilde{B}})-system for 𝔻()\mathbb{D}(\mathbb{Q})_{\mathbb{Q}}. By Corollary 5.5, this contradicts (2)(2)^{\prime}. Thus, ff must be PCF.
(3)(3)^{\prime} \Rightarrow (1):

Assume that ff is not PCF. We use the notation and ideas in the case of Section 6.2 with τ\tau the complex conjugate. As in the proof of Section 6.2, we get a number field LL and a finite set SLS\subseteq\mathcal{M}_{L}. We may assume that SS is invariant under every σGal(L/)\sigma\in{\rm Gal}(L/\mathbb{Q}). By the argument in (2)(2)^{\prime} \Rightarrow (1), there exists a pair ((𝐧L(λ(yi)))i=1,(vi)i=1)((\mathbf{n}_{L}(\lambda(y_{i})))_{i=1}^{\infty},(v_{i})_{i=1}^{\infty}) satisfy the following conditions:

vi(𝐧L(λ(xi)))>0 for all i1;\displaystyle v_{i}(\mathbf{n}_{L}(\lambda(x_{i})))>0\text{ for all }i\geq 1;
σ(vj)(𝐧L(λ(xi)))=0 for all ij and σGal(L/).\displaystyle\sigma(v_{j})(\mathbf{n}_{L}(\lambda(x_{i})))=0\text{ for all }i\neq j\text{ and }\sigma\in{\rm Gal}(L/\mathbb{Q}).

Then we can deduce a contradiction similar to the proof of (2)(2)^{\prime} \Rightarrow (1), hence ff is PCF.

References

  • [ABR11] Ekaterina Amerik, Fedor Bogomolov, and Marat Rovinsky. Remarks on endomorphisms and rational points. Compositio Math., 147:1819–1842, 2011.
  • [AC08] Ekaterina Amerik and Frédéric Campana. Fibrations méromorphes sur certaines variétés à fibré canonique trivial. Pure Appl. Math. Q., 4(2, part 1):509–545, 2008.
  • [BG06] Enrico Bombieri and Walter Gubler. Heights in Diophantine geometry, volume 4 of New Mathematical Monographs. Cambridge University Press, Cambridge, 2006.
  • [BGHR22] X. Buff, T. Gauthier, V. Huguin, and J. Raissy. Rational maps with integer multipliers. arXiv:2212.03661, 2022.
  • [BGKT12] Robert L. Benedetto, Dragos Ghioca, Pär Kurlberg, and Thomas J. Tucker. A case of the dynamical Mordell-Lang conjecture. Math. Ann., 352(1):1–26, 2012. With an appendix by Umberto Zannier.
  • [BLR90] Siegfried Bosch, Werner Lütkebohmert, and Michel Raynaud. Néron models, volume 21 of Ergebnisse der Mathematik und ihrer Grenzgebiete (3) [Results in Mathematics and Related Areas (3)]. Springer-Verlag, Berlin, 1990.
  • [DH93] Adrien Douady and John H. Hubbard. A proof of Thurston’s topological characterization of rational functions. Acta Math., 171(2):263–297, 1993.
  • [GNY18] Dragos Ghioca, Khoa D. Nguyen, and Hexi Ye. The dynamical Manin-Mumford conjecture and the dynamical Bogomolov conjecture for endomorphisms of (1)n(\mathbb{P}^{1})^{n}. Compos. Math., 154(7):1441–1472, 2018.
  • [GTZ11] Dragos Ghioca, Thomas J. Tucker, and Shou-Wu Zhang. Towards a dynamical Manin-Mumford conjecture. Int. Math. Res. Not. IMRN, 22:5109–5122, 2011.
  • [GX18] Dragos Ghioca and Junyi Xie. The Dynamical Mordell–Lang Conjecture for Skew–Linear Self-Maps. Appendix by Michael Wibmer. International Mathematics Research Notices, page rny211, 2018.
  • [Hug23] Valentin Huguin. Rational maps with rational multipliers. J. Éc. polytech. Math., 10:591–599, 2023.
  • [JX23a] Zhuchao Ji and Junyi Xie. DAO for curves. arXiv preprint arXiv:2302.02583, 2023.
  • [JX23b] Zhuchao Ji and Junyi Xie. Homoclinic orbits, multiplier spectrum and rigidity theorems in complex dynamics. Forum Math. Pi, 11:Paper No. e11, 37, 2023.
  • [Lev14] A. Levy. Aim workshop postcritically finite maps in complex and arithmetic dynamics, 2014.
  • [McM87] Curt McMullen. Families of rational maps and iterative root-finding algorithms. Ann. of Math. (2), 125(3):467–493, 1987.
  • [Mil06] John Milnor. On Lattès maps. Dynamics on the Riemann Sphere: A Bodil Branner Festschrift, page 9, 2006.
  • [MS14] Alice Medvedev and Thomas Scanlon. Invariant varieties for polynomial dynamical systems. Ann. of Math. (2), 179(1):81–177, 2014.
  • [Nar04] Władysł aw Narkiewicz. Elementary and analytic theory of algebraic numbers. Springer Monographs in Mathematics. Springer-Verlag, Berlin, third edition, 2004.
  • [Pak23] Fedor Pakovich. Invariant curves for endomorphisms of 1×1\mathbb{P}^{1}\times\mathbb{P}^{1}. Math. Ann., 385(1-2):259–307, 2023.
  • [Poo17] Bjorn Poonen. Rational points on varieties, volume 186 of Graduate Studies in Mathematics. American Mathematical Society, Providence, RI, 2017.
  • [Sil98] Joseph H. Silverman. The space of rational maps on 1\mathbb{P}^{1}. Duke Math. J., 94(1):41–77, 1998.
  • [Sil07] Joseph H. Silverman. The arithmetic of dynamical systems, volume 241 of Graduate Texts in Mathematics. Springer-Verlag, New York, 2007.
  • [Sil12] Joseph H. Silverman. Moduli spaces and arithmetic dynamics, volume 30 of CRM Monograph Series. American Mathematical Society, Providence, RI, 2012.
  • [Tuc14] T. Tucker. Problem 6 in the problem list of the aim workshop postcritically finite maps in complex and arithmetic dynamics, 2014.
  • [Xie17] Junyi Xie. The existence of Zariski dense orbits for polynomial endomorphisms of the affine plane. Compos. Math., 153(8):1658–1672, 2017.
  • [Xie22] Junyi Xie. The existence of Zariski dense orbits for endomorphisms of projective surfaces (with an appendix in collaboration with T. Tucker). J. Amer. Math. Soc., 2022. published online.
  • [Xie23] Junyi Xie. Remarks on algebraic dynamics in positive characteristic. J. Reine Angew. Math., 797:117–153, 2023.
  • [XY23] Junyi Xie and Xinyi Yuan. Partial heights and the geometric Bombieri-Lang conjecture. arXiv:2305.14789, 2023.
  • [Yua08] Xinyi Yuan. Big line bundles over arithmetic varieties. Invent. Math., 173(3):603–649, 2008.
  • [Zdu14] Anna Zdunik. Characteristic exponents of rational functions. Bulletin of the Polish Academy of Sciences. Mathematics, 62(3), 2014.
  • [Zha95] Shou-Wu Zhang. Small points and adelic metrics. J. Algebraic Geom., 4(2):281–300, 1995.
  • [Zha98] Shou-Wu Zhang. Equidistribution of small points on abelian varieties. Ann.of Math. (2), 147(1998), 147:159–165, 1998.